DIPLOMARBEIT

Effects of the atypical antipsychotic clozapine on recombinantly expressed

GABAA subtypes zur Erlangung des akademischen Grades Dr. med. univ. ausgefuehrt am Zentrum fuer Hirnforschung unter der Anleitung von Prof. Dr. Margot Ernst, PhD

Luca Leandro Silva Pita

05/2018 Contents

0.1 Acknowledgements ...... i 0.2 Abbreviations ...... ii 0.3 Abstract ...... iv 0.4 Abstract [german] ...... v

1 Introduction 1 1.1 GABA Metabolism ...... 2 1.2 GABA Transport ...... 3

1.3 GABAA Receptors ...... 5 1.3.1 Structure and Distribution ...... 5 1.3.2 Drug-receptor Interactions ...... 12

1.3.3 Endogenous GABAA Receptor Ligands ...... 13 1.3.3.1 GABA ...... 13 1.3.3.2 Other Endogenous Ligands ...... 14

1.3.4 Exogenous GABAA Receptor Ligands ...... 15 1.3.4.1 ...... 15 1.3.4.2 Anesthetics ...... 17

1.3.4.3 Other GABAA Receptor Ligands ...... 20

1.4 GABAB Receptors ...... 22 1.5 Schizophrenia ...... 25 1.5.1 Epidemiology ...... 26 1.5.2 Pathomechanism ...... 26 1.5.3 Pharmacological Management ...... 27 1.6 Clozapine ...... 29

2 Materials and Methods 32 2.1 Two-Electrode Voltage Clamp ...... 32 2.1.1 Xaenopus laevis oocytes ...... 32 2.1.2 Preparation of oocytes ...... 32 2.1.3 Electrophysiological recordings ...... 34

3 Results 36

3.1 α1β2 ...... 37

3.2 α2β3 ...... 38

3.3 α1β2γ2 ...... 39

3.4 α2β3γ1 ...... 40

3.5 α2β3γ2 ...... 41 3.6 Summary ...... 42

4 Discussion 44 4.1 Exploring a possible intra-subunit binding pocket ...... 44 4.2 Agranulocytosis ...... 49

5 Summary and Outlook 50

List of tables 51

List of figures 53

Bibliography 55 0.1 Acknowledgements

I wish to thank, first and foremost, Margot Ernst for the generous integration into her research group. Her encouraging way to allow for sovereign pursuit of emerging interests while providing a reliable framework with continuous, knowledgeable and heartfelt support promotes the formation of new ideas. I feel privileged having worked with someone who radiates true curiousity and dedication about her field. Moreover I want to share the credit of this work with everyone I encoun- tered during my time at the CBR - thank you for interesting conversations, fun times and emerging friendships.

i 0.2 Abbreviations

2AG ...... 2-arachidonylglycerol

THDOC ...... 3 α5α-tetrahydrodeoxycorticosterone

DHP ...... 5 α-dihydroprogesterone

AMPA ...... α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid

GABA ...... γ-aminobutyric acid

BBB ...... blood brain barrier

BZD ...... benzodiazepines

CNS ...... central nervous system

CLZ ...... clozapine

D2 ...... dopamine D2 receptor

DZP ......

DMSO ...... dimethylsulfoxide

EC ...... effective concentration

ECD ...... extracellular domain

FMZ ...... flumazenil

GPCR ...... G -coupled receptor

GAT ...... GABA transporter

GAD ...... glutamate decarboxylase

GlyR ......

ii ICD ...... intracellular domain

MDD ...... major depressive disorder mRNA ...... messenger ribonucleic acid

NMDAR ...... N-methyl-D-aspartate receptor

PKA ...... protein kinase A

RT-PCR ...... real-time polymerase chain reaction

SLC6 ...... solute carrier family 6

TMD ...... transmembrane domain

TEV ...... two-electrode voltage clamp

VFTD ...... venus fly-trap domain

VGAT ...... vesicular GABA transporter

VGCC ...... voltage-gated calcium channel

WHO ...... World Health Organization

XR ...... xaenopus ringer

iii 0.3 Abstract

Clozapine (CLZ) is a tetracyclic compound with outstanding clinical ef- fectiveness in the treatment of psychotic symptoms even in patients resis- tant to high doses of typical neuroleptics [Naber and Lambert, 2009,Attard and Taylor, 2012]. While CLZ has been shown to interact with multiple targets [Ashby and Wang, 1996, Michel and Trudeau, 2000, Korpi et al.,

1995, Khokhar et al., 2018], its action at subtypes of the GABAA receptor family presents a unique feature when compared to the typical antipsychotic haloperidol [Korpi et al., 1995]. Treatment with CLZ is limited by its serious and potentially fatal side effects, which include increased seizure suscepti- bility and agranulocytosis [Pacia and Devinsky, 1994, Alvir et al., 1993].

As GABAergic interneurons form an undisputed core component of cor- ticolimbic circuity, defects in GABA transmission have become a central element in a number of models for psychiatric disease such as schizophrenia and bipolar disorder [Benes and Berretta, 2001].

In this work we studied CLZ’s effects on five different recombinantly expressed GABAA receptor subtypes in electrophysiological experiments on Xenopus laevis oocytes. We show extensive negative allosteric modulation of GABA elicted currents at micromolar CLZ concentrations in all studied subtypes. Our results confirm that the existence of a γ subunit is no premise for CLZ sensitivity. The maximum mean inhibition at 100umol CLZ is significantly higher in α2 containing assemblies (α2β3γ2, α2β3γ1 and α2β3) compared to α1 containing assemblies (α1β2 and α1β2γ2).

In light of hints gained from structural models of analog , we explore a tentative intra-subunit binding site candidate for CLZ and related molecules. Our findings prompt for a deeper examination of a possible link between some of CLZs clinical effects and its negative modulatory action on

GABAA receptors.

iv 0.4 Abstract [german]

Clozapin (CLZ) ist ein tetracyclischer Wirkstoff mit herausragender klinis- cher Wirksamkeit in der Behandlung psychotischer Symptome in Patienten, die nicht auf typische Neuroleptika ansprechen [Attard and Taylor, 2012]. Waehrend der Wirkmechanismus von CLZ unzureichend geklaert ist, kon- nten Interaktionen mit verschiedenen Zielmolekuelen experimentell gezeigt werden [Ashby and Wang, 1996, Michel and Trudeau, 2000, Korpi et al.,

1995]. Die Beeinflussung von Stroemen in Subtypen der GABAA Rezeptor- familie stellt eine einzigartige Eigenschaft im Vergleich zum typischen Neu- roleptikum Haloperidol dar [Korpi et al., 1995]. Die Behandlung mit CLZ ist begrenzt durch ihre schweren und potentiell toelichen Nebenwirkungen [Pa- cia and Devinsky, 1994, Alvir et al., 1993]. In dieser Arbeit wurde die Wirkung von CLZ auf fuenf verschiedene rekombinant exprimierte GABAA Rezeptor Subtypen in elektrophysiologis- chen Experimenten an Xaenopus laevis Oozyten untersucht. In allen un- tersuchten Subtypen zeigen wir eine ausgepraegte Reduktion von GABA induzierten Stroemen durch mikromolare CLZ Konzentrationen. Die Ergeb- nisse zeigen, dass die γ Untereinheit keine Vorraussetzung fuer CLZ Sensitiv- itaet ist. Die maximale mittlere Inhibition bei 100umol CLZ ist signifikant hoeher in α2 beinhaltenden Subtypen (α2β3γ2, α2β3γ1 und α2β3) verglichen mit α1 beinhaltenden Subtypen (α1β2 und α1β2γ2). Im Licht der Erkenntnisse aus Vergleichen mit Strukturmodellen analoger Proteine schlagen wir eine intra-subunit Bindestelle als Kandidat fuer die Vermittlung der negativ modulatorischen Effekte von CLZ vor. Die Ergeb- nisse regen zu einer tieferen Auseinandersetzung mit dem potentiellen Zusam- menhang zwischen negativ modulatorischen Effekten auf GABAA Rezep- toren und einigen von CLZs klinischen Effekten an.

v 1. Introduction

The human CNS is a network comprised of approximately 100 billion neu- ral cells connected through a large number of synaptic interfaces. These synapses allow for phasic signal transmission between cells through neuro- transmitter release. trigger excitatory (depolarizing) or inhibitory (hyperpolarizing) changes in the postsynaptic cell as a result of interactions with target molecules (receptors). release and action also occurs at extrasynaptic locations, leading to an overall change in the system’s excitability; a principle referred to as tonic signaling. Once the net depolarization of a neural cell reaches a certain threshold, an action potential is triggered, ultimately enabling subsequent signal transmission.

In healthy brains, excitation and inhibition are finely attuned, con- tributing to an environment promotive for information flow and cellular health [Wolfe et al., 2010,Schonfeld-Dado and Segal, 2009]. A disruption of this equilibrium can be observed in a multitude of neuropsychiatric disor- ders, including schizophrenia and epilepsy [Schnitzler and Gross, 2005].

While glutamic acid constitutes the principal excitatory transmitter, γ- aminobutyric acid (GABA) and glycine carry out inhibitory neurotrans- mission in mammals. GABA transmission occurs at an estimated 30% of mammalian synapses, making it a principle signalling mechanism. Its ac- tion goes beyond simple inhibition, as GABA can also exhibit excitatory, shunting and modulatory effects [Ernst et al., 2018]. Exemplary, GABAer- gic interneurons are involved in the integration of synaptic inputs through

1 oscillatory modulation of membrane potentials [Schnitzler and Gross, 2005]. The early hypothesized purely inhibitory nature of GABA was first chal- lenged by observation of depolarizing effects in immature neurons [Ben-Ari et al., 1989]. Later these effects were found to be present and important even in the adult CNS [Spitzer, 2010].

Besides the GABA systems fundamental physiological roles, GABA re- ceptors constitute molecular targets of essential and widely employed drugs such as benzodiazepines and anesthetics. Research into the GABA sys- tem has become a fundamental focus, revealing complex arrangements het- erogeneous in structure and function. Findings suggest major parts in the formation of disease [Benes and Berretta, 2001, Brickley and Mody, 2012] and promising potential for novel pharmacological intervention is dis- cussed [Rudolph and Mohler, 2014, Charych et al., 2009].

This work investigates the role of a subset of GABAA receptor family members as molecular targets of the established atypical antipsychotic drug clozapine.

1.1 GABA Metabolism

GABA is a non proteinogenic amino acid formed by decarboxylation of glutamic acid. The reaction uses peridoxal phosphate and is catalyzed by glutamate decarboxylase (GAD).

Figure 1.1: Simplified decarboxylation of glutamic acid yielding GABA

2 In mammals, two variants of GAD (GAD67 and GAD65) are encoded in seperate . The transcription of GAD67 starts during early em- bryonic development and shows an even intracellular distribution. While experiments with knockout mice reveal the importance of both isoforms during early developmental stages, deficits are more severe in mice lacking GAD67 [Kakizaki et al., 2015]. Initial transcription of GAD65 takes place during later developmental stages and is localized primarily in nerve end- ings. Both forms of GAD appear to be expressed solely in cells that use GABA for synaptic signal transmission [Olsen RW, 1999]. Enzymatic breakdown of GABA exclusively takes place intracellularly under the presence of α–ketoglutarate and is catalyzed by GABA transami- nase. During this reaction, α–ketoglutarate is converted into glutamic acid.

Figure 1.2: Simplified transamination of GABA, yielding succinic semialdehyde

This closed-loop reaction process is referred to as the GABA shunt, re- sponsible for synthesis, conservation and metabolism of GABA [Olsen RW, 1999].

1.2 GABA Transport

As breakdown of GABA is confined to intracellular locations, clearance from the synaptic cleft relies on a group of functional proteins termed GABA transporters (GAT). GAT belong to a family of neurotransmitter:sodium symporters known as the solute carrier 6 family (SLC6) in humans [Scimemi, 2014]. The SLC6 consists of 20 members, of which six have been identified

3 Figure 1.3: GABA transporter stoichiometry

to transport GABA (GAT1, GAT2, GAT3, BGT1, CT1 and TauT). While all six accept GABA as a substrate and rely on the presence of chloride and sodium, substrate specificity and distribution differs among subtypes. As the transport direction follows the sodium gradient, the main mode is inward. Labeling studies show a widespread expression of GAT1 and GAT3 in the rat CNS, whereas GAT2 expression is confined to the leptomeninges. GAT1 reaches its highest levels of expression in hippocampal areas and the olfactory bulb, while GAT3 is mainly found in the olfactory bulb, thalamus and hippothalamus. The cellular and sub-cellular distribution also varies pertinently among brain regions, suggesting specific roles in re-uptake and uptake [Scimemi, 2014]. Figure 1.4

Before its release, GABA is sequestered into vesicles in the presynaptic neuron. Uptake of GABA as well as glycine into vesicles is facilitated by the vesicular GABA transporter (VGAT). A reversible interaction between GAD65 and VGAT has been shown, indicating that presynaptic synthesis from glutamate and vesicular uptake is a highly coupled process. [Jin et al., 2003]

GABA’s vital role during embryonic development is further reinforced by knockout experiments with GAD67, GAD65 and VGAT. Knockout of VGAT

4 Figure 1.4: A) Schematized morphology of cortical pyramidal neuron (top) and of the distribution of GAT1 (green) and GAT3 (blue) at synaptic contacts onto these cells (bottom). B),C) As in A, for Purkinje and thalamic relay neurons. [Scimemi, 2014] results in the most severe embryonic malformations due to near complete loss of signaling ability [Kakizaki et al., 2015].

1.3 GABAA Receptors

1.3.1 Structure and Distribution

GABAA receptors form a heterogeneous family of membrane bound, ligand gated anion channels formed by an arrangement of five subunits. Each subunit’s extracellular domain (ECD) is composed of a large N-terminal and a smaller C-terminal, connected by four α-helical transmembrane domains (TMD) and an intracellular domain (ICD) of variable length between TM3 and TM4 [Schofield et al., 1987]. 19 different subunit genes (six α, three β, three γ, three ρ, one δ, one , one θ and one π) have been identified in mammals, located on seven different (1, 3, 4, 5, 6, 15, X)

5 in humans [Sigel and Steinmann, 2012]. Amino acid sequence overlap is greatest within a single subunit class, whereas it decreases when comparing sequences across subunit classes [Ernst et al., 2018]. The phylogenetic tree shown in Figure 1.5 depicts degrees of sequence similarity as well as location [Ernst et al., 2018].

Figure 1.5: The phylogenetic tree depicts the degree of similarity (based on % sequence similarity) of the 19 GABR genes. In humans, 5 harbours a gene cluster with GABRa1, GABRa6, GABRb2, GABRg2 and GABRp; Chro- mosome 4 harbours a homologous cluster with GABRa2, GABRa4, GABRb1 and GABRg1; on Chromosome 15, we find GABRa5, GABRb3 and GABRg3, the X- Chromosome contains GABRa3, GABRq and GABRe. [Ernst et al., 2018]

While we lack complete information on occurence and distribution of specific pentameric assemblies in human brains, most prevalent GABAA receptors consist of two α, two β and one γ subunit [Sieghart, 2015] in alternating order [Tretter et al., 1997]. Detection of subunit messenger ribunucleic acid (mRNA) by real-time polymerase chain reaction (RT-PCR), as well as immunocytological charac- terizations on adult rat brain slices have provided extensive information

6 on estimated subunit distribution among brain regions [H¨ortnaglet al., 2013,Pirker et al., 2000,M¨ohleret al., 2002]. The knowledge obtained from these studies, as well as electron microscopy [Nayeem et al., 1994], homol- ogy models of the extracellular domain and experiments with concatenated subunits allow for likely estimates about major existing asseblies: α1β2γ2

(60%), α2β3γ2 (15-20%), α3βnγ2 (10-15%), α4βnγ/δ (5%), α5β2γ2 (<5%) and α6β2/3γ2 (<5%) [M¨ohleret al., 2002, Olsen and Sieghart, 2008]. For a recent, comprehensive summary of proposed, high likelyhood and tentative receptor assemblies see Table 1.1 published by [Ernst et al., 2018].

The α1 subunit is unequivocally dominant in prevalence and ubiquity in mammalian brains, whereas expression of other α subunits appears to be more confined to certain brain regions [Pirker et al., 2000]. High level expression of the α6 subunit, for instance, is clearly restricted to the cere- bellum in immunochemical mouse brain mappings [H¨ortnagl et al., 2013]. (Figure 1.6) The observed heterogeneity in mRNA expression appears to be strongest during early embryonic development opposed to the adult stage [Ernst et al., 2018].

While conventional αβγ containing assemblies are mainly involved in direct synaptic transmission, δ as well as α4-6 containing expression patterns were often identified in extrasynaptic locations [Wisden et al., 1992, Wang, 2011]. A tonic form of inhibition mediated by δ containing subtypes in cerebellar regions was first shown in-vivo by Chadderton et al. [Chadderton et al., 2004]. Opposed to the fast, transmitter-release dependent inhibition seen at synapses, extrasynaptic GABAA receptors enable a tonic form of inhibition governed by the continuous presence of ligands [Kullmann et al., 2005].

A three dimensional protein structure has been resolved via x-ray crys- tallography for the ECD and TMD of a β3 homopentameric assembly by Miller et al. [Miller and Aricescu, 2014]. Interfaces between two neighbour-

7 Table 1.1: Receptor subunits and compositions proposed to exist (blue), or assumed to exist with high probability (red) or tentative (green). Note that the tentative arrangements are exemplary mentions and not complete. Modified from [Ernst et al., 2018]

Gene/subunit Receptor composition

GABRa1 α1 α1βxγ2, α1β2δ, α1βxδ, α1αxβxγx, α1αxβxδ, α1βx

GABRa2 α2 α2βxγ2, α2β1γ1, α2αxβxγx

GABRa3 α3 α3βxγ2, α3βx, α3βxθ

GABRa4 α4 α4βxγ2, α4βxδ, α4αxβxγx, α4αxβx

GABRa5 α5 α5βxγ2, α5αxβxγx

GABRa6 α6 α6βxγ2, α6α1βxγ2, α6βxδ, α6α1βxδ

GABRb1 β1 αxβ1γ2, α1β1δ, α4β1δ, α2β1γ1, αxαxβ1βx

GABRb2 β2 α1β2γ2, αxβ2γ2, α4,6β2δ, α1β2δ, αxαxβ2βx

GABRb3 β3 α1β3γ2, αxβ3γ2, α4,6β3δ, αxαxβ3βx

GABRg1 γ1 αxβxγ1

GABRg2 γ2 αxβxγ2, αxβxγ2

GABRg3 γ3 αxβxγ3

GABRd δ α1βxδ, α1β2δ, α4βxδ, α6βxδ, α1α6βxδ

GABRe  α3βx, α3θ

GABRq θ α3βxθ GABRp π nothing known GABRr1 ρ1 ρ subunits are thought to coassemble with each other...

GABRr2 ρ2 ...(e.g. ρ1 homopentamers, ργ2)... GABRr3 ρ3 ...see above

ing subunits form binding sites for several known ligands. The space be- tween two extracellular domains contains several accessible pockets, which

8 Figure 1.6: Distribution of GABAA receptor mRNAs (α1-6,β1-3,γ1-2 and δ) in horizontal sections of the mouse brain at the level of Bregma -2.56 to -2.36 mm. Scale bar = 2 mm [H¨ortnaglet al., 2013]

harbour binding sites that serve for agonist recognition and allosteric mod- ulation [Ernst et al., 2018].

9 Figure 1.7: Topology of the typical GABAA receptor subunit polypeptides, based on 4COF and 4PIR crystal structures. The dashed line indicates the ICD fragment that is highly variable in structure sequence and legth between subtypes. The cys-loop is shown in yellow in the otherwise lilac ECD, ECD interface binding site- forming segments A-G are indicated with green capital letters. The junction zone loops and the termini are colored in grey, the four TMD helices and the HX helix, which in homologous cation channels is located in the TMD as well, are coloured in red, while the ICD comprising structures are coloured green. [Ernst et al., 2018]

The extracellular domain comprises a beta-stranded fold, where seven segments (traditionally called ”loops” A-G) form the two interfaces where each subunit is in contact with its neighbors (the principal- or plus- side, and the complementary- or minus- side, respectively). (Figure 1.7 and Figure 1.8)

10 Figure 1.8: This figure shows a side view of the interface between the β3 (yellow ribbon) and α2 (grey ribbon) subunits in a homology model of an α2β3γ2 receptor. Extracellular segments or ”loops” are highlighted by color. The model is based on the available β3 homopentamer structure 4COF [Miller and Aricescu, 2014].

Based on the available structures of related proteins of the same super- family, homology models have been employed to identify and characterise putative binding sites for various ligands [Puthenkalam et al., 2016].

To date, a wide range of binding sites for exogenous ligands such as , , , benzodiazepines, , , clozapine, and some antidepressants have been described and partly confirmed [Ernst et al., 2018, Sieghart, 2015, Squires and Saederup, 1998, Korpi et al., 1995].

11 Figure 1.9: Ten small ligand binding sites are found in atomic structures of

GABAA receptor homologs. The figure shows a side view of a superposition of available structural models listed in the original study. Two subunits of 4COF are shown in ribbon representation (gray). The ligands are depicted in space-filling representation. [Puthenkalam et al., 2016]

Opposed to the direct agonistic effect of GABA, substances with al- losteric activity modulate the receptors sensitivity to the effects of GABA by inducing a conformational change.

1.3.2 Drug-receptor Interactions

This section serves the introduction of basic pharmacologic terminology re- garding drug-receptor interactions. Affinity is a measure for the binding ability of a ligand to a site on the target molecule. It is descriptive of the inter-molecular forces to form a complex. Intrinsic activity describes the probability of an already formed ligand-receptor complex to result in a functional change.

12 Agonists are ligands capable of inducing a functional change in the re- ceptor that corresponds to the main physiological ligand. (E.g. increased channel conductivity in case of GABAA receptors) Antagonists block or dampen the receptors response/state that can be triggered by agonists. In- verse agonists elict a functional change opposite to that of an agonist. The primary, physiological binding site on a receptor is referred to as or- thosteric, while sites at other locations are termed allosteric. Binding at allosteric sites can induce conformational changes that directly affect recep- tor function or mediate modulation of other ligands effects. (Negative al- losteric modulation (NAM) or positive allosteric modulation (PAM)). If two ligands have a common binding site, they are competetive, whereas allosteric effects are non-competitive.

1.3.3 Endogenous GABAA Receptor Ligands

1.3.3.1 GABA

A high affinity, extracellular, orthosteric agonist binding site of GABA was identified at the β(+)/α(-) interface present in the most abundant existing assemblies [Smith and Olsen, 1995]. Binding of GABA induces a confor- mational change, selectively allowing chloride and carbonate anions to pass through the receptors central pore. The HCO3-/Cl- permeability ratio is 0.2-0.4 [Ernst et al., 2018]. Most common receptors are composed of two α, two β and one γ subunit, thus presenting two β(+)/α(-) interfaces. A spe- cific model of putative interactions between GABA and its binding site has been obtained by combining information available from structural models of related proteins from [Ernst et al., 2003, Bergmann et al., 2013]. (Fig- ure 1.8) Simultaneous binding of GABA on both sites greatly enhances the likelihood of channel opening [Baumann et al., 2003]. GABA’s action mode explains the inhibitory nature of postsynaptic GABAA receptors, as chloride equilibrium potential and gradient in the mature neuron lead to an influx of

13 chloride. Immature neurons lack the extent of chloride transporters neces- sary to generate such gradient, thus allowing conditions under which GABA has excitatory properties [Olsen and Sieghart, 2008]. This mechanism is not confined to the developing CNS, but plays a crucial role for the integration of newly born neurons in the adult nervous system [Spitzer, 2010]. Affinity of GABA to its site is highly pH dependent, making adjustments an im- perative prequesite for experimental characterization [Huang et al., 2004]. An additional, low affinity binding site for GABA was recently proposed at the extracellular β(+)/δ(-) interface in αβδ containing receptors [Lee et al., 2016].

Subunit composition unambiguously constitutes a determining factor for GABA sensitivity and therefore determines the physiological function of a receptor. [Olsen and Sieghart, 2009]

1.3.3.2 Other Endogenous Ligands

Apart from GABA, numerous other endogenous substances are known to in- teract with GABAA receptors. It was proposed that most, if not all, GABAA receptor ligands can interact with more than one binding site [Sieghart, 2015].

Histamine is a neurotransmitter primarily known for mediating inflam- matory response. Many additional central nervous functions exist and a his- taminergic model of treatment resistant schizophrenia has been suggested [Ito,

2004]. A direct agonistic effect on β3 homopentamers [Hoerbelt et al., 2016] has been shown. Additionally, positive modulatory properties were observed in extrasynaptic and synaptic α4 containing receptors [Bianchi et al., 2011, Saras et al., 2008]. Based on the available structure of the β3 homopentamer, a putative binding site was described in a pocket located at the extracellular interface between two β3 subunits [Ernst et al., 2018].

14 The endocannabinoid 2-arachidonylglycerol (2AG) has been shown to interact in a modulatory way with β2 containing GABAA receptors [Sigel et al., 2011]. A novel binding site involving the β2 TM4 was suggested [Sigel et al., 2011]. Later docking studies strongly support the proposed endo- cannabinoid binding site localized between TM4 and TM3 [Baur et al., 2013]. present a class of endogenous steroids interacting directly with GABAA receptors. Their effect is determined by the specific char- cateristics of the , as well as the brain region, type of neuron, receptor location and subunit composition. Pregnenolone, progesterone, 5α- dihydroprogesterone (DHP), , adiol, androstane and 3α5α- tetrahydrodeoxycorticosterone (THDOC) were shown to exhibit positive modulatory effects, whereas sulfated and 3β-OH steroids posess negative allosteric properties [Wang, 2011]. Findings suggest that binding of neu- roactive steroids takes place at the lower TMD interface irrespective of the subunit type [Ernst et al., 2018]. Based on recent crystal structures of the

GLIC-GABAAR α1 chimera bound to pregnolone sulfate and THDOC, two TMD binding sites were described. [Laverty et al., 2017].

1.3.4 Exogenous GABAA Receptor Ligands

1.3.4.1 Benzodiazepines

Figure 1.10: structure

15 Benzodiazepines (BZD) present a group of GABAA receptor modula- tors introduced into clinical use in the 1960s. BZD have since then found widespread clinical adoption, constituting frequently prescribed and essen- tial drugs valued for their anxiolytic, anticonvulsant, sedative and muscle re- laxant properties while maintaining favorable toxic profiles [Sieghart, 2015].

A high affinity binding site for the action of BZDs has been identified at the extracellular α(+)/γ(-) interface [Sieghart, 2015]. Binding of most BZD induces a conformational change that increases the affinity of GABA to its orthosteric binding site at the β(+)/α(-) interface. BZDs thus positively modulate GABAA receptors in the presence of GABA [Sieghart, 2015]. Ad- ditionally, an increase of channel opening frequency has been demonstrated in electrophysiological experiments [Bianchi et al., 2009]. There are ligands of the benzodiazepine binding site such as flumazenil (FMZ), that exhibit negative allosteric modulatory (NAM) effects, thus competetively blocking the effects of classical benzodiazepines. FMZ is clinically employed to an- tagonize unwanted or excessive BZD effects. This NAM mechanism might also be responsible for the risk of paradoxical effects of some BZDs under certain circumstances.

Classical 1,4-benzodiazepines such as diazepam, flunitrazepam, temazepam and lorazepam act on receptors containing α1,2,3,5+/γ2,3- interfaces. Op- posed to that, the chemically distinct alternative ligand of the same bind- ing site, , displays relative selectivity for α1γ2 containing recep- tors [Ernst et al., 2018].

Advances in the understanding of subtypes and their specific function have fueled the promising investigation into potentially novel, selective agents with more specific therapeutic profiles [Sieghart et al., 2012, Rudolph and Mohler, 2014].

16 Figure 1.11: Etomidate

1.3.4.2 Anesthetics

General anesthetics present a group of drugs employed for their hypnotic, immobilizing, anesthetic and amnestic properties. While most anesthetic agents positively modulate GABAA receptors, etomidate constitutes a sub- stance with relatively specific action on GABAA receptors containing β2 or

β3 subunits [Uchida et al., 1995,Forman, 2011,Martin et al., 2009]. Etomi- date evokes potent amnestic effects at relatively low doses. This presents a prominent feature when compared to other anesthetic drugs. It was shown that etomidate has negative effects on the plasticity of glutamatergic excita- tory transmission in murine hippocampal slices [Martin et al., 2009]. After experiments with null mutant mice, the amnestic effects can be correlated with α5 containing receptors [Martin et al., 2009]. This finding is compat- ible with the murine mRNA expression pattern of α5 subunits, which is markedly enhanced in the hippocampi. (Figure 1.6) An interaction and possible binding site for etomidate was identified at the β+/α- and β+/β- interfaces in the upper TMD [Chiara et al., 2012].

Barbiturates present a class of drugs that, opposed to the positive mod- ulatory action of benzodiazepines, exhibits direct agonistic properties on

GABAA receptors even in the absence of GABA. Two binding sites have been proposed for at the α+/β- and γ+/β- interfaces of the upper TMD [Chiara et al., 2013].

17 Figure 1.12: Phenobarbital

Propofol is a lipid-soluble anesthetic approved for intravenous applica- tion. Due to its rapid-onset and short effect duration, it is not only em- ployed for induction and maintenance of general anesthesia but also utilized for procedural sedation.

Figure 1.13: Propofol

In the TMDs of the bacterial GLIC, a protein homologous to GABAA receptors, an intrasubunit pocket for propofol exists. Results from photolabeling studies however indicate, that the intra-subunit binding site is not present in GABAA receptors [Sieghart, 2015]. It was shown that propofol causes potentiation and direct activation via binding sites at the

TMD α+/β-, γ+/β- and β+/α- interfaces of GABAA receptors with little or no specific subunit interface selectivity [Sieghart, 2015]. An additional possible binding site has been proposed in the intracellular loop of α1β2 receptors, based on experiments with mutated receptors [Moraga-Cid et al., 2011].

18 Figure 1.14: Ketamine

Ketamine is a water soluble phencyclidine (PCP) derivative with two enantiomers: The S(+) isomer and the R(-) isomer. Since its introduction into clinical use in the 1970s, ketamine has sustained its role as an important intravenous general anesthetic and induction agent. Opposed to anesthetics of the opioid or class, ketamine exhibits stimulatory effects on the cardiovascular system and bronchodilatation. Moreover, its newly de- scribed neuroprotective, antiinflammatory and antitumor effects implicate further potential benefits and applications [Kurdi et al., 2014]. Addition- ally, ketamine treatment for depression has recently gained popularity, as it provides rapid symptom relief compared to other antidepressant medica- tions [Murrough et al., 2013].

While ketamines main anesthetic mechanism is likely caused by strong, non-competetive antagonism at N-methyl-D-aspartate receptors (NMDAR), interactions with a multitude of other receptors are evident. Observa- tions in mice reveal that the metabolite (2R,6R)-hydroxynorketamine might contribute to the antidepressant effects by interaction with AMPA recep- tors [Zanos et al., 2016]. Another experiment with mice suggests the ac- tivation of GABAA receptors as a potential pathway for ketamines antide- pressive action [Rosa et al., 2016]. A potentiation of GABAergic conduc- tance arising from α6-containing GABAA receptors was shown in dissociated granule neurons [Hevers et al., 2008]. Based on the bacterial GLIC crystal

19 structure bound to R-ketamine, a putative binding site was described at the extracellular interface between two subunits [Puthenkalam et al., 2016]. (Figure 1.9) Ketamine effects entail a strong, psychoactive component, that is per- ceived as undesirable and intimidating by many patients. The dissociative and hallucinatory state caused by ketamine is used as a model for the study of positive, negative, and cognitive symptoms, dopaminergic and GABAer- gic dysfunction, age of onset, functional dysconnectivity, and abnormal cor- tical oscillations observed in acute schizophrenia [Frohlich and Van Horn, 2014].

Figure 1.15: Desflourane

Numerous inhalative anesthetics, such as desflourane, isoflourane, en-

flourane, halothane, and ethanol positively modulate GABAA receptors. Acumulated evidence suggests that a binding pocket for these compounds exists within the four α-helices of the α subunit TMD [Sieghart, 2015].

1.3.4.3 Other GABAA Receptor Ligands

A vast variety of other compounds, many of which are naturally ocurring and plant derived, interact with GABAA receptors. , found in plants of the Corydalis species, as well as , isolated from Amanita muscaria, present alkaloids directly interacting via the GABA binding site [Johnston et al., 1968].

20 Many natural and synthetically derived flavonoids display modulatory ef- fects via the high affinity BZD binding site [Baureithel et al., 1997]. presents a flavonoid with potent positive allosteric properties, exhibiting an- ticonvulsive effects in mice [Kavvadias et al., 2004]. Terpenes form a large, heterogenous class of compounds almost ubiqui- tous in plants. They are likely responsible for the therapeutic effects of many traditional, plant derived medicines. Exemplary, , present in plants of the Valeriana genus, mediates anxiolysis and sedation through al- losteric modulation of GABAA receptors [Becker et al., 2014]. Currently, cannabinoids such as cannabidiol (CBD) and tetrahydrocannabi- nol (THC) are increasingly recognized, used and investigated for their broad therapeutic potential. Plant cannabinoids modulate GABAA receptors in differential and complex ways, possibly constituting a core pathway for some of their therapeutic effects. Several binding sites for small molecules and atoms such as Br-, Ba2+ and Br-ethanol have been identified [Puthenkalam et al., 2016]. Figure 1.9 Other known synthetic modulators include some tricyclic antidepressants such as amoxapine, several anticonvulsive drugs such as , the antiparasitic , as well as the atypical antipsychotics chlorpromazin and clozapine [Squires and Saederup, 1998,Sieghart, 2015,Korpi et al., 1995].

21 1.4 GABAB Receptors

Opposed to the ionotropic nature of GABAA receptors, GABAB receptors present a family of G protein-coupled receptors (GPCRs). They are com- posed of two principal subunits, each formed by seven transmembrane do- mains and a large extracellular ’venus fly-trap domain’ (VFTD) that con- tains the ligand-binding site [Gassmann and Bettler, 2012]. The principal subunit compositions are GABAB1a,2 and GABAB1b,2. G-protein coupling occurs intracellularly at the GABAB2 subunit, whereas extracellular ligand binding takes place at the VFTD of GABAB1a or GABAB1b [Rondard et al., 2011]. Additionally, the auxiliary subunits KCTD8, KCTD12, KCTD12b and KCTD16 were identified, which posess modular structure and feature a con- served tetramerization T1 domain as well as one or two carboxy-terminal ’homology’ domains (H1 and H2) of unknown functions [Schwenk et al., 2010]. Located intracellularly, close to the G-protein binding site, these auxiliary subunits take great part in the receptors kinetic and pharmacolog- ical profile [Gassmann and Bettler, 2012].

22 Figure 1.16: a) GABAB1a and GABAB1b are principal subunit isoforms that differ by the presence of two amino-terminal sushi domains in GABAB1a. Whereas these subunits contain the GABA binding site, GABAB2 subunits couple to the G protein. b) The principal subunits associate with the sequence-related auxiliary subunits KCTD8, KCTD12, KCTD12b or KCTD16. The T1 domains form homotetramers that bind to GABAB2. [Gassmann and Bettler, 2012]

2+ GABAB receptors in presynaptic locations inhibit Ca influx via Gβγ release and inhibition of voltage-gated calcium channels (VGCCs). Addi- tionally Gβγ was shown to inhibit vesicle fusion and therefore spontaneous neurotransmitter release through direct interaction with SNARE (soluble N-ethylmaleimide-sensitive factor attachment protein receptor) [Gassmann and Bettler, 2012]. Release of the larger Gα protein also reduces spontaneous neurotransmitter release through inhibition of adenylyl cyclase [Gassmann and Bettler, 2012].

23 Different from its main effectors in presynaptic cells, postsynaptic GABAB receptors mediate inhibition mainly through activation of G protein-coupled inwardly-rectifying potassium channels (GIRKs) via released Gβγ [Gassmann and Bettler, 2012]. Activated GIRK reduce neuronal excitability by gener- ating slow IPSPs [Gassmann and Bettler, 2012].

Figure 1.17: a) Downstream effectors of presynaptic GABAB receptors are de- picted by arrows indicating the effect on target proteins, physiological roles are represented in the boxes. b) Downstream effectors of postsynaptic GABAB recep- tors are depicted by arrows indicating the effect on target proteins, physiological roles are represented in the boxes. [Gassmann and Bettler, 2012]

24 Additional downstream effectors such as the potassium channel TREK2 and N-methyl-D-aspartate receptor (NMDAR) have been identified and are depicted in Figure 1.17 from [Gassmann and Bettler, 2012]. Downregu- lation of protein kinase A (PKA) as a result of adenylyl cyclase inhibition was shown to take part in gene regulation [Fukui et al., 2008].

Post mortem studies reveal decreased expression of GABAB receptors in bipolar disorder, major depressive disorder (MDD) and schizophrenia [Fatemi et al., 2011]. A significant link between polymorphisms in the KCTD12 aux- iliary subunit gene and risk of bipolar disorder was identified in a chinese population [Lee et al., 2011]. Knockout studies with mice support the hy- pothesis that KCTD12 plays a central role in mood regulation [Cathomas et al., 2015].

GABAB receptors have been proposed as novel drug targets for various neuropsychiatric disorders [Kumar et al., 2013].

1.5 Schizophrenia

The World Health Organization (WHO) classifies schizophrenia in code F20 of the International Statistical Classification of Diseases and Related Health Problems (ICD10).

Schizophrenia presents a spectrum of chronic syndromes characterized by severe deviations from ordinary perception and thought. Sense of con- sensual reality is greatly disturbed, impairing the ability to interact in mu- tually comprehensible ways. Delusions, hallucinations and unusual thought processing contribute to an overly intense and often threatening experience of reality. Additionally, a disruption of cognitive and affective functioning contributes to the syndromes severely debilitating effects on many aspects of life. Symptoms are typically divided into two categories, positive and negative. (Table 1.2)

25 Table 1.2: Schizophrenia symptom categories

positive negative

delusions affective flattening hallucinations alogia disorganized speech anhedonia disorganized behaviour avolition

1.5.1 Epidemiology

With a global prevalence between 0.5% and 1.6%, the syndrome poses a major global health challenge. Yearly incidence lies at 0.02%, greatly vary- ing among regions [Messias et al., 2007]. About 50% of cases present an acute initial onset, while the other half is preceded by a prodromal phase. Incidence peaks at an age range of 15-24 years for both sexes, but males have a 30%-40% increased lifetime incidence risk. Positive symptoms oc- cur close to the first hospitalization, while negative symptoms emerge on average five years later [Messias et al., 2007]. Schizophrenia is associated with a weighted average of 14.5 years of potential life lost [Hjorthøj et al., 2017]. Besides a markedly increased risk for suicide, comorbid conditions such as cardiovascular disease contribute to the significantly reduced life expectancy [Skalicky, 2017, Piotrowski et al., 2017].

1.5.2 Pathomechanism

Over past decades, extensive research has been conducted in order to eluci- date pathogenetic mechanisms of schizophrenia. While substantial advances lead to an improved management, identification of distinct underlying path- ways remains elusive. The most persistent and long-standing hypothesis is based on excessive dopaminergic signalling. It was reinforced by post

26 mortem findings, the resemblance between states evoked by dopaminergic, stimulant drugs and positive symptoms, as well as the antipsychotic efficacy of dopamine D2 blocking agents. The link between increased dopamine signalling and schizophrenia remains unequivocal, albeit not satisfactory in light of the current body of evidence. Significant familial coincidence, as high as 41% - 61% among identi- cal twins, lead to extensive efforts to characterize the syndromes genetic makeup. While unequivocal and substantial, the genetic components of schizophrenia are inherited in complex and non-mendelian ways [Harrison, 2015], limiting distinct conclusions. See table 1.2 for select identified loci and genes associated with schizophrenia. Accumulated evidence suggests a complex, likely multicausal etiology involving several neurotransmitter systems beyond dopamine, such as gluta- mate, serotonin and GABA [Schwartz et al., 2012,Balu et al., 2013,Charych et al., 2009, Benes and Berretta, 2001, Verrall et al., 2010, Barkus et al., 2014]. In efforts to pharmacologically address some of schizophrenias cog- nitive symptoms, evidence suggests that α2 containing GABAA receptor assemblies might play a special role as a target [Charych et al., 2009]. Fol- lowing this lead, we focused on α2 containing assemblies for characterization of CLZ effects.

1.5.3 Pharmacological Management

First generation antipsychotic drugs are characterized by strong antagonis- tic effects on dopamine D2 receptors. As a consequence of strong D2 block- ade, these typical antipsychotics evoke extrapyramidal side effects [Meltzer, 2013]. Subsequently, new antipsychotic compounds were discovered that proved to be effective even with relatively weak dopamine D2 blockade. The action mode of these atypical antipsychotics remains largely elusive [Khokhar et al., 2018].

27 Table 1.3: Select loci and genes showing association with schizophrenia. (Modified from [Harrison, 2015])

Locus Implicated Name of Notes Reference to gene gene gene/product biology

12p13.33 CACNA1C L-type calcium Important in neuronal function. Muta- [Bhat et al., 2012] channel α sub- tions cause Timothy syndrome and Bru- unit, type 1c gada syndrome.

12q24.11 DAO D-amino acid oxi- Enzyme which degrades the NMDA re- [Verrall et al., 2010] dase ceptor co-agonist D-serine. Expression and activity increased in schizophrenia. Not GWAS significant.

1q42.2 DISC1 Disrupted in Identified in a large Scottish pedigree [Brandon and Sawa, schizophrenia-1 with a chromosome 1:11 translocation. 2011] A multifunctional scaffolding protein. Not GWAS significant.

11q23.2 DRD2 Dopamine D2 re- Long known to be the key target of an- [Beaulieu and Gainet- ceptor tipsychotic drugs, GWAS data now in- dinov, 2011] dicate that the DRD2 gene may play a role in schizophrenia.

2q33-34 ERBB4 Receptor tyrosine Receptor for neuregulin 1 and some [Mei and Xiong, 2008] kinase erbB4 other ligands. Mutations can cause can- cers. Not GWAS significant.

5q33.2 GRIA1 AMPA recep- The subunit influences properties of the [Barkus et al., 2014] tor subunit 1 AMPA receptor, and affects synaptic (GluA1; GluR1) plasticity and behaviour.

16p13.2 GRIN2A NMDA recep- The subunit influences properties of the [Paoletti et al., 2013] tor subunit 2A NMDAR, including synaptic localisation (GluN2A; NR2A) and channel conductance.

7q21.11- GRM3 Metabotropic Group II metabotropic glutamate recep- [Harrison et al., 2008] 12 glutamate recep- tor (along with mGlu2), acting primar- tor 3 (mGlu3) ily as inhibitory autoreceptors.

1p21.3 MIR137 MicroRNA 137 Non-protein-coding gene. A microRNA, [Pasquinelli, 2012] which regulates other genes by binding to the 3’ untranslated region of their transcripts.

8p12 NRG1 Neuregulin 1 Growth factor, involved in many as- [Mei and Nave, 2014] pects of nervous system development and plasticity. Not GWAS significant.

17p13.3 SRR Serine racemase Enzyme which synthesises D-serine from [Balu et al., 2013] L-serine.

18q21.2 TCF4 Transcription fac- Basic helix-loop-helix transcription fac- [Forrest et al., 2014] tor 4 tor. Haploinsufficiency causes Pitt- Hopkins syndrome.

2q32.1 ZNF804A finger pro- Putative transcription factor. [Hess and Glatt, tein 804A 2014]

28 Up to 30% of patients with schizophrenia show no or only partial re- sponse to treatment with at least two different antipsychotic medications [Con- ley and Kelly, 2001]. Identified predictive factors for treatment resistance include younger age, symptom severeness, previous antidepressant medica- tion, suicide attempt and affective symyptoms [Wimberley et al., 2016]. In those patients, the atypical antipsychotic clozapine was unequivocally con- firmed to be the single most effective treatment choice, providing additional relief from some negative symptoms compared to other antipsychotics [Naber and Lambert, 2009, Kumar et al., 2017].

1.6 Clozapine

Clozapine (CLZ) is a tetracyclic compound with outstanding clinical effec- tiveness in the treatment of psychotic symptoms even in patients resistant to high doses of typical neuroleptics [Attard and Taylor, 2012,Kumar et al., 2017].

Figure 1.18: Clozapine

Clozapine was the first agent to challenge the prevailing dogma of po- tent dopamine D2 antagonism as a premise for antipsychotic efficacy in the 1970s [Stille et al., 1971]. Its mechanism of action remains elusive, mak-

29 ing its study a central opportunity for the identification of potentially novel targets for antipsychotic pharmacotherapy.

While clozapine shows superior antipsychotic efficacy, has effects on neg- ative symptoms and evokes no extrapyramidal side effects opposed to classi- cal antipsychotics, it remains a second or third line treatment. The incidence of drug induced agranulocytosis poses a potentially fatal hazard of clozap- ine treatment, substantially limiting its application [Wahlbeck et al., 1999]. While novel and effective genetic screenings for individual risk prediction were developed [Girardin et al., 2018], mandatory blood monitoring narrows the drug’s potential from an economic and patient compliance perspective. Clozapine treatment is strongly associated with weight gain through yet un- known pathways. A potential correlation between genetic variations in the GABRA2 gene and the risk of weight gain was recently identified [Zai et al., 2015]. With an incidence of 1.3%, clozapine treatment greatly increases the risk of generalized tonic-clonic seizures in a dose-dependant way [Pacia and Devinsky, 1994]. Additionally, risk for potentially fatal cardiomyopathy and myocarditis is markedly increased by CLZ treatment [Merrill et al., 2005]. Premature cardiovascular death presents a principal cause of morbidity in schizophrenic populations [Skalicky, 2017].

While long-term treatment with clozapine as well as haloperidol were found to upregulate GAT1 and GAD67 in rats [Zink et al., 2004a, Zink et al., 2004b], the increase of muscimol binding, reflecting GABAA receptor GABA binding sites in the infralimbic cortex remains a unique feature of clozapine [Zink et al., 2004b]. Korpi et al. proposed an antagonistic effect of clozapine on GABAA receptors based on radioligand studies [Korpi et al.,

1995]. Assemblies containing α6 subunits were mainly unaffected by CLZ with the exception of the β1 containing assembly α6β1γ2. The widespread

α1β2γ2 assembly was inhibited by CLZ and effects were not altered by ap- plication of flumazenil, a compound with benzodiazepine-antagonistic prop-

30 erties [Korpi et al., 1995]. Additional in-vivo experiments further confirmed the GABAAR inhibiting effects of clozapine as well as one of its meabo- lites, N-desmethylclozapine, at high micromolar concentrations [Wong et al., 1996]. Another distinguishing property of CLZ is its effect on the noradrenergic system. It was proposed that α-2 receptor antagonism [Ashby and Wang, 1996], as well as the ability to block reuptake of norepinephrine following synaptic release [Khokhar et al., 2015] might take an important part in its unique clinical profile [Khokhar et al., 2018].

31 2. Materials and Methods

2.1 Two-Electrode Voltage Clamp

Two-electrode voltage clamp (TEVC) is a conventional electrophysiological technique suited for current measurements in large cells. Two electrodes are placed intracellularly. One electrode is used to inject current into the cell and adjust the membrane potential to desired values (voltage clamp). The other electrode is used to record changes in voltage resulting from the cell’s ion channel behaviour under a desired experimental setting.

2.1.1 Xaenopus laevis oocytes

Xenopus laevis oocytes present an established choice for the expression and study of plasma membrane proteins. Due to their physiological role, oocytes are rich in substrates and organelles, making them suitable expression sys- tems. Additionally, Xenopus laevis oocytes express a relatively small num- ber of ”idiogenic” ion channels, making them ideal for the study of recom- binantly expressed channels.

2.1.2 Preparation of oocytes

The methods for isolation, culturing, injection and defolliculation of oocytes were identical with those described by E. Sigel [Sigel et al., 1990].

32 Stage 5-6 oocytes were obtained from the ovary lobe of previously anes- thesized and decapitated mature female Xenopus laevis (Nasco, Fort Atkin- son, WI, USA).

For separation from the sorrounding follicle layer, previously isolated oocytes were incubated in a solution of 1mg/ml collagenase (type IA, Sigma, MO, USA) in xenopus ringer (XR) composed of 90mM NaCl, 1mM MgCl2, 1mM KCl, 1mM CaCl2 and 5mM HEPES-NaOH (pH 7.4) for 60-90 min- utes. Removal of follicular layers after enzymatic digestion was achieved by pipetting cells through sterile pasteur pipettes. For storage periods a solu- tion of 100 U/ml and 100µg/ml streptomycin in XR, referred to as XR+, was used.

Figure 2.1: Defolliculation process - modified from [Sigel and Minier, 2005]

After 24 hours of resting at 18◦C the cells were transferred to XR medium for injection of mRNA. Injection pipettes were pulled from glass capillaries using a P-97 puller (Sutter Instrument, California, USA) und subsequently polished under a Micro Forge MF830 microscope (NARISHIGE SCIEN- TIFIC, Tokyo, Japan). Between 2.5 ng and 4 ng of mRNA were injected into each cell using a Drummond Nanoject 2 injection device (Drummond Scientific Company, PA, USA).

33 For α β γ receptors, a subunit mRNA ratio of 1:1:5 was used, following the approach of Z. Varagic [Varagic et al., 2013], as described previously by E. Sigel and X. Li [Sigel et al., 1990, Li et al., 2003]. During incubation, cells were regularly checked to ensure prompt removal of dead or disintegrating cells from the bath. XR solutions were replaced daily.

2.1.3 Electrophysiological recordings

After incubation for at least 24 hours depending on subunit composition, oocytes were placed on a nylon-grid in a bath of XR solution. The oocytes were impaled with two glass microelectrodes (2-3 Mohm), which were filled with 2 M KCl. A constant flow of XR (6 mL/min) was then applied that could be switched to XR containing GABA and/or clozapine. The perfusion sys- tem is based on a gravity flow mechanism, aiming at constantly washing the cell while keeping mechanical stress at a minimum. Two valves provide the switches between delivery of washing solution and drug-containing solution to the oocyte. An Ismatec Reglo Digital I30 peristaltic pump (Cole-Parmer GmbH, Wertheim, Germany) removes excess solution from the oocyte bath at a constant rate. To test for modulation of GABA induced currents by clozapine, measure- ments were performed in presence of a GABA concentration that triggers 3-5 percent of the respective maximum GABA-elicited current of the individual oocyte (EC3-5). Determination of EC3-5 was performed via measurement of a full GABA dose response curve. The current measured during application of 1mM GABA was referenced as the maximum current. Clozapine was pre-dissolved in dimethyl sulfoxide (DMSO) to 30mM stock solutions and diluted in escalating concentrations in previously deter- mined GABA EC3-5 XR.

34 Figure 2.2: GABA EC3-5 determination sequence

Between two applications, oocytes were washed with XR for up to 15 min to ensure full recovery of sensitivity. Before each clozapine sequence, two control measurements using GABA EC3-5 were performed. All record- ings were performed at room temperature at a holding potential of -60 mV using a Warner OC-725C TEV (Warner Instrument, Hamden, CT, USA) or a Dagan TEV-200A TEV Oocyte clamp (Dagan Corporation, Mineapolis, MN, USA). Data was digitized, recorded and measured using a Digidata 1322A data acquisition system (Axon Instruments, Union City, CA, USA). Data was analyzed using Prism 6 (GraphPad, La Jolla, CA, USA). After elimination of outliers (CI95%) statistical significance was tested with paired t-test and 95% confidence interval. Oocytes expressing assemblies with the known high-affinity benzodi- azepine binding site at the α+/γ2- interface were tested by application of diazepam at GABA EC3-5 to confirm receptor configuration.

35 3. Results

In preliminary screenings we applied 100µmol CLZ at different GABA con- centrations, observing a maximum inhibition of GABA elicted currents at an EC3-5 of GABA. The observations are suggestive of a right shift of the GABA dose-response curve. Additionally, we gathered hints for a reduction of the maximum evoked GABA current at 1 µMol GABA. These prelimi- nary findings are compatible with an allosteric (noncompetitive), negative modulatory mechanism of action. On this basis, further measurements were performed at an EC3-5 of GABA, in order to best portray CLZs effects.

Inhibition of GABA currents by CLZ in α1β1,2,3γ2 receptors was already demonstrated in earlier experiments [Korpi et al., 1995]. We successfully re- produced the described inhibitory effects on α1β2γ2 receptors at micromolar concentrations of CLZ. In order to aid the understanding of potential mechanisms of action and binding sites, we studied CLZs effects on four additional assemblies (α1β2,

α2β3γ2, α2β3γ1 and α2β3) with previously uncertain sensitivity. We focused especially on α2 containing assemblies due to their possible role as targets for drugs effective in treating some of schizophrenias symptoms [Charych et al., 2009]. A summary of the effects on all five measured receptor subtypes is given in Section 3.6.

36 3.1 α1β2

In α1β2 receptors, significant inhibition of GABA elicted currents was ob- served at CLZ concentrations of 10 µmol, 30 µmol and 100 µmol. The mean inhibition reaches a maximum of 59% at 100 µmol CLZ concentration.

Figure 3.1: The range of observed currents at each CLZ concentration is repre- sented by dark blue bars. After elimination of outliners (Q=1%), interpolation of a four parameter logistic regression curve was conducted. A 95% coinfidence interval (CI) of the curve is represented by the dotted line.

Table 3.1: α1β2

CLZ concentration mean inhibition significance

1 µmol 7% 0.6042 3 µmol 26% 0.0765 10 µmol 32% 0.0369* 30 µmol 38% 0.0118* 100 µmol 59% 0.0011*

37 3.2 α2β3

In α2β3 receptors, significant inhibition of GABA elicted currents was ob- served at CLZ concentrations of 30 µmol and 100 µmol. The mean inhibition reaches a maximum of 107.5% at 100 µmol CLZ concentration.

Figure 3.2: The range of observed currents at each CLZ concentration is repre- sented by dark blue bars. After elimination of outliners (Q=1%), interpolation of a four parameter logistic regression curve was conducted. A 95% coinfidence interval (CI) of the curve is represented by the dotted line.

Table 3.2: α2β3

CLZ concentration mean inhibition significance

1 µmol 2% 0.8228 3 µmol 9% 0.5635 10 µmol 37% 0.2407 30 µmol 74% 0.0258* 100 µmol 107.5% 0.0912

38 3.3 α1β2γ2

In α1β2γ2 receptors, significant inhibition of GABA elicted currents was observed at CLZ concentrations of 10 µmol, 30 µmol and 100 µmol. The mean inhibition reaches a maximum of 32% at 100 µmol CLZ concentration.

Figure 3.3: The range of observed currents at each CLZ concentration is repre- sented by dark blue bars. After elimination of outliners (Q=1%), interpolation of a four parameter logistic regression curve was conducted. A 95% coinfidence interval (CI) of the curve is represented by the dotted line.

Table 3.3: α1β2γ2

CLZ concentration mean inhibition significance

1 µmol 4% 0.2952 3 µmol 8% 0.2048 10 µmol 20% 0.0086* 30 µmol 29% 0.0029* 100 µmol 32% 0.001*

39 3.4 α2β3γ1

In α2β3γ1 receptors, significant inhibition of GABA elicted currents was observed at CLZ concentrations of 10 µmol, 30 µmol and 100 µmol. The mean inhibition reaches a maximum of 81% at 100 µmol CLZ concentration.

Figure 3.4: The range of observed currents at each CLZ concentration is repre- sented by dark blue bars. After elimination of outliners (Q=1%), interpolation of a four parameter logistic regression curve was conducted. A 95% coinfidence interval (CI) of the curve is represented by the dotted line.

Table 3.4: α2β3γ1

CLZ concentration mean inhibition significance

1 µmol 12% 0.6401 3 µmol 13% 0.1682 10 µmol 44% 0.0002* 30 µmol 69% 0.0003* 100 µmol 81% 0.0001*

40 3.5 α2β3γ2

In α2β3γ2 receptors, significant inhibition of GABA elicted currents was observed at CLZ concentrations of 10 µmol, 30 µmol and 100 µmol. The mean inhibition reaches a maximum of 79% at 100 µmol CLZ concentration.

Figure 3.5: The range of observed currents at each CLZ concentration is repre- sented by dark blue bars. After elimination of outliners (Q=1%), interpolation of a four parameter logistic regression curve was conducted. A 95% coinfidence interval (CI) of the curve is represented by the dotted line.

Table 3.5: α2β3γ2

CLZ concentration mean inhibition significance

1 µmol 3% 0.4779 3 µmol 7% 0.1851 10 µmol 33% 0.0062* 30 µmol 53% 0.0001* 100 µmol 79% 0.0004*

41 3.6 Summary

While all investigated subtypes responded to CLZ, we show significantly higher maximum mean inhibition in α2 containing assemblies compared to

α1 containing assemblies. The observed inhibition of currents in αβ assem- blies proves, that the existence of a γ subunit is no premise for CLZ sensitiv- ity. Inhibition remained incomplete in all investigated subtypes, supporting the assumed uncompetitive mechanism of action.

Figure 3.6: Mean inhibition of GABA EC3-5 elicted currents by increasing con- centrations of CLZ in all five subtypes

42 Table 3.6: Mean inhibition observed at 100µmol CLZ

receptor assembly mean inhibition significance

α1β2γ2 32% 0.001*

α1β2 59% 0.0011*

α2β3γ2 79% 0.0004*

α2β3γ1 81% 0.0001*

α2β3 107.5% 0.0912

The mean inhibition at 100µmol CLZ is on average 42.07% stronger in α2 containing receptors compared to α1 containing receptors. (P-value 0.001***)

43 4. Discussion

We observed extensive negative modulatory effects on a number of GABAA receptor assemblies. These findings prompt speculation on a possible con- nection to some of CLZs clinical effects.

4.1 Exploring a possible intra-subunit binding pocket

Given the unequivocal link between GABAA receptor inhibition and seizure susceptibility, mediation of CLZs secondary pro-convulsive effects via nega- tive modulation of GABAA receptors becomes a feasible hypothesis. Carry- ing the idea forward, we find many structurally similar molecules known to inhibit GABAA receptors and lower seizure threshold [Pacia and Devinsky, 1994, Squires and Saederup, 1998]. One of these compounds is the typi- cal antipsychotic chlorpromazine, a tricyclic phenothiazine derivative. A 3-dimensional structure of CPZ bound to the bacterial analog ELIC was re- cently resolved [Nys et al., 2016]. The newly found allosteric binding pocket might also be present in GABAA receptors. Figure 4.1 depicts a LigPlot diagram of the residues in the ELIC-CPZ complex involved in CPZ bind- ing. Figure 4.2 shows a rendering of the three-dimensional drug-receptor complex with highlighted representations of the residues from Figure 4.1.

44 Figure 4.1: LigPlot of molecular interactions between ELIC and chlorpromazine in the 5LG3 structure from [Nys et al., 2016]. The structure of CPZ is found in the center, bonds are colored lilac and atoms according to the elements. Half-red circles represent non-ligand residues involved in the hydrophobic contacts to atoms which are surrounded by red lines.

45 Figure 4.2: ELIC-chlorpromazine complex 5LG3 [Nys et al., 2016], highlighting interacting residues from figure 4.1. One chain of the receptor is represented by blue ribbon. Interacting residues are additionally represented by licorice structures colored according to elements. Chlorpromazin is represented in the same way but with slightly thicker bonds.

46 Of the 11 residue pairs compared in table 4.1, we only see two identical ones and two with shared functional characteristics.

Table 4.1: Comparing residues from the bacterial analog structure 5LG3 involved in CPZ binding with structurally aligned residues from 4COF.

5LG3 4COF

Ile23 Ile46 Phe126 Cys150 Val147 Gly177 Thr149 Ile181 Glu150 - Asn151 - Glu155 Glu182 Asp158 Pro184 Trp160 Phe186 Ile162 Ser187 Ala197 Leu214

Legend: Green : amino acids with hydrophobic side chains Purple : amino acids with polar uncharged side chains Red : amino acids with negatively charged side chains Yellow : amino acids with special side chain characteristics

When looking at the two superpositioned structures in Figure 4.3, we find marked difference in the position of a segment in the F-loop. This might be a conformational change caused by binding of CPZ. The remaining con- formation around the pocket remains largely intact in the 4COF structure.

47 Figure 4.3: This rendering shows a superposition of two structures. One chain from the ELIC-chlorpromazine complex (5LG3) is represented by blue ribbon and the corresponding chain from the β3 homopentamer structure structure 4COF is represented by red ribbon. Note the divergence of the F-Loop, possibly a structural change caused by binding of CPZ to the pocket.

48 While highly tentative, the exploration allows for speculation on three ideas:

• The existence of the intra-subunit CPZ binding pocket present in the

bacterial ELIC in GABAA receptors.

• A common molecular target for related drugs with known negative modulatory effects such as clozapine and amoxapine.

• Mediation of the drugs pro-convulsive effects via that target.

4.2 Agranulocytosis

White blood cells and their precursors express GABAA receptors [Sanders et al., 2015], and GABA is in many (non-neuronal) cells involved in proper cell differentiation [Fukui et al., 2008]. While highly speculative, CLZs ef- fects on GABAA receptors might present a pathway for its secondary effect of agranulocytosis.

49 5. Summary and Outlook

Our experiments show significant inhibition of GABA evoked currents by micromolar concentrations of clozapine in five different recombinantly ex- pressed GABAA receptor subtypes: α1β2γ2, α1β2, α2β3γ2, α2β3γ1 and

α2β3. The inhibition is significantly stronger in α2 containing assemblies compared to α1 containing assemblies. Given the likely negative allosteric modulatory mechanism, further ex- periments should be carried out at an EC30 of GABA to best portray ef- fects. Measurements with additional assemblies are needed to substantiate our findings of an increased effect in α2 containing subtypes compared to α1 containing compositions. For subsequent study of the tentative intra-subunit binding pocket ana- log to the newly described site in in the 5LG3 structure [Nys et al., 2016], studies with mutated subunit genes could provide important indications.

50 List of Tables

1.1 Receptor subunits and compositions proposed to exist (blue), or assumed to exist with high probability (red) or tentative (green). Note that the tentative arrangements are exemplary mentions and not complete. Modified from [Ernst et al., 2018]8 1.2 Schizophrenia symptom categories ...... 26 1.3 Select loci and genes showing association with schizophrenia. (Modified from [Harrison, 2015]) ...... 28

3.1 α1β2 ...... 37

3.2 α2β3 ...... 38

3.3 α1β2γ2 ...... 39

3.4 α2β3γ1 ...... 40

3.5 α2β3γ2 ...... 41 3.6 Mean inhibition observed at 100µmol CLZ ...... 43

4.1 Comparing residues from the bacterial analog structure 5LG3 involved in CPZ binding with structurally aligned residues from 4COF...... 47

51 List of Figures

1.1 Simplified decarboxylation of glutamic acid yielding GABA .2 1.2 Simplified transamination of GABA, yielding succinic semi- aldehyde ...... 3 1.3 GABA transporter stoichiometry ...... 4 1.4 Subcellular distribution of GAT isoforms in rats [Scimemi, 2014] ...... 5

1.5 Phylogenetic tree of GABAA receptor subunit genes [Ernst et al., 2018] ...... 6

1.6 GABAA receptor subunit mRNA immunoreactivities in hor- izontal sections of the mouse brain [H¨ortnaglet al., 2013] . .9 1.7 Topology diagram of typical subunit polypeptides [Ernst et al., 2018] ...... 10

1.8 Side view of the extracellular β3+/α2- interface ...... 11 1.9 Small molecule binding sites described by [Puthenkalam et al., 2016] ...... 12 1.10 Benzodiazepine structure ...... 15 1.11 Etomidate ...... 17 1.12 Phenobarbital ...... 18 1.13 Propofol ...... 18 1.14 Ketamine ...... 19 1.15 Desflourane ...... 20

52 1.16 GABAB receptor principal subunit composition [Gassmann and Bettler, 2012] ...... 23

1.17 GABAB receptor downstream effectors and their physiologi- cal roles [Gassmann and Bettler, 2012] ...... 24 1.18 Clozapine ...... 29

2.1 Defolliculation process - modified from [Sigel and Minier, 2005] 33 2.2 GABA EC3-5 determination sequence ...... 35

3.1 Results diagram for α1β2 receptors ...... 37

3.2 Results diagram for α2β3 receptors ...... 38

3.3 Results diagram for α1β2γ2 receptors ...... 39

3.4 Results diagram for α2β3γ1 receptors ...... 40

3.5 Results diagram for α2β3γ2 receptors ...... 41 3.6 Summary of electrophysiological results ...... 42

4.1 LigPlot of molecular interactions between ELIC and chlor- promazine in the 5LG3 structure from [Nys et al., 2016]. The structure of CPZ is found in the center, bonds are colored lilac and atoms according to the elements. Half-red circles repre- sent non-ligand residues involved in the hydrophobic contacts to atoms which are surrounded by red lines...... 45 4.2 ELIC-chlorpromazine complex 5LG3 [Nys et al., 2016] . . . . 46 4.3 Superposition of 4COF and the ELIC-chlorpromazine com- plex (5LG3) (LigPlot) [Nys et al., 2016] ...... 48

53 54 Bibliography

[Alvir et al., 1993] Alvir, J. M., Lieberman, J. A., Safferman, A. Z., Schwimmer, J. L., and Schaaf, J. A. (1993). Clozapine-induced agranulocytosis. incidence and risk factors in the united states. The New England journal of medicine, 329:162–167.

[Ashby and Wang, 1996] Ashby, C. R. and Wang, R. Y. (1996). Pharmacological actions of the atypical antipsychotic drug clozapine: a review. Synapse (New York, N.Y.), 24:349–394.

[Attard and Taylor, 2012] Attard, A. and Taylor, D. M. (2012). Comparative effec- tiveness of atypical antipsychotics in schizophrenia: what have real-world trials taught us? CNS drugs, 26:491–508.

[Balu et al., 2013] Balu, D. T., Li, Y., Puhl, M. D., Benneyworth, M. A., Basu, A. C., Takagi, S., Bolshakov, V. Y., and Coyle, J. T. (2013). Multiple risk pathways for schizophrenia converge in serine racemase knockout mice, a mouse model of hypofunction. Proceedings of the National Academy of Sciences, 110(26):E2400–E2409.

[Barkus et al., 2014] Barkus, C., Sanderson, D., Rawlins, J., Walton, M., Harrison, P., and Bannerman, D. (2014). What causes aberrant salience in schizophrenia? a role for impaired short-term habituation and the (glua1) subunit. Molecular psychiatry, 19(10):1060.

[Baumann et al., 2003] Baumann, S. W., Baur, R., and Sigel, E. (2003). Individual properties of the two functional agonist sites in gaba(a) receptors. The Journal of neuroscience : the official journal of the Society for Neuroscience, 23:11158– 11166.

55 [Baur et al., 2013] Baur, R., Kielar, M., Richter, L., Ernst, M., Ecker, G. F., and Sigel, E. (2013). Molecular analysis of the site for 2-arachidonylglycerol (2-ag) on the beta2 subunit of gaba(a) receptors. Journal of neurochemistry, 126:29–36.

[Baureithel et al., 1997] Baureithel, K. H., B¨uter,K. B., Engesser, A., Burkard, W., and Schaffner, W. (1997). Inhibition of benzodiazepine binding in vitro by amentoflavone, a constituent of various species of hypericum. Pharmaceutica Acta Helvetiae, 72(3):153–157.

[Beaulieu and Gainetdinov, 2011] Beaulieu, J.-M. and Gainetdinov, R. R. (2011). The physiology, signaling, and pharmacology of dopamine receptors. Pharmaco- logical reviews, 63(1):182–217.

[Becker et al., 2014] Becker, A., Felgentreff, F., Schr¨oder, H., Meier, B., and Brattstr¨om,A. (2014). The anxiolytic effects of a extract is based on valerenic acid. BMC complementary and alternative medicine, 14(1):267.

[Ben-Ari et al., 1989] Ben-Ari, Y., Cherubini, E., Corradetti, R., and Gaiarsa, J. (1989). Giant synaptic potentials in immature rat ca3 hippocampal neurones. The Journal of physiology, 416(1):303–325.

[Benes and Berretta, 2001] Benes, F. M. and Berretta, S. (2001). Gabaergic in- terneurons: implications for understanding schizophrenia and bipolar disorder. Neuropsychopharmacology : official publication of the American College of Neu- ropsychopharmacology, 25:1–27.

[Bergmann et al., 2013] Bergmann, R., Kongsbak, K., Sørensen, P. L., Sander, T., and Balle, T. (2013). A unified model of the gaba(a) receptor comprising agonist and benzodiazepine binding sites. PloS one, 8:e52323.

[Bhat et al., 2012] Bhat, S., Dao, D. T., Terrillion, C. E., Arad, M., Smith, R. J., Soldatov, N. M., and Gould, T. D. (2012). Cacna1c (cav1. 2) in the pathophys- iology of psychiatric disease. Progress in neurobiology, 99(1):1–14.

[Bianchi et al., 2009] Bianchi, M. T., Botzolakis, E. J., Lagrange, A. H., and Mac- donald, R. L. (2009). Benzodiazepine modulation of gaba(a) receptor opening frequency depends on activation context: a patch clamp and simulation study. Epilepsy Res, 85(2-3):212–20.

56 [Bianchi et al., 2011] Bianchi, M. T., Clark, A. G., and Fisher, J. L. (2011). The wake-promoting transmitter histamine preferentially enhances alpha-4 subunit- containing gabaa receptors. Neuropharmacology, 61(4):747–52.

[Brandon and Sawa, 2011] Brandon, N. J. and Sawa, A. (2011). Linking neurode- velopmental and synaptic theories of mental illness through disc1. Nature Reviews Neuroscience, 12(12):707.

[Brickley and Mody, 2012] Brickley, S. G. and Mody, I. (2012). Extrasynaptic gaba(a) receptors: their function in the cns and implications for disease. Neuron, 73:23–34.

[Cathomas et al., 2015] Cathomas, F., Stegen, M., Sigrist, H., Schmid, L., Seifritz, E., Gassmann, M., Bettler, B., and Pryce, C. R. (2015). Altered emotionality and neuronal excitability in mice lacking kctd12, an auxiliary subunit of gaba b receptors associated with mood disorders. Translational psychiatry, 5(2):e510.

[Chadderton et al., 2004] Chadderton, P., Margrie, T. W., and H¨ausser,M. (2004). Integration of quanta in cerebellar granule cells during sensory processing. Na- ture, 428:856–860.

[Charych et al., 2009] Charych, E. I., Liu, F., Moss, S. J., and Brandon, N. J. (2009). Gaba(a) receptors and their associated proteins: implications in the eti- ology and treatment of schizophrenia and related disorders. Neuropharmacology, 57:481–495.

[Chiara et al., 2012] Chiara, D. C., Dostalova, Z., Jayakar, S. S., Zhou, X., Miller, K. W., and Cohen, J. B. (2012). Mapping general anesthetic binding site(s) in human alpha1beta3 gamma-aminobutyric acid type a receptors with [(3)h]tdbzl- etomidate, a photoreactive etomidate analogue. Biochemistry, 51(4):836–47.

[Chiara et al., 2013] Chiara, D. C., Jayakar, S. S., Zhou, X., Zhang, X., Savechenkov, P. Y., Bruzik, K. S., Miller, K. W., and Cohen, J. B. (2013). Specificity of intersubunit general anesthetic-binding sites in the transmembrane domain of the human alpha1beta3gamma2 gamma-aminobutyric acid type a (gabaa) receptor. J Biol Chem, 288(27):19343–57.

[Conley and Kelly, 2001] Conley, R. and Kelly, D. (2001). Management of treat- ment resistance in schizophrenia. Biological Psychiatry, 50(10):898–911.

57 [Ernst et al., 2003] Ernst, M., Brauchart, D., Boresch, S., and Sieghart, W. (2003). Comparative modeling of gabaa receptors: limits, insights, future developments. Neuroscience, 119(4):933–943.

[Ernst et al., 2018] Ernst, M., Steudle, F., and Bampali, K. (2018). Gabaa recep- tors. eLS.

[Fatemi et al., 2011] Fatemi, S. H., Folsom, T. D., and Thuras, P. D. (2011). Deficits in gabab receptor system in schizophrenia and mood disorders: A post- mortem study. Schizophrenia research, 128(1):37–43.

[Forman, 2011] Forman, S. (2011). Clinical and molecular pharmacology of etomi- date. Anesthesiology, pages 695–707.

[Forrest et al., 2014] Forrest, M. P., Hill, M. J., Quantock, A. J., Martin-Rendon, E., and Blake, D. J. (2014). The emerging roles of tcf4 in disease and develop- ment. Trends in molecular medicine, 20(6):322–331.

[Frohlich and Van Horn, 2014] Frohlich, J. and Van Horn, J. D. (2014). Review- ing the ketamine model for schizophrenia. Journal of psychopharmacology, 28(4):287–302.

[Fukui et al., 2008] Fukui, M., Nakamichi, N., Yoneyama, M., Ozawa, S., Fujimori, S., Takahata, Y., Nakamura, N., Taniura, H., and Yoneda, Y. (2008). Modulation of cellular proliferation and differentiation through gabab receptors expressed by undifferentiated neural progenitor cells isolated from fetal mouse brain. Journal of cellular physiology, 216(2):507–519.

[Gassmann and Bettler, 2012] Gassmann, M. and Bettler, B. (2012). Regulation of neuronal gaba(b) receptor functions by subunit composition. Nat Rev Neurosci, 13(6):380–94.

[Girardin et al., 2018] Girardin, F. R., Poncet, A., Perrier, A., Vernaz, N., Pletscher, M., F Samer, C., Lieberman, J. A., and Villard, J. (2018). Cost- effectiveness of hla-dqb1/hla-b pharmacogenetic-guided treatment and blood monitoring in us patients taking clozapine. The pharmacogenomics journal.

[Harrison et al., 2008] Harrison, P., Lyon, L., Sartorius, L., Burnet, P., and Lane, T. (2008). The group ii metabotropic 3 (mglur3, mglu3,

58 grm3): expression, function and involvement in schizophrenia. Journal of psy- chopharmacology, 22(3):308–322.

[Harrison, 2015] Harrison, P. J. (2015). Recent genetic findings in schizophrenia and their therapeutic relevance. Journal of Psychopharmacology, 29(2):85–96.

[Hess and Glatt, 2014] Hess, J. L. and Glatt, S. J. (2014). How might znf804a variants influence risk for schizophrenia and bipolar disorder? a literature review, synthesis, and bioinformatic analysis. American Journal of Medical Genetics Part B: Neuropsychiatric Genetics, 165(1):28–40.

[Hevers et al., 2008] Hevers, W., Hadley, S. H., L¨uddens,H., and Amin, J. (2008). Ketamine, but not phencyclidine, selectively modulates cerebellar gabaa recep- tors containing α6 and δ subunits. Journal of Neuroscience, 28(20):5383–5393.

[Hjorthøj et al., 2017] Hjorthøj, C., St¨urup,A. E., McGrath, J. J., and Nordentoft, M. (2017). Years of potential life lost and life expectancy in schizophrenia: a systematic review and meta-analysis. The Lancet Psychiatry, 4(4):295–301.

[Hoerbelt et al., 2016] Hoerbelt, P., Ramerstorfer, J., Ernst, M., Sieghart, W., Thomson, J. L., Hough, L. B., and Fleck, M. W. (2016). Mutagenesis and compu- tational docking studies support the existence of a histamine binding site at the extracellular beta3+beta3- interface of homooligomeric beta3 gabaa receptors. Neuropharmacology, 108:252–63.

[Huang et al., 2004] Huang, R.-Q., Chen, Z., and Dillon, G. H. (2004). Molecular basis for modulation of recombinant alpha1beta2gamma2 gabaa receptors by protons. Journal of neurophysiology, 92:883–894.

[H¨ortnaglet al., 2013] H¨ortnagl,H., Tasan, R. O., Wieselthaler, A., Kirchmair, E., Sieghart, W., and Sperk, G. (2013). Patterns of mrna and protein expression for 12 subunits in the mouse brain. Neuroscience, 236:345–372.

[Ito, 2004] Ito, C. (2004). The role of the central histaminergic system on schizophrenia. Drug news & perspectives, 17:383–387.

[Jin et al., 2003] Jin, H., Wu, H., Osterhaus, G., Wei, J., Davis, K., Sha, D., Floor, E., Hsu, C.-C., Kopke, R. D., and Wu, J.-Y. (2003). Demonstration of functional coupling between gamma -aminobutyric acid (gaba) synthesis and vesicular gaba

59 transport into synaptic vesicles. Proceedings of the National Academy of Sciences of the United States of America, 100:4293–4298.

[Johnston et al., 1968] Johnston, G., Curtis, D., De Groat, W., and Duggan, A. (1968). Central actions of and muscimol. Biochemical pharmacology, 17(12):2488–2489.

[Kakizaki et al., 2015] Kakizaki, T., Oriuchi, N., and Yanagawa, Y. (2015). Gad65/gad67 double knockout mice exhibit intermediate severity in both cleft palate and omphalocele compared with gad67 knockout and vgat knockout mice. Neuroscience, 288:86–93.

[Kavvadias et al., 2004] Kavvadias, D., Sand, P., Youdim, K. A., Qaiser, M. Z., Rice-Evans, C., Baur, R., Sigel, E., Rausch, W.-D., Riederer, P., and Schreier, P. (2004). The flavone hispidulin, a benzodiazepine receptor ligand with positive al- losteric properties, traverses the blood–brain barrier and exhibits anticonvulsive effects. British journal of pharmacology, 142(5):811–820.

[Khokhar et al., 2015] Khokhar, J. Y., Chau, D. T., Dawson, R., and Green, A. I. (2015). Clozapine reconstructed: Haloperidol’s ability to reduce alcohol intake in the syrian golden hamster can be enhanced through noradrenergic modulation by desipramine and idazoxan. Drug & Alcohol Dependence, 152:277–281.

[Khokhar et al., 2018] Khokhar, J. Y., Henricks, A. M., Sullivan, E. D. K., and Green, A. I. (2018). Unique effects of clozapine: A pharmacological perspective. Advances in pharmacology (San Diego, Calif.), 82:137–162.

[Korpi et al., 1995] Korpi, E. R., Wong, G., and L¨uddens,H. (1995). Subtype specificity of gamma-aminobutyric acid type a receptor antagonism by clozapine. Naunyn-Schmiedeberg’s archives of pharmacology, 352:365–373.

[Kullmann et al., 2005] Kullmann, D. M., Ruiz, A., Rusakov, D. M., Scott, R., Semyanov, A., and Walker, M. C. (2005). Presynaptic, extrasynaptic and axonal gabaa receptors in the cns: where and why? Progress in biophysics and molecular biology, 87:33–46.

[Kumar et al., 2013] Kumar, K., Sharma, S., Kumar, P., and Deshmukh, R. (2013). Therapeutic potential of gabab receptor ligands in drug addiction, anxiety, de-

60 pression and other cns disorders. Pharmacology Biochemistry and Behavior, 110:174–184.

[Kumar et al., 2017] Kumar, M., Chavan, B., Sidana, A., and Das, S. (2017). Efficacy and tolerability of clozapine versus quetiapine in treatment-resistant schizophrenia. Indian journal of psychological medicine, 39(6):770.

[Kurdi et al., 2014] Kurdi, M. S., Theerth, K. A., and Deva, R. S. (2014). Ke- tamine: current applications in anesthesia, pain, and critical care. Anesthesia, essays and researches, 8(3):283.

[Laverty et al., 2017] Laverty, D., Thomas, P., Field, M., Andersen, O. J., Gold, M. G., Biggin, P. C., Gielen, M., and Smart, T. G. (2017). Crystal structures of a gaba a-receptor chimera reveal new endogenous neurosteroid-binding sites. Nature Structural and Molecular Biology, 24(11):977.

[Lee et al., 2016] Lee, H. J., Absalom, N. L., Hanrahan, J. R., van Nieuwenhuijzen, P., Ahring, P. K., and Chebib, M. (2016). A pharmacological characterization of gaba, thip and ds2 at binary alpha4beta3 and beta3delta receptors: Gaba activates beta3delta receptors via the beta3(+)delta(-) interface. Brain Res, 1644:222–30.

[Lee et al., 2011] Lee, M., Chen, C., Lee, C., Chen, C., Chong, M., Ouyang, W., Chiu, N., Chuo, L., Chen, C., Tan, H., et al. (2011). Genome-wide association study of bipolar i disorder in the han chinese population. Molecular psychiatry, 16(5):548.

[Li et al., 2003] Li, X., Cao, H., Zhang, C., Furtmueller, R., Fuchs, K., Huck, S., Sieghart, W., Deschamps, J., and Cook, J. M. (2003). Synthesis, in vitro affinity, and efficacy of a bis 8-ethynyl-4h-imidazo[1,5a]- [1,4]benzodiazepine analogue, the first bivalent alpha5 subtype selective bzr/gaba(a) antagonist. Journal of medicinal chemistry, 46:5567–5570.

[Martin et al., 2009] Martin, L. J., Oh, G. H. T., and Orser, B. A. (2009). Eto- midate targets alpha5 gamma-aminobutyric acid subtype a receptors to regulate synaptic plasticity and memory blockade. Anesthesiology, 111:1025–1035.

[Mei and Nave, 2014] Mei, L. and Nave, K.-A. (2014). Neuregulin-erbb signaling in the nervous system and neuropsychiatric diseases. Neuron, 83(1):27–49.

61 [Mei and Xiong, 2008] Mei, L. and Xiong, W.-C. (2008). Neuregulin 1 in neural de- velopment, synaptic plasticity and schizophrenia. Nature Reviews Neuroscience, 9(6):437.

[Meltzer, 2013] Meltzer, H. Y. (2013). Update on typical and atypical antipsychotic drugs. Annual review of medicine, 64:393–406.

[Merrill et al., 2005] Merrill, D. B., Dec, G. W., and Goff, D. C. (2005). Adverse cardiac effects associated with clozapine. Journal of clinical psychopharmacology, 25(1):32–41.

[Messias et al., 2007] Messias, E. L., Chen, C.-Y., and Eaton, W. W. (2007). Epi- demiology of schizophrenia: review of findings and myths. Psychiatric Clinics, 30(3):323–338.

[Michel and Trudeau, 2000] Michel, F. J. and Trudeau, L. E. (2000). Clozapine in- hibits synaptic transmission at gabaergic synapses established by ventral tegmen- tal area neurones in culture. Neuropharmacology, 39:1536–1543.

[Miller and Aricescu, 2014] Miller, P. S. and Aricescu, A. R. (2014). Crystal struc- ture of a human gabaa receptor. Nature, 512:270–275.

[Moraga-Cid et al., 2011] Moraga-Cid, G., Yevenes, G. E., Schmalzing, G., Peo- ples, R. W., and Aguayo, L. G. (2011). A single phenylalanine residue in the main intracellular loop of α1 γ-aminobutyric acid type a and glycine receptors influences their sensitivity to propofol. Anesthesiology: The Journal of the Amer- ican Society of Anesthesiologists, 115(3):464–473.

[Murrough et al., 2013] Murrough, J. W., Iosifescu, D. V., Chang, L. C., Al Jurdi, R. K., Green, C. E., Perez, A. M., Iqbal, S., Pillemer, S., Foulkes, A., Shah, A., et al. (2013). Antidepressant efficacy of ketamine in treatment-resistant major depression: a two-site randomized controlled trial. American Journal of Psychiatry, 170(10):1134–1142.

[M¨ohleret al., 2002] M¨ohler,H., Fritschy, J. M., and Rudolph, U. (2002). A new benzodiazepine pharmacology. The Journal of pharmacology and experimental therapeutics, 300:2–8.

[Naber and Lambert, 2009] Naber, D. and Lambert, M. (2009). The catie and cutlass studies in schizophrenia. CNS drugs, 23(8):649–659.

62 [Nayeem et al., 1994] Nayeem, N., Green, T. P., Martin, I. L., and Barnard, E. A. (1994). Quaternary structure of the native gabaa receptor determined by electron microscopic image analysis. Journal of neurochemistry, 62:815–818.

[Nys et al., 2016] Nys, M., Wijckmans, E., Farinha, A., Yoluk, O., Andersson, M., Brams, M., Spurny, R., Peigneur, S., Tytgat, J., Lindahl, E., et al. (2016). Allosteric binding site in a cys-loop receptor ligand-binding domain unveiled in the crystal structure of elic in complex with chlorpromazine. Proceedings of the National Academy of Sciences, 113(43):E6696–E6703.

[Olsen and Sieghart, 2008] Olsen, R. W. and Sieghart, W. (2008). International union of pharmacology. lxx. subtypes of gamma-aminobutyric acid(a) receptors: classification on the basis of subunit composition, pharmacology, and function. update. Pharmacological reviews, 60:243–260.

[Olsen and Sieghart, 2009] Olsen, R. W. and Sieghart, W. (2009). Gaba a recep- tors: subtypes provide diversity of function and pharmacology. Neuropharma- cology, 56(1):141–8.

[Olsen RW, 1999] Olsen RW, D. T. (1999). GABA Synthesis, Uptake and Re- lease, volume Basic Neurochemistry: Molecular, Cellular and Medical Aspects. Philadelphia: Lippincott-Raven, 6 edition.

[Pacia and Devinsky, 1994] Pacia, S. V. and Devinsky, O. (1994). Clozapine- related seizures experience with 5,629 patients. Neurology, 44(12):2247–2247.

[Paoletti et al., 2013] Paoletti, P., Bellone, C., and Zhou, Q. (2013). Nmda receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease. Nature Reviews Neuroscience, 14(6):383.

[Pasquinelli, 2012] Pasquinelli, A. E. (2012). Micrornas and their targets: recogni- tion, regulation and an emerging reciprocal relationship. Nature Reviews Genet- ics, 13(4):271.

[Piotrowski et al., 2017] Piotrowski, P., M Gondek, T., Krolicka-Derkegowska, A., Misiak, B., Adamowski, T., and Kiejna, A. (2017). Causes of mortality in schizophrenia: An updated review of european studies. Psychiatria Danubina, 29(2):108–120.

63 [Pirker et al., 2000] Pirker, S., Schwarzer, C., Wieselthaler, A., Sieghart, W., and Sperk, G. (2000). Gaba(a) receptors: immunocytochemical distribution of 13 subunits in the adult rat brain. Neuroscience, 101:815–850.

[Puthenkalam et al., 2016] Puthenkalam, R., Hieckel, M., Simeone, X., Suwattana- sophon, C., Feldbauer, R. V., Ecker, G. F., and Ernst, M. (2016). Structural studies of gabaa receptor binding sites: Which experimental structure tells us what? Front Mol Neurosci, 9:44.

[Rondard et al., 2011] Rondard, P., Goudet, C., Kniazeff, J., Pin, J.-P., and Pr´ezeau,L. (2011). The complexity of their activation mechanism opens new possibilities for the modulation of mglu and gabab class cg protein-coupled re- ceptors. Neuropharmacology, 60(1):82–92.

[Rosa et al., 2016] Rosa, P. B., Neis, V. B., Ribeiro, C. M., Moretti, M., and Ro- drigues, A. L. S. (2016). Antidepressant-like effects of ascorbic acid and ke- tamine involve modulation of gabaa and gabab receptors. Pharmacological Re- ports, 68(5):996–1001.

[Rudolph and Mohler, 2014] Rudolph, U. and Mohler, H. (2014). Gabaa recep- tor subtypes: Therapeutic potential in down syndrome, affective disorders, schizophrenia, and autism. Annu Rev Pharmacol Toxicol, 54:483–507.

[Sanders et al., 2015] Sanders, R. D., Grover, V., Goulding, J., Godlee, A., Gurney, S., Snelgrove, R., Ma, D., Singh, S., Maze, M., and Hussell, T. (2015). Immune cell expression of gabaa receptors and the effects of diazepam on influenza infec- tion. Journal of neuroimmunology, 282:97–103.

[Saras et al., 2008] Saras, A., Gisselmann, G., Vogt-Eisele, A. K., Erlkamp, K. S., Kletke, O., Pusch, H., and Hatt, H. (2008). Histamine action on vertebrate gabaa receptors: direct channel gating and potentiation of gaba responses. J Biol Chem, 283(16):10470–5.

[Schnitzler and Gross, 2005] Schnitzler, A. and Gross, J. (2005). Normal and pathological oscillatory communication in the brain. Nature reviews. Neuro- science, 6:285–296.

[Schofield et al., 1987] Schofield, P. R., Darlison, M. G., Fujita, N., Burt, D. R., Stephenson, F. A., Rodriguez, H., Rhee, L. M., Ramachandran, J., Reale, V.,

64 and Glencorse, T. A. (1987). Sequence and functional expression of the gaba a receptor shows a ligand-gated receptor super-family. Nature, 328:221–227.

[Schonfeld-Dado and Segal, 2009] Schonfeld-Dado, E. and Segal, M. (2009). Activity-dependent survival of neurons in culture: a model of slow neurodegener- ation. Journal of neural transmission (Vienna, Austria : 1996), 116:1363–1369.

[Schwartz et al., 2012] Schwartz, T. L., Sachdeva, S., and Stahl, S. M. (2012). Glu- tamate neurocircuitry. theoretical underpinnings in schizophrenia. Frontiers in Pharmacology, 3:195.

[Schwenk et al., 2010] Schwenk, J., Metz, M., Zolles, G., Turecek, R., Fritzius, T., Bildl, W., Tarusawa, E., Kulik, A., Unger, A., Ivankova, K., et al. (2010). Native gaba b receptors are heteromultimers with a family of auxiliary subunits. Nature, 465(7295):231.

[Scimemi, 2014] Scimemi, A. (2014). Structure, function, and plasticity of gaba transporters. Frontiers in cellular neuroscience, 8:161.

[Sieghart, 2015] Sieghart, W. (2015). Allosteric modulation of gabaa receptors via multiple drug-binding sites. Advances in pharmacology (San Diego, Calif.), 72:53–96.

[Sieghart et al., 2012] Sieghart, W., Ramerstorfer, J., Sarto-Jackson, I., Varagic, Z., and Ernst, M. (2012). A novel gaba(a) receptor pharmacology: drugs in- teracting with the alpha(+) beta(-) interface. British journal of pharmacology, 166:476–485.

[Sigel et al., 2011] Sigel, E., Baur, R., R´acz,I., Marazzi, J., Smart, T. G., Zimmer, A., and Gertsch, J. (2011). The major central endocannabinoid directly acts at gaba(a) receptors. Proceedings of the National Academy of Sciences of the United States of America, 108:18150–18155.

[Sigel et al., 1990] Sigel, E., Baur, R., Trube, G., M¨ohler,H., and Malherbe, P. (1990). The effect of subunit composition of rat brain gabaa receptors on channel function. Neuron, 5:703–711.

[Sigel and Minier, 2005] Sigel, E. and Minier, F. (2005). The xenopus oocyte: sys- tem for the study of functional expression and modulation of proteins. Molecular nutrition & food research, 49:228–234.

65 [Sigel and Steinmann, 2012] Sigel, E. and Steinmann, M. E. (2012). Structure, function, and modulation of gaba(a) receptors. J Biol Chem, 287(48):40224–31.

[Skalicky, 2017] Skalicky, D. (2017). Cardiovascular disease in patients with schizophrenia. The Medical journal of Australia, 207(4):179.

[Smith and Olsen, 1995] Smith, G. B. and Olsen, R. W. (1995). Functional do- mains of gabaa receptors. Trends in pharmacological sciences, 16:162–168.

[Spitzer, 2010] Spitzer, N. C. (2010). How gaba generates depolarization. The Journal of physiology, 588(5):757–758.

[Squires and Saederup, 1998] Squires, R. F. and Saederup, E. (1998). Clozapine and several other antipsychotic/antidepressant drugs preferentially block the same ’core’ fraction of gaba(a) receptors. Neurochemical research, 23:1283–1290.

[Stille et al., 1971] Stille, G., Lauener, H., and Eichenberger, E. (1971). The phar- macology of 8-chloro-11-(4-methyl-1-piperazinyl)-5h-dibenzo(b,e)(1,4)diazepine (clozapine). Il Farmaco.

[Tretter et al., 1997] Tretter, V., Ehya, N., Fuchs, K., and Sieghart, W. (1997). Sto- ichiometry and assembly of a recombinant gabaa receptor subtype. The Journal of neuroscience : the official journal of the Society for Neuroscience, 17:2728– 2737.

[Uchida et al., 1995] Uchida, I., Kamatchi, G., Burt, D., and Yang, J. (1995). Eto- midate potentiation of gabaa receptor gated current depends on the subunit composition. Neuroscience letters, 185:203–206.

[Varagic et al., 2013] Varagic, Z., Ramerstorfer, J., Huang, S., Rallapalli, S., Sarto- Jackson, I., Cook, J., Sieghart, W., and Ernst, M. (2013). Subtype selectivity of alpha+beta- site ligands of gabaa receptors: identification of the first highly spe- cific positive modulators at alpha6beta2/3gamma2 receptors. Br J Pharmacol, 169(2):384–99.

[Verrall et al., 2010] Verrall, L., Burnet, P., Betts, J., and Harrison, P. (2010). The neurobiology of d-amino acid oxidase and its involvement in schizophrenia. Molecular psychiatry, 15(2):122.

66 [Wahlbeck et al., 1999] Wahlbeck, K., Cheine, M., Essali, A., and Adams, C. (1999). Evidence of clozapine’s effectiveness in schizophrenia: a systematic re- view and meta-analysis of randomized trials. American Journal of Psychiatry, 156(7):990–999.

[Wang, 2011] Wang, M. (2011). Neurosteroids and gaba-a receptor function. Front Endocrinol (Lausanne), 2:44.

[Wimberley et al., 2016] Wimberley, T., Støvring, H., Sørensen, H. J., Horsdal, H. T., MacCabe, J. H., and Gasse, C. (2016). Predictors of treatment resistance in patients with schizophrenia: a population-based cohort study. The Lancet Psychiatry, 3(4):358–366.

[Wisden et al., 1992] Wisden, W., Laurie, D. J., Monyer, H., and Seeburg, P. H. (1992). The distribution of 13 gabaa receptor subunit mrnas in the rat brain. i. telencephalon, diencephalon, mesencephalon. The Journal of neuroscience : the official journal of the Society for Neuroscience, 12:1040–1062.

[Wolfe et al., 2010] Wolfe, J., Houweling, A. R., and Brecht, M. (2010). Sparse and powerful cortical spikes. Current opinion in neurobiology, 20:306–312.

[Wong et al., 1996] Wong, G., Kuoppam¨aki, M., Hietala, J., L¨uddens, H., Syv¨alahti, E., and Korpi, E. R. (1996). Effects of clozapine metabolites and chronic clozapine treatment on rat brain gabaa receptors. European journal of pharmacology, 314:319–323.

[Zai et al., 2015] Zai, C. C. H., Tiwari, A. K., Chowdhury, N. I., Brandl, E. J., Shaikh, S. A., Freeman, N., Lieberman, J. A., Meltzer, H. Y., M¨uller,D. J., and Kennedy, J. L. (2015). Association study of gabaa alpha2 receptor sub- unit gene variants in antipsychotic-associated weight gain. Journal of clinical psychopharmacology, 35:7–12.

[Zanos et al., 2016] Zanos, P., Moaddel, R., Morris, P. J., and Georgiou (2016). Nmdar inhibition-independent antidepressant actions of ketamine metabolites. Nature, 533(7604):481.

[Zink et al., 2004a] Zink, M., Schmitt, A., May, B., M¨uller,B., Braus, D. F., and Henn, F. A. (2004a). Differential effects of long-term treatment with clozapine or haloperidol on gaba transporter expression. Pharmacopsychiatry, 37:171–174.

67 [Zink et al., 2004b] Zink, M., Schmitt, A., May, B., M¨uller, B., Demirakca, T., Braus, D. F., and Henn, F. A. (2004b). Differential effects of long-term treatment with clozapine or haloperidol on gabaa receptor binding and gad67 expression. Schizophrenia research, 66:151–157.

68