<<

Tolerance and Antagonism to Effects in the

Rat CNS

Şahruh Türkmen

Umeå University Medical Dissertations New Series No 1026 – ISSN 0346-6612 – ISBN 91-7264-073-1

From the Department of Clinical Sciences, Obstetrics and Gynecology Umeå University

Tolerance and antagonism to allopregnanolone effects in the rat CNS

Şahruh Türkmen

Umeå 2006

1

Department of Clinical science, Obstetrics & Gynecology. Umeå University SE-901 85 Umeå, Sweden

Cover illustration:

Copyright © 2006 by Sahruh Turkmen

ISSN 0346-6612

ISBN 91-7264-073-1

Printed by Print & Media, Umeå University, 2006

2

To my family

……. Tolerance, a view as politically pernicious, philosophically false and scientifically ambiguous……..

3

Contents

Pages

ABSTRACT …………………………………………………………..6

LIST OF ORIGINAL PAPERS ……………………………………..7

ABBREVIATIONS …………………………………………………..8

1. INTRODUCTION …………………………………………………9 1.1. SEX AND CNS ………………………………....9 1.1.1 Genomic and non-genomic mechanisms …………………..11 1.1.2 and Neuroactive .…………………...11 1.1.3 Allopregnanolone ………………………………………….12 1.1.4 Allopregnanolone …………………………..13 1.2. GABA-A RECEPTORS ………………………………………..15 1.2.1 Structure …………………………………………………....15 1.2.2 Distributions and functions ………………………………...17 1.2.3 GABA-A and antagonist ...... ….18 1.2.4 Allopregnanolone and interaction with the GABA-A receptor ……………………………………………………20 1.3. CLINICAL CONDITIONS RELATED TO THE SEX SETROID HORMONES EFFECT IN THE ………….21 1.4. TOLERANCE ………………………………………………….24 1.4.1 The GABA-A receptor and tolerance …………………….25 1.4.2 Changes in the GABA-A receptor during tolerance ……...26 1.4.3 Tolerance to allopregnanolone ……………………………27 1.5. AND .………………………………..28 1.5.1 Spatial learning and memory ……………………………..28 1.5.2 The GABA-A receptor modulators and spatial learning and memory ……………………………………………….29

4

2. AIMS ……………………………………………………………….31

3. METHODS…………………………………………………………32 3.1 In situ hybridization………………………………………….32 3.2 Immunohistochemistry……………………………………....32 3.3 Chloride ion uptake assay …………………………………...33 3.4 EEG threshold method ………………………………………33 3.5 Morris Water Maze test ……………………………………..35 3.6 assays ……………………………………………..37 3.7 Statistical analysis …………………………………………...37

4. RESULTS AND DISCUSSIONS …………………………………39 4.1 Antagonism of an allopregnanolone effect ………………….39 4.2 Chronic allopregnanolone tolerance …………………………42 4.3 Acute allopregnanolone tolerance and the GABA-A receptor subunits .……………………………………………46 4.4 Persistence of allopregnanolone tolerance .………………….50

5. GENERAL DISCUSSIONS .………………………………………55

6. CONCLUSIONS .…………………………………………………..57

7. ACKNOWLEDGMENTS .………………………………………...58

8. REFERENCES .…………………………………………………….60

APPENDIX - Papers I – IV

5

ABSTRACT Many studies have suggested a relationship between sex steroids and negative mental and mood changes in women. Allopregnanolone, a potent endogenous of the GABA-A receptor and a of , is one of the most accused neuroactive steroids. Variations in the levels of neuroactive steroids that influence the activity of the GABA-A receptor cause a vulnerability to mental and emotional pathology. In women, there are physiological conditions in which allopregnanolone production increases acutely (e.g. ) or chronically (e.g. , ), thus exposing the GABA-A receptor to high allopregnanolone . In such conditions, tolerance to allopregnanolone probably develops. We have evaluated the 3β-hydroxy steroid UC1011 as a functional antagonist to allopregnanolone-induced negative effects in rats. In vivo, we used the Morris Water Maze (MWM) test of learning and, in vitro, we studied chloride ion uptake into cortical and hippocampal membrane preparations. The steroid UC1011 reduces the allopregnanolone-induced learning impairment in the MWM and the increase in chloride ion uptake induced by allopregnanolone. To detect whether chronic tolerance develops to an allopregnanolone-induced condition, male rats were pretreated with allopregnanolone injections for three or seven days. These rats were then tested in the Morris Water Maze for five days and compared with relevant controls. Rats with seven days’ allopregnanolone pretreatment experienced improved performance compared with the acutely allopregnanolone-exposed group, reflecting chronic tolerance development. To study the GABA-A receptor changes in acute allopregnanolone tolerance, we used the silent second (SS) anaesthesia threshold method. At acute tolerance, 90 minutes of anaesthesia, the abundance of the GABA-A receptor α4 subunit and the expression of the α4 subunit mRNA in the thalamus ventral-posteriomedial (VPM) nucleus were reduced. There was also a significant negative correlation between the increase in the allopregnanolone dose needed to maintain anaesthesia and the α4 mRNA in the VPM nucleus. We also investigated whether allopregnanolone tolerance was still present one or two days after the end of the anaesthesia-induced acute tolerance. Tolerance persisted to one day, but not two days, after the treatment and the α4 subunit mRNA expression in the VPM nucleus was negatively related to the allopregnanolone doses needed after one day. In conclusion, the current thesis shows that the substance UC1011 can reduce the allopregnanolone-induced negative effects in the water maze test. Chronic allopregnanolone tolerance can develop to the effects of allopregnanolone. Allopregnanolone tolerance persists one day after the induction of acute allopregnanolone tolerance. The GABA-A receptor α4 subunit in the thalamus might be involved in the development and persistence of acute tolerance to allopregnanolone.

Key words: Allopregnanolone, GABA-A receptor, UC1011, Spatial memory, Morris Water Maze, Tolerance, Silent second, mRNA

6

LIST OF ORIGINAL PAPERS:

I) 3beta-20beta-dihydroxy-5alpha-pregnane (UC1011) antagonism of the GABA potentiation and the learning impairment induced in rats by allopregnanolone. Turkmen S., Lundgren P., Birzniece V., Zingmark E., Backstrom T., and Johansson I.M. Eur J Neurosci. 2004: 20, 1604-1612.

II) Tolerance development to Morris water maze test impairments induced by acute allopregnanolone. Türkmen S., Löfgren M., Birzniece V., Bäckström T., and Johansson IM. Neuroscience. 2006: 139(2): 651-9

III) GABA-A receptor changes in acute allopregnanolone tolerance. Birzniece* V., Türkmen* S., Lindblad C., Zhu D., Johansson IM., Bäckström T., Wahlström G. Eur J Pharmacol. 2006: 535(1-3):125-34. * The first two authors have contributed equally in this work.

IV) Remaining anaesthetic allopregnanolone tolerance in male rat. Türkmen Ş., Wahlström G., Bäckström T., Johansson I.M. Manuscript.

These papers are reproduced with the permission of the copyright holders.

7

ABBREVIATIONS

ANOVA analysis of variance BZP CC celite chromatography CNS central DG dentate gyrus DHEAS sulphate DHP dihydroprogesterone DNA deoxyribonucleic acid EEG electroencephalography GABA γ-aminobutyric acid HPLC high performance liquid chromatography HRT hormone replacement therapy HSD dehydrogenase i.m. intramuscular i.p. intraperitoneal i.v. intravenous LSD least significant difference MDR maintenance dose rate MPA acetate NMDA N-methyl-D-aspartate P450scc p450 side-chain cleavage PMDD premenstrual dysphoric disorder PMS RIA radioimmunoassay SEM standard error of the mean SS silent second THP tetrahydroprogesterone UC1011 3β-20β-dihydroxy-5α-pregnane VPM ventral posteriomedial thalamic nucleus

8

1. INTRODUCTION

Many women report mental or emotional changes when the concentrations of ovarian hormones fluctuate during the menstrual cycle, pregnancy and the post-partum period. In addition, problems become apparent with exogenous steroid treatment such as contraceptive usage and hormone replacement therapy. Some women also point out that their cognitive performance varies to some degree with the hormonal changes in the menstrual cycle and at . The sex steroid hormones, and progesterone, influence a variety of behaviours in vertebrates and have crucial effects on female reproductive functions such as the development and maturation of reproductive organs, carbohydrate and osmotic equilibrium (Whitfield et al., 1999). Animal studies have also shown that the steroid hormones of gonadal origin act on the foetal (CNS) tissues and permanently alter the and behaviours they will mediate in the future (Grady et al., 1965; Phoenix et al., 1959). After , the female reproductive system undergoes cyclical changes, the menstrual cycle. The most obvious of these changes is the that occurs with the cyclic changes in hormonal production of the -pituitary-ovarian axis (Adashi, 1992). The mechanisms by which the ovarian hormones affect brain functions are not completely understood. In the present thesis, we investigated the mechanisms of steroid and receptor interactions and receptor changes at exposure to a neuroactive steroid.

1.1. SEX HORMONES AND CNS Glands secrete hormones, which trigger changes in expression in distant parts of the body. There are two main structural classes of hormones, 1) steroid hormones and 2) hormones. Steroid hormones are synthesised in steroidogenic tissues such as the , adrenal glands and . Some steroid hormones are also synthesised in the nervous system (Baulieu and Robel, 1990; Speroff, 1999). The steroid hormones all contain the same structure on the A-ring with a keto group at the 3 position and a double bond between carbon 4 and 5; the only exception is the . The precursor of steroid hormones (with the exception of retinoic

9

acid) is . Moreover, with the exception of , they all contain a basic skeletal structure of four fused rings, referred to as perhydrocyclopentanophenanthrene, and the same atomic numbering system as for cholesterol (Fig 1 B). Members of the superfamily and hydroxysteroid dehydrogenase (HSD) are involved in the of the steroid hormones (Mellon and Griffin, 2002; Serra, 1983; Speroff, 1999). The first step in steroidogenesis is the conversion of cholesterol to by the side-chain cleavage enzyme cytochrome P450scc located in the inner mitochondrial membrane (Speroff, 1999). Pregnenolone is then converted to progesterone by 3βHSD located in the (Speroff, 1999). From this point, progesterone can enter a range of biochemical pathways to synthesise a variety of steroid hormones, such as the main neuroactive steroid, allopregnanolone (Fig. 1).

A) Cholesterol

3αHSD-II/III 3α5α-THP(allopregnanolone) 5α- DHP P450scc 3α5β-THP 3βHSD 5α reductase-I/II (isoallopregnanolone) Pregnenolone

3αHSD 3βHSD-I/II 3α5β-THP () 5β- DHP Progesterone 3β5β-THP 3βHSD 5β- reductase (isopregnanolone)

B) C)

Figure 1: A) One steroid synthesis pathway (progesterone derivates) (adapted from Compagnone and Mellon, 2000). B) The perhydrocyclopentanophenanthrene structure and numbering. C) The allopregnanolone (5α-pregnane-3α-ol-20-one) structure. DHP, dihydroprogesterone; THP, tetrahydroprogesterone; P450scc, P450 side-chain cleavage enzyme; HSD, hydroxysteroid dehydrogenase

10

1.1.1 Genomic and non-genomic mechanisms of action Steroid hormones are known to act through intracellular receptors, located in the nucleus and cytoplasm that acts as ligand-activated transcription factors in the regulation of gene expression (genomic mechanism). However, this classical mechanism of action is unable to account for a variety of rapid steroid effects that occur within milliseconds to seconds. Steroid hormones also have effects through membrane-bound receptors such as ion channels (non-genomic or non-nuclear mechanism), a mechanism in which receptor binding to DNA and changed RNA synthesis is not required (Baulieu and Robel, 1995; Frye et al., 1992; McEwen, 1994; Rupprecht, 2003). However, steroid actions on membrane-bound ion channels can secondarily activate the so that a change in gene expression occurs. This may be the case during tolerance development, for example, when the target receptor changes to be less sensitive to the drug of action (Biggio et al., 2003). A full understanding of these mechanisms has not yet been obtained. The (ER) exists as an alpha (ERα) and a beta (ERβ) isoform; progesterone also has two intracellular receptors, PR-A and PR-B. It was long believed that steroid hormones acted exclusively through these classical receptors. However, both estrogen and progesterone can also affect the target cells by non-genomic action (Coiret et al., 2005; Duras et al., 2005). In most tissues co-expressing estrogen and progesterone receptors, estrogen controls the amount of progesterone receptors (up-regulation by estrogen) and thereby also the sensitivity to progestins (steroids with progesterone-like effects). The is down-regulated by exposure to endogenous progesterone (Bouchard, 1999). The presence of these hormone receptors within different regions of the brain has been described in several studies (Genazzani et al., 2000; Jung-Testas et al., 1992; McEwen et al., 1982). The non-genomic actions of steroid hormones at membrane receptors are often associated with signalling pathways. Using such pathways, steroids rapidly regulate multiple cellular functions (Christ et al., 1995). The rapid time course and the insensitivity of the effect to transcription and synthesis inhibitors are essential conditions in the non-genomic action (for a review see Schmidt et al., 2000).

1.1.2 Neurosteroids and Neuroactive Steroids The term “” refers to steroids formed in the nervous system. The name was created by Dr. E.E. Baulieu and colleagues, when they showed

11

that the brain has the capacity to synthesise its own steroids in situ (Corpechot et al., 1981). Neurosteroids can be synthesised de novo in the brain from cholesterol or from steroid precursors that cross the blood-brain barrier. The closely related term “neuroactive steroids” includes all the steroids that are neuronal modulators (Gasior et al., 1999; Paul and Purdy, 1992). The terms “neurosteroid” and “neuroactive steroid” are often used interchangeably. The term “GABAergic steroid” is used for the group of steroids that act on the γ-amino butyric acid type-A (GABA-A) receptor (Majewska, 1990). Neurosteroids can have non-genomic actions (Paul and Purdy, 1992), thereby rapidly altering neuronal excitability by binding and modulating membrane receptors for inhibitory or excitatory (Rupprecht and Holsboer, 1999). A non-genomic neurosteroid effect is the interaction between the GABA-A receptor and progesterone , such as the neurosteroid allopregnanolone (Gee, 1988; Majewska et al., 1986).

1.1.3 Allopregnanolone The neurosteroid allopregnanolone (3α-hydroxy-5α-pregnanone-20-one or tetrahydroprogesterone) is an A-ring reduced pregnane steroid and one of the main circulating metabolites of progesterone (Fig. 1). Allopregnanolone is synthesised in the of the , in the and in the central nervous system (Holzbauer et al., 1985; Ottander et al., 2005; Paul and Purdy, 1992). Allopregnanolone is an allosteroic modulator, which, at nanomolar concentrations potentiates the GABA effect at the GABA-A receptor. It is assumed that the modulatory action of allopregnanolone is mediated by binding to the GABA-A receptor (Majewska et al., 1986; Rupprecht, 2003). Recently, one preliminary study reported the presence of two distinct allopregnanolone binding sites on the GABA-A receptor (Hosie, 2005). Allopregnanolone is synthesised from progesterone by a two-step reaction (Fig. 1A). The first step is irreversible and is mediated by 5α- reductase that converts progesterone to 5α-hydroxyprogesterone (5α- pregnanedione). The second step is reversible and is mediated by 3αHSD that converts 5α-hydroxyprogesterone to allopregnanolone (Karavolas and Hodges, 1990; Krieger and Scott, 1984; Mellon and Griffin, 2002). The positive allosteric modulation of the GABA-A receptors by allopregnanolone exerts , , anaesthetic, and effects, at least in animals (Arafat et al., 1988; Backstrom et al., 2003a; Bitran et al., 1991; Carl et al., 1990; Frye and Duncan, 1994; Lancel et al., 1997; Reddy, 2004). In humans, allopregnanolone may play a critical role in neuropsychiatric disorders, in particular the disruption of

12

memory processes and sleep (Reddy, 2002; Schumacher et al., 2003). Moreover, the possibility has to be considered that the neurosteroid allopregnanolone might not only selectively modulate the GABA-A receptor but might also affect the NMDA and receptors, which could induce clinical effects other than those induced by the potentiation of GABA-A receptors (Johansson and Le Greves, 2005; Landgren et al., 1998; Widmer et al., 2003).

1.1.4 Allopregnanolone concentrations In both humans and animals, the allopregnanolone levels in brain and plasma change during various physiological and pathological conditions, i.e. stress, pregnancy and the menstrual cycle (in animals, the estrus cycle) (Bixo et al., 1997; Bixo et al., 1986; Paul and Purdy, 1992). Allopregnanolone is rapidly redistributed from the brain and the first compartment half-life is very short (Purdy et al., 1990b). However, the long-term is slow, probably due to a large accumulation in fat tissue (Timby et al., 2005; Zhu et al., 2004). In ovariectomised female rats, the progesterone half-life and clearance are around 75 and 165 min (respectively) (Gangrade et al., 1992). The of medroxyprogesterone acetate, a progestin, are affected by , i.e. increased substance and, as a result, lower medroxyprogesterone acetate doses are suggested for women of more advanced age (<60) (Jarvinen et al., 2004). In post-menopausal women, plasma concentrations of allopregnanolone increased in a dose-dependent manner and their time course closely followed the time course of the progesterone concentrations reached after treatment. The peak plasma concentrations of allopregnanolone after injecting (i.m.) the 25, 50 and 100 mg doses of progesterone was 5 ng/ml, 8.07 ng/ml, and 10.3 ng/ml and these peaks occurred seven, four and six hours after injections (respectively) (de Wit et al., 2001). In this female category, a low oral dose (20 mg) of progesterone 12 hours after treatment produces an allopregnanolone concentration comparable to that achieved physiologically in premenopausal women (Andreen et al., 2006), while the allopregnanolone plasma concentrations were increased to 15.9 and 19.5 ng/ml after two hours when 400 or 800 mg (respectively) progesterone were administered vaginally (Andreen et al., 2005). In women with regular a single dose of progesterone (100 mg; i.m.), increased the peak plasma concentration of allopregnanolone to 12.83 ng/ml after three hours which then decreases to 10.2 ng/ml by 13 hours (de Wit et al., 2001). The brain and plasma concentrations of allopregnanolone are partly regulated by independent mechanisms (Paul and Purdy, 1992). The brain,

13

like the gonads and adrenals, is a steroidogenic organ which is able to produce allopregnanolone, since the enzymes 5α-reductase and 3αHSD that are needed for the production of allopregnanolone from progesterone are present in many brain areas, notably in the cortex and the (Cheney et al., 1995; Compagnone and Mellon, 2000). The brain concentration of allopregnanolone is significantly higher than the plasma level in both cycling and pregnant female rats and both brain and plasma levels of allopregnanolone temporally follow those of progesterone (Corpechot et al., 1993). In humans, allopregnanolone accumulates in the brain and increases in plasma during the of the menstrual cycle (Bixo et al., 1997). The serum allopregnanolone level does not differ significantly between young fertile women in the follicular phase and age- matched men (0.25 vs. 0.23 ng/ml), whereas in women during the luteal phase the allopregnanolone level is significantly higher (approximately 1.2- 3.8 ng/ml). In non-pregnant women, plasma pregnanolone and allopregnanolone concentrations of between 25 to 50 ng/ml cause and 168-223 ng/ml cause anaesthesia (Carl et al., 1990; Sundstrom et al., 1999; Timby et al., 2005). In postmenopausal women, the serum allopregnanolone level does not change with age, while a progressive decrease is observed in men, reaching the lowest level after 60 years of age (Bicikova et al., 1998; Genazzani et al., 1998). The serum allopregnanolone level increases during pregnancy, with the highest levels (> 50 ng/ml) at term (Luisi et al., 2000). The allopregnanolone concentrations in maternal and foetal sera are similar in 20-week pregnant women (3.2 ng/ml), while levels about three times lower are present in the amniotic fluid (Bicikova et al., 2002). In the plasma of intact adult male rats, allopregnanolone is barely detectable, whereas the plasma concentrations in intact cycling female rats on random days of the estrus cycle are much higher (approximately 2 ng/ml) (Corpechot et al., 1993). The allopregnanolone concentration in the brain of male rats is 0.6-1.2 ng/, while the concentration in the female rat brain (whole) on random days of the estrus cycle is approximately 11 ng/g protein (Corpechot et al., 1993). The allopregnanolone concentration in the rat brain cortex can reach 3-10 ng/g protein after swimming stress (Paul and Purdy, 1992; Purdy et al., 1991) and 30 ng/g protein 30 minutes after acute electroshock (Barbaccia et al., 1997). The allopregnanolone concentration in the brain differs between regions and depends on the phase of estrus cycle, i.e., in the cortex, 4 ng/g protein in pro-estrus and 6 ng/g protein in the estrus phase of the cycle has been measured (Paul and Purdy, 1992; Purdy et al., 1991). In pregnant rats, the allopregnanolone concentration in plasma shows

14

a maximum increase on day 19 (approximately 11 ng/ml) and returns to control values on day 21, while, in the brain cortex, the allopregnanolone value is greatly increased (> 30 ng/g protein) (Corpechot et al., 1993). It has also been indicated that there is a significant regional and interspecies diversity in the progesterone metabolic pathways and allopregnanolone accumulation in the brain, in which the incubation of rat brain slices with progesterone leads to the accumulation of 5α-DHP and allopregnanolone, whereas the incubation of mouse brain slices leads to the accumulation of 5α-DHP and 20α-DHP and that of monkey brain slices leads to the accumulation of only 20α-DHP (Korneyev et al., 1993). In the MWM test, the allopregnanolone concentration 511 ng/g in the rat frontal cortex impaired learning (eight minutes after an i.v. injection), the brain levels then decreased by 30-60% during the next 12 min (Johansson et al., 2002). Rats remain anaesthetised as long as the allopregnanolone level in the brain frontal cortex is above 955 ng/g (Korneyev and Costa, 1996; Wang et al., 1996). The anti- effect of allopregnanolone in the cortex has been determined to be in the range of 9.5-10.5 ng/g (Bitran et al., 1993).

1.2 GABA-A RECEPTORS 1.2.1 Structure GABA is the major inhibitory in the CNS. GABA is released from GABAergic after synthesis from glutamate by the enzyme decarboxylase (GAD). GABA activates three classes of receptors, the GABA-A, GABA-B and GABA-C receptors. In this paper, only the GABA-A receptor will be discussed, as allopregnanolone modulates this receptor. The transmembrane heteropentameric GABA-A receptor is composed of five subunits that belong to different subunit classes. So far, six α, four β, four γ, one δ, one ε, one π and one θ subunit have been identified. Some of the subunits also exist as splice variants, such as the γ2 subunit that is expressed as γ2S (short form) and γ2L (long form). There is 30-40% sequence identity between the GABA-A receptor subunit families and the other ligand gated ion channels (i.e. receptor) and all the sequences within each subunit family are homologous, with about 70- 80% sequence identity. The subunits form a symmetrical structure around the , each subunit contributing to the wall of the channel (Fig. 2) (DeLorey and Olsen, 1992; Olsen and Tobin, 1990; Schofield et al., 1987). The pentameric combination of the GABA-A receptors most often includes two α, two β and one γ subunit. The α1, β2/3, and γ2 subunits co-exist in many GABA-A receptors and the composition of

15

the receptor varies between different brain regions (Backus et al., 1993; Benke et al., 1994; Chang et al., 1996). It is assumed that GABA-A receptor expression on the surface is controlled by a dynamic mechanism. The GABA-A receptor subunits are translated on the endoplasmic reticulum and then assemble into receptor complexes. Unassembled receptor subunits are degraded, whereas assembled receptors are targeted via the Golgi apparatus to the plasma membrane. Surface receptors cluster at synapses. Receptors can also dissociate from the synaptic clusters and re-enter the mobile pool by endocytosis (Barnes, 2001). After the translation in the endoplasmic reticulum, GABA-A receptors slowly appear on the cell surface (Gorrie et al., 1997). It has been suggested that GABA-A receptor endocytosis is a mechanism that underlies tolerance development in physical dependence on , a GABA-A receptor positive modulator substance (Poisbeau et al., 1997; Tehrani et al., 1997).

Figure 2: GABA-A receptor structure and chloride ion channel.

The GABA-A receptor contains specific binding sites for at least GABA, , , benzodiazepines and anaesthetic steroids (Fig. 2). These substances bind to the GABA-A receptor and regulate the gating (opening and closing) of the chloride ion channel. The GABA-A receptor exhibits distinct pharmacological and electrophysiological properties depending on the subunit composition (Mehta and Ticku, 1999; Olsen and Tobin, 1990; Whiting et al., 1999). The localisation of the various binding sites has been investigated. The (a GABA analogue) is located at the interface of the α and β subunits of the receptor (Pregenzer et al., 1993). The β subunit has also been suggested as a requirement for the GABA-A receptor blockers TBPS (t-

16

butylbicyclophosphorothionate) and picrotoxin (Slany et al., 1995). The benzodiazepine binding site is located at the interface of the α and γ subunits of GABA-A receptors (Wong et al., 1992).

1.2.2 Distributions and functions The individual subunits exhibit distinct distributions throughout the nervous system and the expression of multiple subunits in the same neurons shows the existence of a large variety of GABA-A receptor subunits in the brain neurons. This suggests that different receptor subunits might have specific functions (Fritschy et al., 1998; Wisden et al., 1992). Studies to determine the function of the GABA-A receptor subunits indicate that the sedative, amnesic and probably ataxic and partly anticonvulsive actions of the benzodiazepines are mediated by the receptors containing the α1 subunit, whereas the benzodiazepine anxiolytic action takes place via receptors that contain the α2 subunit (McKernan et al., 2000). The specific localisation of the α5 subunit in the hippocampus indicates that the memory-impairing effects of benzodiazepines are possibly mediated by the α5 subunit. This conclusion is supported by the recent finding that α5 subunit-deficient animals display increased learning and memory abilities (Collinson et al., 2002; Fritschy et al., 1998; McKernan et al., 2000). An increase in α4 subunit level is associated with conditions characterised by increased neuronal excitability, increased susceptibility and higher levels of anxiety (Frye and Bayon, 1999; Gallo and Smith, 1993; Gulinello et al., 2001). On the other hand, the receptors in the brain containing the α6 subunit are exclusively located in cerebellar granule cells and, like the α4 subunit, make the GABA-A receptors insensitive to benzodiazepines (Khan et al., 1996; Mehta and Shank, 1995). It is suggested that the GABA-A receptors containing the α6 subunit are responsible for the motor control of the body (Korpi et al., 1993). The γ2 subunit is involved in synaptic targeting and clustering and allopregnanolone sensitivity. The loss of the GABA-A receptor clusters results in the loss of the synaptic clustering molecule and synaptic GABAergic function (Essrich et al., 1998). In the GABA-A receptor assembly, the γ2 subunit also enhances inhibitory synaptic transmission by reducing the desensitisation of the GABA-A receptor ion channel (Tia et al., 1996). In contrast to synaptic receptors that contain the γ2 subunits, receptors containing the δ subunit are located extrasynaptically and mediate tonic inhibition. The γ2 and δ subunits compete for assembly with some of the α and β subunits in the GABA-A receptor complex, to regulate the balance of synaptic and extrasynaptic receptors (Luscher and Keller, 2004; Peng et al.,

17

2002). Functionally, the δ subunit affects the sensitivity of the GABA-A receptors to neurosteroids (Mihalek et al., 1999). It has been suggested that the β2 subunit is present in about 50% of the GABA-A receptors in the CNS (McKernan and Whiting, 1996) and the β2 subunit enhances the GABA-A receptor-evoked response to some anaesthetic drugs (e.g. ) but not to the neurosteroid allopregnanolone (Belelli et al., 2002; Reynolds et al., 2003).

1.2.3 GABA-A receptor agonist and antagonist steroids Certain steroids interact directly with the GABA-A receptor (GABA- steroids), as either allosteroic or antagonists. Allopregnanolone and tetrahydrodeoxycorticosterone (5α-pregnan-3α,21-diol-20-one, or THDOC) display GABA agonistic features (Harrison et al., 1987; Majewska et al., 1986). Plasma and brain concentrations of the neurosteroids allopregnanolone and THDOC increase during stress, but this increase can be prevented by adrenalectomy. Allopregnanolone and THDOC potentiate the binding of the GABA-A receptor agonist muscimol and the benzodiazepine and they allosterically inhibit the binding of the TBPS. Both neurosteroids also stimulate chloride uptake and currents in synaptoneurosomes and neurons (Majewska, 1990; Reddy, 2002; Schwartz et al., 1987). Positive GABA modulator steroids display different potencies in different regions of the brain, since the regionally variable subunit composition of the GABA-A receptor is reflected in the neurosteroid response (Jussofie, 1993b; Schwabe et al., 2005). The reduction of by 5α-reductase generates 5α-, which is then converted to 3α- (3α-Diol) by 3αHSD. 3α-androstanediol is also a powerful GABA-A receptor-modulating neurosteroid, with allopregnanolone-like properties (Frye et al., 1996; Martini et al., 1993; Reddy and Kulkarni, 2000). Pregnenolone sulphate (PS) and dehydroepiandrosterone sulphate (DHEAS) are also naturally occurring GABA-A active steroids (Paul and Purdy, 1992). PS exhibits mixed agonistic features, increasing muscimol binding at nanomolar concentrations and reducing it at micromolar concentrations, but it behaves as an allosteroic GABA-A (Majewska et al., 1988). In pre-optic cells, PS has no effect by itself on GABA-A receptor activation by spontaneously released GABA but inhibits the allopregnanolone-induced chloride flux enhancement (Haage et al., 2005). PS can also activate the and is therefore excitatory in two ways. The neurosteroid DHEAS reversibly blocks GABA-induced currents and behaves as an allosteroic GABA antagonist at the GABA-A receptor

18

(Majewska et al., 1990). The mechanism of the antagonistic neurosteroids is not completely clear. Neuroactive steroids, which are antagonists against GABA-A receptor steroid agonists, also exist. Studies show that 3β-pregnane steroids are non- competitive GABA antagonists (Maitra and Reynolds, 1998; Wang et al., 2002). The 3β-hydroxy of allopregnanolone, isoallopregnanolone (3β-hydroxy-5α-pregnane-20-one, Fig. 1), however, is not active by itself at the GABA-A receptor with GABA concentrations of less than 15 μM (Wang et al., 2002). The action of isoallopregnanolone is dependent on GABA-A receptor activation (Lundgren et al., 2003; Wang et al., 2000). Isoallopregnanolone functions as an allopregnanolone antagonist at the GABA-A receptor in the brain (Lundgren et al., 2003; Wang et al., 2000) and with receptors expressed in the Xenopus Lavis oocytes (Wang et al., 2002). Isoallopregnanolone is produced in the brain in a similar way to allopregnanolone, with the exception that the 3βHSD enzyme is needed instead of the 3αHSD enzyme (Weir et al., 2004). In vitro, isoallopregnanolone selectively inhibits the allopregnanolone-induced chloride ion uptake into rat cortical tissue (Lundgren et al., 2003). Isoallopregnanolone is also able to block the allopregnanolone-induced inhibition of the population spike in hippocampus slices (Wang et al., 2000). The fact that a 3β-pregnane steroid also functions as an allopregnanolone antagonist in vivo was demonstrated for the first time in Paper I of this thesis. Recently, the in vivo inhibitory action of isoallopregnanolone was shown on the anaesthetic action of allopregnanolone (Backstrom et al., 2005). The neuroactive steroid also has an in vivo inhibitory impact on the neuroactive steroid effects on operant self- administration in rats (O'Dell et al., 2005). Different results relating to the effects on GABAergic neurons in the CNS have been reported (Morrow et al., 1990b; Woodbury, 1952). One study showed that , , and their reduced metabolites at low nanomolar concentrations potentiate TBPS binding at the GABA-A receptor and slightly reduce it at high nanomolar and micromolar concentrations (Majewska, 1987; Majewska et al., 1985). Contrary to these results, studies in our laboratory showed that 3α-hydroxy 5α-reduced metabolites of both cortisol and cortisone reduce GABA- mediated chloride ion uptake by themselves. However, when interactions between allopregnanolone, THDOC and the glucocorticoid metabolites were studied, an enhancement of the allopregnanolone effect was noted (Stromberg et al., 2005).

19

1.2.4 Allopregnanolone interaction with the GABA-A receptor The neurosteroid allopregnanolone enhances the GABA effect on the GABA-A receptor. This can be observed at nanomolar concentrations of the neurosteroid. The threshold concentrations for enhancement is 10-30 nM, but at higher concentrations (μM) allopregnanolone may in neurons also directly open the GABA-A receptor chloride channel (Belelli and Lambert, 2005). However, neurosteroid interactions with the GABA-A receptor are not completely understood. It is suggested that the allopregnanolone enhancement of GABA at the GABA-A receptor results from an increase in opening frequency and burst duration (Macdonald and Olsen, 1994; Majewska, 1992). In clinical and other in vivo situations, the effects of allopregnanolone on the CNS via the GABA-A receptor can be divided into three categories: 1) direct effects on the receptor function; 2) induction of tolerance; and 3) abstinence (withdrawal) (Backstrom et al., 2003a). The direct effect of allopregnanolone on mood and behaviour displays a biphasic pattern. At high concentrations, allopregnanolone induces a general enhancement of the GABA-A receptor and thereby induces sedative, hypnotic and even anaesthetic effects in both human and animals (Carl et al., 1990; Norberg et al., 1987; Sundstrom and Backstrom, 1998). In certain individual, however, a low concentration of allopregnanolone induces the loss of impulse control, negative mood and / (Backstrom et al., 2003a; Miczek et al., 2003). The modulatory effect of allopregnanolone depends on hormonal variations (Reddy and Kulkarni, 1999). In addition, studies of recombinant GABA-A receptors provide evidence of a subunit selective heterogeneity of steroid sensitivity, in which the γ3 subunit increases the maximum efficacy of allopregnanolone compared with other γ subunit subtypes (Maitra and Reynolds, 1999). Another study also indicates that both the α and γ subunits are important determinants of allopregnanolone sensitivity (Maitra and Reynolds, 1998). The δ subunit predominantly co-assembles with the α4 and α6 subunits, the second of which is limited to cerebellar granule cells (Huh et al., 1996; Quirk et al., 1995). Disruption of the δ subunit produces a considerable attenuation of behavioural responses to neuroactive steroids but not to other neuromodulatory drugs (Mihalek et al., 1999). It has been shown that the expression of GABA-A receptor subunits changes during the estrus cycle. In late diestrus, the number of α4, β1 and δ subunit immunopositive cells increases in the midbrain of rats. Quick changes in the expression level of subunits at the beginning of the estrogen peak (late diestrus) also indicate that the expression of the GABA-A receptor can

20

change within a few hours (Landgren et al., 1998; Lovick et al., 2005; Reddy and Kulkarni, 1999). In the hippocampal CA1 region, for example, but not in the dentate gyrus region the effect of allopregnanolone changes with the estrus cycle (Landgren et al., 1998). Phosphorylation is suggested as an additional mechanism involved in the dynamic regulation of allopregnanolone interaction with the GABA-A receptors (Brussaard et al., 2000; Fancsik et al., 2000; Harney et al., 2003; Lambert et al., 2003). The outcome of GABA-A receptor phosphorylation depends on the nature of the kinase, the subunit composition of the receptor and the residues that are phosphorylated. In magnocellular oxytocin neurons of the hypothalamus, the promotion of phosphorylation suppresses the inhibitory allopregnanolone response, whereas, in the hippocampus, phosphorylation increases the interaction between allopregnanolone and the GABA-A receptor (Harney et al., 2003; Koksma et al., 2003).

1.3 CLINICAL CONDITIONS RELATED TO THE EFFECT OF SEX HORMONES IN THE FEMALE BRAIN The sex steroid hormones act on neuronal and glial cells, modulating neurotransmitter synthesis, neurotransmitter receptor expression and synaptic transmission (Compagnone and Mellon, 2000). The sex steroids may therefore exert clinically significant effects on the central nervous system. For example, the menstrual cycle influences the occurrence of in some female epileptics and the performance of certain cognitive tests (Backstrom, 1976; Postma et al., 1999; Reddy, 2004). However, the appearance of mood and behavioural symptoms during the menstrual cycle, oral contraceptive usage, hormone replacement therapy, pregnancy and postpartum has been extensively reported. The mechanisms underlying these symptoms are not clear.

Pre-Menstrual Dysphoric Disorder (PMDD): one of the most obvious interactions between sex steroid hormones and the CNS is found in the premenstrual syndrome (PMS). The term “premenstrual dysphoric disorder” is used for a more severe condition, when the psychological symptoms are so disruptive that they impair the function of the woman and there is a need for treatment (AmericanPsychiatricAssociation., 1994). PMDD is a psychoneuroendocrine disorder characterised by cyclical changes in physical, psychological and behavioural symptoms that are related to the formation of the corpus luteum and fluctuations in gonadal hormones (Backstrom et al., 2003b; Sundstrom Poromaa et al., 2003). The etiology of

21

PMS is not yet clearly known, but the ovarian steroids, as well as changes in GABAergic and serotoninergic functions, are involved (AC et al., 2005; Backstrom et al., 1983; Halbreich, 1990; Hammarback et al., 1985; Hammarback et al., 1991; Sundstrom Poromaa et al., 2003). In one study, different estrogen and progesterone levels and a lower luteal phase allopregnanolone concentration have been shown and the possibility of the defective synthesis of GABA agonist steroids in PMDD was put forward (Monteleone et al., 2000). The majority of studies, including those from our laboratory, show that pregnanolone, PS, 5α-DHP, allopregnanolone and estradiol plasma levels, in the luteal phase of PMS patients and control subjects, are similar. However, within individuals, there is a relationship between symptom severity and the serum concentrations of estradiol, allopregnanolone and PS (Wang et al., 1996). PMDD patients during the luteal phase less sensitive to the effects of neurosteroids, benzodiazepines and , and show reduced GABA-A receptor sensitivity to these substances in the CNS (Nyberg et al., 2004; Sundstrom et al., 1997; Sundstrom and Backstrom, 1998). In rats, it has been shown that short-term exposure to allopregnanolone reduces the sensitivity to benzodiazepines, with changes in the composition of the GABA-A receptor and in association with increased anxiety (Gulinello et al., 2001). Similar changes are suggested as the basic mechanism behind PMDD.

Pregnancy and postpartum mood disorders: pregnancy results in marked alterations in both the psychology and the of women. Memory impairment, cognitive changes and mood changes during pregnancy and in the postpartum period have been documented (Davey, 1995). Estrogen increases 1,000 times during pregnancy and progesterone levels also rise compared with the level before conception (Brett and Baxendale, 2001; Casey, 1993). Allopregnanolone is the progesterone metabolite that mainly increases during pregnancy (Morrow et al., 1987; Paul and Purdy, 1992). Chronic allopregnanolone treatment results in reduced receptor responses to this neurosteroid (Yu et al., 1996b). If this down-regulation occurs during pregnancy, the rapid postpartum fall in progesterone metabolites will increase the excitability (hormone withdrawal effect), due to the down- regulated receptor. This could be an explanation for postpartum mood disorders and increased seizure attacks during labour and delivery among women with (Tomson, 1997; Yu et al., 1996a; Yu et al., 1996b). The memory deficit might be related to allopregnanolone rather than to progesterone itself (Brett and Baxendale, 2001; Johansson et al., 2002). It has also been suggested that hormone-related phenomena such as PMDD are

22

related to the occurrence of postpartum mood disorders (Bloch et al., 2005). The etiology of postpartum mood disorders might then be related to differential neurosteroid sensitivity.

Oral contraceptive usage: negative mood symptoms are well-known side- effects of oral contraceptives (OC) in women (Rosenberg and Waugh, 1998). Women who experience adverse mood symptoms during OC usage appear to be the same women as those with PMDD (Cullberg, 1972). It appears that women with mild PMDD might benefit from OC, whereas women with severe PMDD develop more negative mood symptoms as a result of the treatment (Backstrom, 1996). This might be caused by a different brain sensitivity to ovarian hormones, in which the GABA-A receptor is less sensitive in women with PMDD compared with controls (Sundstrom Poromaa et al., 2003). It has also been shown that long-term daily administration of OC induces a persistent and marked decrease in the synthesis and accumulation of progesterone and the metabolite allopregnanolone, and at least in the rats changes the GABA-A receptor composition in the brain (Follesa et al., 2001).

Hormone Replacement Therapy (HRT): negative mood changes are clinically well-known side effects of combined HRT (estrogen + progesterone) and constitute a major problem in relation to compliance. Women on estrogen-only HRT do not develop negative mood changes (Hammarback et al., 1985). All progestagens (hormones with a similar effect to that of progesterone) appear to be able to induce negative mood changes (Backstrom, 1996). In fact, various studies of the effect of estrogen have shown significant improvements in mood in postmenopausal women (Montgomery et al., 1987; Rasgon et al., 2002). In contrast, negative mood symptoms begin shortly after adding progesterone to the treatment and continue to rise in severity until the progesterone treatment has been ended (Bjorn et al., 2000; Hammarback et al., 1985; Panay and Studd, 1997). In postmenopausal women with a previous PMDD history, the lower dose of progestagen increased negative mood symptoms (Andreen et al., 2003; Bjorn et al., 2002). Menopausal women with higher pre-treatment serum estradiol levels reported more negative effects on mood than women with low estrogen levels when given a combination of estrogen-progesterone (Klaiber et al., 1997). On the other hand, another study proposed that, together with the progesterone, a higher dose of estrogen caused significantly more negative mood symptoms than a lower dose (Bjorn et al., 2003). Allopregnanolone is the main metabolite of progesterone and the

23

chronic administration of allopregnanolone results in the down-regulation of the GABA-A receptor. In addition, these receptors become less sensitive to neurosteroids (Yu et al., 1996a). As a result, in HRT, reduced sensitivity to neurosteroids caused by repeated exposure to progestagen is suggested as a possible mechanism underlying negative mood changes (Andreen et al., 2005; Bjorn et al., 2002). However, the previously held belief that estrogen has a positive effect on memory function and performance has not been confirmed in recent controlled studies. The results of the Women’s Health Initiative Memory Study (WHIMS) revealed that estrogen therapy alone did not reduce dementia or the incidence of mild cognitive impairment; on the contrary, there was an increased risk from estrogen plus progestin therapies (Shumaker et al., 2004; Shumaker et al., 2003).

1.4 TOLERANCE It is known that regular drug users and alcohol drinkers generally become able to tolerate larger amounts of drug or ethanol after repeated exposure. Pharmacologically, tolerance is defined as a decrease over time in the ability of the drug to produce the same degree of pharmacological effect. Tolerance is classified with respect to: 1) possible mechanisms (pharmacodynamic, or pharmacokinetic) and 2) rate of development (acute, chronic, or rapid) (Kalant et al., 1971). Pharmacodynamic tolerance (functional tolerance, or receptor site tolerance) includes reductions in the sensitivity of the target tissue to the same degree of drug exposure. This type of tolerance usually tends to reduce the clinical drug effects, as the duration of drug exposure proceeds and behavioural tolerance is generally of this type. On the other hand, pharmacokinetic tolerance (dispositional or metabolic tolerance) comprises changes in drug absorption, distribution, and metabolism, which lead to reductions in the concentration and duration of drugs in the target tissues. The term usually describes the phenomenon of increased drug clearance. In acute tolerance, there is a reduction in sensitivity to a drug within the duration of a single exposure to the drug (intra-sessional), while, in chronic tolerance, a decrease in sensitivity to a drug occurs after multiple exposures to the drug (inter-sessional). “Rapid tolerance” is a term suggested by Kalant and colleagues and is described as a decrease in sensitivity to a drug after the duration of a single exposure to that drug (Kalant, 1993). A distinction is also made between tolerance to a specific drug, to related drugs (selective or homologous cross-tolerance), or to drugs from unrelated drug classes

24

(heterologous cross-tolerance) (Greenblatt and Shader, 1978; Kalant et al., 1971).

1.4.1 The GABA-A receptor and tolerance The role of the GABA-A receptors in the development of tolerance to positive modulator substances is mostly investigated using benzodiazepines (BZPs) and ethanol. Both substances produce effects by interacting with the GABA-A receptors. BZPs are widely prescribed for the treatment of disorders, such as anxiety and , but tolerance to their effects develops rapidly. After a long period of BZP usage, withdrawal symptoms can also appear following the cessation of the intake of the drug (Harris et al., 1992; Roy-Byrne, 2005). Withdrawal symptoms take place when there is a decline in the blood and tissue concentration of BZP or another dependence-forming substance to which an individual has been continuously exposed. Symptoms of BZP withdrawal are time limited (1-2 weeks), but the duration varies according to the drug and the individual (Roy-Byrne, 2005). The first report of acute tolerance was in relation to ethanol (Mellanby, 1919). Acute tolerance development has also been reported for many other GABA modulatory substances, such as barbiturates and BZPs in both humans and animals, and for the neurosteroid allopregnanolone in rats (Bolander and Wahlstrom, 1988; Dripps et al., 1956; Ihmsen et al., 2004; Kroboth et al., 1993; Zhu et al., 2004). The time needed for tolerance to develop varies. One day of continuous intracerebroventricular exposure reduces the duration of the loss of righting reflex, but tolerance to the hypothermic effects of thiopental and progressed after seven days (Saunders et al., 1990). Increased sensitivity to -induced seizures was first observed after three days of pentobarbital exposure. The same study proposed that the acute effects of barbiturates on the GABA-A receptor are reversible and the pharmacokinetic and pharmacodynamic tolerances are separable (Saunders et al., 1990). It is well established that, following chronic exposure to benzodiazepines and alcohol, there are changes in GABAergic . These changes contribute to the symptoms of tolerance, dependence and withdrawal. However, the nature and mechanism of these changes are not clear (Allison and Pratt, 2003; Kan et al., 2004; O'Brien C, 2005). The treatment of cortical neurons in cultures with benzodiazepines results in GABA-A receptor down-regulation. Although the exact molecular mechanisms responsible for the changes in GABA-A receptor function induced by the positive modulator exposure remain unclear, post-

25

translational modifications, receptor trafficking, effects on receptor density and changes in subunit expression have been shown (Devaud et al., 1997; Grobin et al., 1998; Kumar et al., 2002). GABA-A receptor down-regulation is supposed to take place in the following order; desensitisation (tachyphylaxis) that is characterised by decreased receptor currents during continuous GABA exposure, the sequestration (endocytosis) of subunit polypeptides and the uncoupling of allosteric interactions between the GABA and benzodiazepine binding sites, subunit polypeptide degradation and the repression of subunit gene expression (Barnes, 1996). Consequently, the end point of these changes can be observed as changes in GABA-A receptor subunit mRNA expression and finally BZP tolerant condition.

1.4.2 Changes in the GABA-A receptor during tolerance Chronic treatment with positive allosteroic modulators that act at different sites of the GABA-A receptor results in changes in the sensitivity and expression of the receptor subunit mRNA (Holt et al., 1996; Mhatre and Ticku, 1992; Morrow et al., 1990a). In rats, seven days of flunitrazepam treatment induces tolerance and quantitative in situ hybridisation showed that the α1 and β3 subunit mRNAs are decreased, while the β2 subunit mRNA increased in the hippocampus of the tolerant animals (Tietz et al., 1999). In another study, seven days of treatment increased the α4, β1 and γ3 subunit mRNA levels and these changes are sustained at 14 days’ treatment, along with increases in α3 and α5 and a decrease in γ2 subunit mRNA (Holt et al., 1996). The treatment of rats with diazepam continuously (s.c.) for three weeks reduced the α1 subunit mRNA level in the hippocampus but not in the cortex or cerebellum. The α5 subunit mRNA level was reduced in the cerebral cortex and hippocampus, while the γ2 subunit mRNA level only decreased in the cortex (Wu et al., 1994). The acute sedative effect of diazepam is mediated via the GABA-A receptor α1 subunit, while it is suggested that tolerance to the sedative action of diazepam is associated with the decreased α5 subunit level in the hippocampal dentate gyrus (van Rijnsoever et al., 2004). These studies revealed that, for the induction of tolerance, the chronic activation of the GABA-A receptors by diazepam is sufficient and changes in subunit mRNAs could possibly account for the mechanism of tolerance development. However, tolerance develops to some, but not all, behavioural actions of BZPs. As mentioned above, tolerance to the anxiolytic action of BZP can develop, but little or no tolerance develops to the anterograde amnesia after chronic treatment in humans or animals (Ghoneim et al., 1981; Hughes et al., 1984). In rats, it has been shown that tolerance can develop to

26

the amnesic effect of diazepam on spatial learning using the Morris Water Maze (MWM) test. The authors revealed that the amnesic effect of diazepam is not secondary to sedation and that the amnesic and sedative effects are independent factors (McNamara and Skelton, 1997). The development of ethanol tolerance and dependence is associated with alterations in many properties of the GABA-A receptors throughout the brain and changes the expression of distinct GABA-A receptor subunit mRNAs and in various brain regions. The levels of mRNA for the GABA-A receptor α1, α2, and α3 subunits are reduced in the cerebral cortex, while the α4, β1, β2, β3, and γ1 subunit peptide and mRNA levels are increased in the cerebral cortex following chronic ethanol exposure (Devaud et al., 1997; Mahmoudi et al., 1997; Matthews et al., 1998; Mhatre and Ticku, 1992). Phosphorylation is also proposed as a mechanism that underlies tolerance development. Although the molecular mechanism is not understood, it is clear that the interaction of positive modulator substances with the GABA-A receptor can be regulated by the relative activity of kinases and phosphatases (Brussaard and Koksma, 2002; Koksma et al., 2003; Song and Messing, 2005).

1.4.3 Tolerance to allopregnanolone Repeated exposure to allopregnanolone results in the development of tolerance to the anticonvulsant effect but not to the somnogenic effect (Czlonkowska et al., 2001; Damianisch et al., 2001). Another study proposed genotype-dependent differences in the mechanism of allopregnanolone tolerance (Palmer et al., 2002). Long-term exposure of cultured cortical neurons to allopregnanolone produces a down-regulation of the GABA and TBPS binding sites and the uncoupling of the GABA-A receptor complex (Yu and Ticku, 1995b). This is associated with a reduction in the efficacy of the GABA-induced chloride ion influx and reduced efficacy of other positive modulators. In these studies, chronic allopregnanolone treatment reduced the α2, and α3, β2 and β3 subunit mRNA levels but did not alter the γ2 subunit mRNA level in cortical neurons (Yu et al., 1996a; Yu et al., 1996b; Yu and Ticku, 1995a; 1995b). Short-term exposure to allopregnanolone and the discontinuation of neurosteroid after long-term exposure (sudden decrease in the neurosteroid level) results in the increased expression of the GABA-A receptor α4 subunit mRNA (Follesa et al., 2000; Gulinello et al., 2001). The increased expression of the α4 subunit correlates with increased anxiety and the reduced sensitivity of the GABA-A receptors to classical benzodiazepine

27

agonists and to , as well as a distinct effect by and other positive and negative modulators (Barnard et al., 1998; Gulinello et al., 2001). Our laboratory has used electroencephalography (EEG) to show that acute allopregnanolone tolerance can develop after just 60 minutes of i.v infusion. Using this threshold method, tolerance was accompanied by increased allopregnanolone concentrations in the hippocampus and the brain stem (Zhu et al., 2004). To the best of our knowledge, there are no earlier studies of GABA-A receptor subunit mRNA expression in acute tolerance to allopregnanolone or any other GABA-A receptor active substance.

1.5 LEARNING AND MEMORY Learning is defined as the acquisition of information or skills, while memory refers to the process of recalling what has been learnt (Gazzaniga et al., 2002). Learning is subdivided into several types, such as simple learning (perceptual learning), associative learning (conditioned learning or stimulus- response learning) and more complex learning such as spatial learning (Gazzaniga et al., 2002). Different brain systems are involved in different types of learning and memory. In rodents, lesions of the hippocampus impair the learning of spatial tasks, in addition to many other abilities, e.g. a subset of configural tasks and working memory tasks (O'Keefe and Nadal, 1978; Olton and Papas, 1979; Scoville and Milner, 1957). In women, cognitive disturbances are reported when a high progesterone level is present in the brain (Brett and Baxendale, 2001). It is also known that progesterone has potent anaesthetic properties and a dampening effect on brain excitability (Backstrom et al., 1984). These rapid negative effects of progesterone are mediated by the neuroactive metabolite allopregnanolone (Paul and Purdy, 1992). As a result, the primary action of progesterone on memory and learning may be due to its metabolite allopregnanolone, which has a highly specific action and increases neuronal inhibition via the GABA-A receptor.

1.5.1 Spatial learning and memory Spatial learning is a behavioural task which requires the animal or person to use multiple pieces of information simultaneously. The hippocampus, striatum, basal forebrain, cerebellum and neocortex are brain regions that are known to be involved in this learning (Cain and Saucier, 1996). The Morris Water Maze (MWM) is one of the most frequently used tests for studies of spatial learning; it takes advantage of the fact that rats are good swimmers

28

(Cain and Saucier, 1996; Morris, 1984). The technique of this test is described in detail in the method part of the current thesis. Several neurotransmitters and receptor systems are proposed in the learning and memory mechanisms, with the percentage of significance calculated as follows: glutamatergic (93%), GABA (81%), (81%), (81%), (55%) and nor-epinephrine (48%) (Myhrer, 2003). GABA is suggested as a critical neurotransmitter with an impact on performance in the water maze task (Myhrer, 2003). GABA, and biogenic amines are thought to be either harmful or not to have any effect on this kind of learning (McNamara and Skelton, 1993; Myhrer, 2003).

1.5.2 The GABA-A receptor modulators and spatial learning and memory GABA-A antagonistic steroids have been found to enhance learning and memory processes, while GABA-A agonistic steroids are most often found to impair these processes (Darnaudery et al., 2000; Johansson et al., 2002; Mayo et al., 1993; Reddy and Kulkarni, 1998) GABA-A receptor active substances, like benzodiazepines (BZP), are widely used as sedative, anxiolytic and anticonvulsant drugs, but unwanted memory deficits frequently occur during treatment. These effects are mediated by the activation of the GABA-A receptor, e.g. in the hippocampus, and the enhancement of inhibitory GABAergic transmission (Barker et al., 2004; Macdonald and Barker, 1978; McNamara et al., 1993). The stimulation of GABA-A receptors with BZPs reduces the neuronal firing of glutamate neurons, which correlates with the spatial memory impairment induced by BZPs (Riedel, 1996). BZP induces GABA-A receptor-mediated synaptic inhibition and affects the NMDA receptors and Ca+ influx. In other words, an interactive mechanism may conduct and regulate the effect of BZPs on spatial learning, by the concurrent affection of the NMDA and glutamate receptors (McNamara et al., 1993). It is known that the excitability of the hippocampal neurons is influenced by powerful inputs that in turn have an effect on cognition. Allopregnanolone participates in the GABAergic modulation of the central cholinergic function, (Landgren et al., 1998) and inhibits basal acetylcholine release from the prefrontal cortex and hippocampus, but not from the striatum, in a dose-dependent manner (Dazzi et al., 1996). Gonadal hormones influence the brain organisation and a variety of reproductive or non-reproductive behaviours, including some cognitive abilities (Williams and Meck, 1991). Dendritic spines are believed to be important sites for the enhancement of synaptic efficacy and networks involved in learning and neural plasticity (Hering and Sheng, 2001). The

29

dendritic spine density in the hippocampus declines slowly as estrogen levels fall, a decline that is much more rapid if progesterone is administered (Woolley and McEwen, 1993). This rapid down-regulation of dendritic spines is probably due to the action of progesterone at the progesterone receptor (PR), as this effect can be inhibited by the progesterone antagonist RU 38486 that exerts its effect on PR (Woolley and McEwen, 1993). There is controversy about the effect of allopregnanolone on learning and memory, but in the majority of studies allopregnanolone impairs spatial learning and memory in animals. It is even possible to observe a progressive memory impairment, a few minutes after an injection of allopregnanolone, due to a robust increase in allopregnanolone levels in both the brain and plasma of rats (Freeman et al., 1993; Johansson et al., 2002; Ladurelle et al., 2000; Paul and Purdy, 1992). The mechanism of the effect of allopregnanolone on memory and learning is not yet clear. Apparently, however, this neuroactive steroid act as a positive of the GABA-A receptor and, like BZPs, increases the frequency and duration of the GABA-gated chloride channel opening. The dose-dependent potentiation of the GABA-evoked Cl¯ currents is expressed as an inhibitory effect on neurons (Paul and Purdy, 1992), especially in the hippocampus, one of the major brain areas involved in learning (Landgren et al., 1998). In rodents, the activity of individual hippocampal pyramidal cells is correlated with the animals’ location in the environment (called place cells) (Muller et al., 1987; O'Keefe and Dostrovsky, 1971). The application of allopregnanolone to the hippocampus prolongs the GABA-mediated inhibitory postsynaptic currents from hippocampal neurons, altering the hippocampal neurophysiology with the inhibition of hippocampal pyramidal neurons (Harrison et al., 1987; Landgren et al., 1998; Tokunaga et al., 2003). Acute allopregnanolone administration to rats, and consequent changes in hippocampal neurobiology, selectively impairs the use of spatial reference memory, while saving the non-spatial memory (Matthews et al., 2002). The selectivity of allopregnanolone in relation to spatial memory tasks is similar to the deficits found following acute ethanol administration (Matthews and Morrow, 2000).

30

2. AIMS

1. To test whether a 3β-hydroxy pregnane steroid can block the negative effects of allopregnanolone on learning in the Morris Water Maze test (Paper I).

2. To study the mechanism of the 3β-hydroxy pregnane steroid antagonism on the negative effects of allopregnanolone on learning (Paper I).

3. To evaluate whether chronic tolerance develops to the negative effects of allopregnanolone on learning (Paper II).

4. To study the mechanism behind acute allopregnanolone tolerance, by quantifying GABA-A receptor subunit mRNA expression in different brain areas (Paper III).

5. To evaluate whether acute tolerance induced by allopregnanolone anaesthesia persists after the end of treatment and whether persistence is related to changes in GABA-A subunit mRNA expression (Paper IV).

31

3. METHODS

3.1 In situ hybridisation This technique can be used to determine the location and for quantification of the expression level of specific RNA or DNA in tissues, cells or chromosomes using specific labelled nucleic acid probes. The nucleic acid hybridisation techniques are based on the fact that, in appropriate conditions, complementary sequences of single-stranded nucleic acid species spontaneously anneal to form a double-stranded hybrid (duplex). To study the distribution or level of expression of specific or gene transcripts, isotopic or non-isotopic markers (enzymatic or photo-chemical) can be incorporated into the probe (Wilkinson, 1995). The sensitivity of in situ hybridisation exceeds that of Southern and Northern blotting techniques, as the target mRNA is then diluted by the relatively large-scale homogenisation of tissue. With in situ hybridisation, mRNA species present at low levels and expressed in few cells can be analysed. In our studies, oligonucleotide probes labelled with the 35S-isotope were used for the determination of the GABA-A receptor α1, α2, α4, α5, α6, β2, γ2(L+S) and δ subunit mRNA expression (Wisden et al., 1992). Denatured oligonucleotide probes for the subunit in 100 μl of hybridisation were used on each slide containing three consecutive fixated sections of half the brain from the same animal. Slides were hybridised over night, followed by washes, and they were then exposed to Hyperfilm Biomax MR-1 film for two to 12 days. The slides were then dipped in Kodak NTB2 photoemulsion, exposed at +4ºC for 0.5 to six months, after which sections were counterstained with Pyronin-Y. Using a special image analysing system (MCID M4 analysing system- Imaging Research Inc. Ontario, Canada), we counted the silver grains in fixed microscopic areas. Counting was performed in blind conditions.

3.2 Immunohistochemistry This technique is a combination of microscopy and immunology, in which an antibody with very specific binding for its antigen is used and the reaction site can be detected by attaching a specific probe to the antigen-antibody complex. Using this technique, antigens in any position within the cell can be identified. There are numerous immunohistochemical techniques that can be used to localise antigens. The selection of a suitable technique is based on parameters such as the type of specimen under investigation and the degree of sensitivity required (Beesley, 1993).

32

We used a Streptoavidin-Biotin immunohistochemistry method for the GABA-A receptor α4 subunit protein (Paper III). Sections were fixated and blocked with normal swine serum. This was followed by incubation with the primary antibody (1:20 goat anti rat α4) for two hours. Subsequently, slides were incubated with secondary antibody (swine anti-goat-biotinylated substance), followed by avidin/HRP. Sections were then incubated with diaminobenzidine/H2O2 and counterstained with Meyer’s hematoxylin. Usual problems, such as unspecific labelling (background), too weak signals (method problem) and wrong results (specificity) (Beesley, 1993), were overcome by changing the type of blocking substance and incubation time with the primary antibody. Using the x40 magnification lens of the light microscope (BX51M- Olympus, Japan), positively stained cells containing the α4 subunit were counted in three distinct random fields of the ventral posteriomedial nucleus region of the thalamus in each of the three sections. The average number of positive cells per slide (animal) was calculated by dividing the sum of cells by the number of areas that were analysed.

3.3 Chloride ion uptake assay Allopregnanolone potentiates the GABA-induced chloride ion influx through the GABA-A receptor (Majewska et al., 1986). The chloride ion uptake assay can be used to determine the GABA-A receptor function and modulation by drugs. Briefly, in this study (Paper 1), microsacs were prepared by the homogenisation of cortical or hippocampal tissue + buffer in a glass-Teflon homogeniser, filtration through a nylon mesh and purification by centrifugation (as described by Lundgren et al., 2003). A Skatron (Lier, Norway) twelve-channel cell harvester and 96-well microplates were used to perform the chloride ion uptake assays. Wells were filled by adding steroids, GABA, buffer and [36Cl-]. By adding homogenate to the wells, the up-take reaction was started and, after five seconds, the reaction was terminated by washes with buffer containing picrotoxin. After rapid filtration of the well content through a fibreglass filter, captured [36Cl-] was measured with a liquid scintillation spectroscope. Each chloride influx test was performed in quadruplicate determinations. The data are expressed as net chloride ion uptake (i.e. minus the basal uptake in the homogenate).

3.4 EEG threshold method In Papers III and IV, acute allopregnanolone tolerance and changes in CNS sensitivity to allopregnanolone were evaluated with the electroencephalographic (EEG) burst suppression threshold test. This

33

method is constituted on the induction of a burst suppression of one second or more in the EEG (silent second, SS) by the i.v. infusion of an anaesthetic drug. The dose needed to induce the criterion of a “silent second” (SS) is defined as the threshold dose of the anaesthetic (Fig. 3).

After 1-2 min infusion

Gradually decreased frequency

First Silent Second after 3-5 min

Figure 3: The EEG of a rat before and during continuous infusion of an anaesthetic drug (adapted from (Korkmaz and Wahlstrom, 1997).

The period of burst suppressions (including SS) is a reversible CNS that is induced by the test drug. Factors like the chemical properties of the drug and vehicle (e.g. water ) and age of the animals affect the results of the test. To minimise the infusion volume and to follow the time, an infusion volume rate is calculated. The optimal infusion rate must be used, i.e. the rate at which the lowest dose is needed to obtain an SS; for allopregnanolone, the dose rate is 4 mg/kg/min (Zhu et al., 2001). In addition, some CNS drugs (e.g. ethanol) are not able to induce SS and some drugs (e.g. barbiturates) may induce convulsive-like episodes during the excitatory phase (inhibition of the CNS inhibition) of the induction (Wahlstrom and Norberg, 1984). The EEG was recorded from subcutaneous stainless steel sutures placed in a bifrontal configuration (Korkmaz and Wahlstrom, 1997).

34

In this thesis (Papers III and IV), allopregnanolone was infused i.v. into one of the tail veins with a syringe pump, until the first SS was recorded in the EEG. When an SS was recorded, the infusion was immediately stopped, while the EEG recording continued. Immediately after the first SS detection, the control animals were decapitated. In the other rats, the next infusion of allopregnanolone started when no SS had been recorded for a period of one minute. The method of stopping and re-starting the infusion continued for a fixed time (30 and 90 min) and was then ended at an SS and the rats were immediately decapitated. The allopregnanolone doses needed to maintain anaesthesia over time (maintenance dose rate, MDR, mg/kg/min) were calculated from the accumulated doses during three different intervals of 20 min. The three intervals were 10-30 minutes (MDR1), 40-60 minutes (MDR2) and 70-90 minutes (MDR3). The calculated individual difference between MDR3 and MDR1 (ΔMDR) was used as an in vivo measure of acute tolerance. In Paper IV, rats in two groups were allowed to recover after the 90 minutes of anaesthesia and, one or two days later, the rats were decapitated immediately after a new SS induction with allopregnanolone.

3.5 Morris Water Maze test The Morris Water Maze test was developed by Richard Morris (Morris, 1984) and is one of the most frequently used tests for evaluating the capacity of rodents to learn and retrieve spatially encoded information. In humans, a computerised (virtual) version of the Morris Water Maze (VMWM) test is used for the laboratory testing of spatial ability (Astur et al., 1998; Hamilton et al., 2002). In animals, the Morris Water Maze test can be used in two different spatial tasks; a place-learning task that requires reference memory (remembering the location of the platform from the previous day) and a matching-to-place task that requires both reference memory and working memory (learning and remembering the new location) (Whishaw, 1985). In our studies, we used a place-learning version of the test. Technically, water provides the in-maze environment during the process and escape from water is used for the motivation of learning. Two advantages of the MWM test over other mazes are 1) the rat wants to escape so it searches and 2) in water, there are no local cues (e.g. scent trails). The task requires that the rat swims in a pool until it finds and climbs onto a small platform which is hidden, slightly submerged in the water. In the place-learning version, the platform place is fixed, whereas the platform place in the matching-to-place task is moved to a new location each day (Whishaw, 1985). The animal is placed in the pool at various points near the wall and,

35

to successfully complete the performance the animal must learn the exact location of the hidden platform in relation to the visual extra-maze cues in the room. The spatial memory is measured on a probe trial, in which the platform is removed. Animals that have learnt the position of the platform will spend more time in the area of the pool where the platform was previously placed. The MWM test with a visible platform can also be used for simple (stimulus-response) learning and the examination of visual sharpness (Gerlai, 2001). In Papers I and II, the water maze was a round black pool that was filled with clear tap water. The circular escape platform made of transparent Plexiglas was placed in the middle of the north-east quadrant. To reduce the stress effect of the test, we daily handled and habituated the rats for five days, until two days before the experiment. This included the same procedures as in the experiment, except for injections. Studies of place navigation with a constant location of the hidden platform were performed. Animals received allopregnanolone or a vehicle injection eight minutes before the test. They were then given four trials every day from four different starting positions, which were varied randomly from day to day but were the same for all the rats every day. The training took place for five or six days. The trial numbers per day (four) was optimal, as the effect of allopregnanolone in the brain is short lasting, because of a short half-life, and more trials in one day would have affected the results as the allopregnanolone level would have diminished. We already knew that the allopregnanolone concentrations in the brain tissue at eight minutes were 1.5 to 2.5 times higher than at 20 minutes after the allopregnanolone injections in the MWM test (Johansson et al., 2002). In each trial, the rats had a maximum of 120 s to search for the platform. Rats that did not find the platform were guided there by hand. All the rats were allowed to sit on the platform for 30 s and were then picked up and dried with a towel for 30 s, until the start of the next trial. The performance of rats was evaluated with an overhead camera connected to an image analyser (HVS imaging, Hampton, UK) with the water maze software program HVS Water 2020. The parameters latency, path length to reach the platform, swimming speed and thigmotaxis are prominent factors for evaluating the learning condition and the motor ability in the MWM test. Latency and path length were used as representatives of learning, while swimming speed was used as an index of changes in spontaneous locomotor activity and sedation (Gallagher et al., 1993; Lindner, 1997; Morris, 1984). It has been suggested that high thigmotaxis shows a low ability to engage in the procedural part of the

36

navigational strategies in the MWM, but it can also be seen as a sign of fear and anxiety (Devan et al., 1999; Simon et al., 1994). In Paper II, to test spatial memory in rats, a probe test was carried out one day after the last training session (day 6), without any injections before the test. The rats swam for 60 s with no platform present and the time spent in the different quadrants was analysed.

3.6 Hormone assays Allopregnanolone in serum (Paper III) or plasma (Paper IV) was extracted with and the extracts then evaporated to dryness under . Neurosteroids in tissues were extracted by ethanol (99.5%) for seven days at +4ºC. The recovery of steroids from the tissue using this procedure has previously been shown to be 100% (Bixo et al., 1984). Plasma and serum samples were analysed in triplicate and brain samples were analysed in duplicate. Allopregnanolone in extracts was purified by either preparative high performance liquid chromatography (HPLC) (Papers I and IV) or celite chromatography (CC) (Paper III) and measured by radioimmunoassay (RIA). The allopregnanolone antiserum used for RIA analyses was raised against 3α-hydroxy-20-oxo-5α-pregnane-11-yl-carboxymethyl ether coupled to bovine . The sensitivity of the assay was 25 pg, the intra- assay coefficient of variation 7% and the inter-assay coefficient of variation 8% (Purdy et al., 1990a). Evaluations of the repeatability, reproducibility and sensitivity of the HPLC purification were tested using aliquots spikes with a known amount of steroid, 100, 200 and 300 pg (0.03, 0.06 and 0.12 pmol). The recovery of steroid after the HPLC separation was 79% for brain tissue and 72% for plasma, while the recovery was 85% for brain tissue and 78% for plasma with CC. The test samples were therefore corrected for this loss. Our comparison of the two methods revealed that the allopregnanolone concentration that was obtained was the same with both chromatographic methods (unpublished data).

3.7 Statistical analysis Repeated measures analysis of variance (ANOVA) was used when the data from individual rats were obtained several times. With significant results (p≤0.05), the Least Significant Differences (LSD) post-hoc test was used. In the MWM test, the factors were days by treatment (groups). Group parameters found to be significantly different (ANOVA+LSD, p≤0.05) were further analysed two by two with the Mann-Whitney U-test to identify the days on which the groups were significantly different. The dose-response

37

data were fitted to the sigmoidal dose-response equation: Y = Minimal response + (Maximum response – Minimal response)/(1+10^((LogEC50- X)*HillSlope). The independent factors were the concentration of enhancing substance (GABA, allopregnanolone) and the concentration of antagonising substance (0–30 µM UC1011) when the chloride ion uptake parameters were evaluated. In the EEG-threshold studies, the between-subject factors were groups and the time of onset of anaesthesia (i.e. am, pm and mid-time in Paper IV). To compare the mean value of a variable with independent observations (between groups), we used one-way ANOVA (number of groups >2) or Student’s t-test (number of groups =2), e.g. MDRs, ΔMDR, SSs and ΔSS. However, to determine how responses were affected by the two independent factors treatment and time of onset of anaesthesia, we used two-way ANOVA. Linear regression coefficients (b) and parametric correlation coefficients (r) were used to evaluate simple statistical relationships in the data set. For Paper III, the chi-square test was used to test significance in frequency tables. The significance level used was p ≤ 0.05. All the values in the text and tables are displayed as means ± SEM. The SPSS statistical package versions 10.0 and 11.0 were used.

38

4. RESULTS AND DISCUSSIONS

The current thesis investigated changes in the sensitivity (tolerance) and plasticity of the GABA-A receptors and the opportunity to inhibit negative behavioural effects induced by allopregnanolone.

4.1 Antagonism of an allopregnanolone effect, Paper I To evaluate the 3β-hydroxypregnane steroid UC1011 (3β-20β-dihydroxy- 5α-pregnane) as a blocker of allopregnanolone-induced effects, both the in vivo MWM test of learning and the in vitro chloride ion uptake into cortical and hippocampal membrane preparations (microsacs) were used. The rats were injected every day either with UC1011 20 mg/kg, allopregnanolone 2 mg/kg, allopregnanolone: UC1011 2:6, 2:8 and 2:20 mg/kg, or with vehicle (10% 2-hydroxypropyl-β-). As the MWM test is hippocampus dependent, we analysed the effect of UC1011 with both cortex- and hippocampus-derived membranes. The increase in chloride ion uptake by the cortical and hippocampal microsacs induced by allopregnanolone was antagonised by the UC1011 dosages of 10 and 30 µM. UC1011 in a concentration of 3-30 µM did not display any significant inhibitory effect on the chloride ion uptake induced by 10 µM GABA alone.

A B 220 200 1000 nM Allo 180 190 160 160 1000 nM Allo uptake uptake -

- 140 Cl

Cl 150 nM Allo (% of control) 130 (% of control) 120

100 100 0 1 3 10 30 Max 0 1 3 10 30 Max [UC1011] (µM) [UC1011] (µM)

Figure 4: The antagonistic effect of UC1011 on two concentrations of allopregnanolone in cortical microsacs (A) and of one concentration of allopregnanolone in hippocampal microsacs (B). The inhibition of allopregnanolone-stimulated chloride ion uptake was performed in the presence of 10 µM of GABA and different concentrations of UC1011. The chloride ion uptake in the presence of 10 μM of GABA was taken as 100%.

39

In Figure 4, the antagonistic effect of UC1011 against two concentrations of allopregnanolone with cortical microsacs and of one concentration of allopregnanolone with hippocampal microsacs is shown. These studies showed that UC1011 significantly reduces the allopregnanolone-induced increase in chloride ion uptake seen in the presence of GABA in both the brain regions that were analysed.

Benzodiazepines are also positive modulators of the GABA-A receptor that, like allopregnanolone, have negative effects on memory and learning and these effects are mediated by the activation of BZP-specific sites within the GABA-A receptor complex (Lister, 1985). The enhancement of spatial learning by the benzodiazepine receptor antagonists and inverse agonists helped to provide an understanding of this mechanism (Raffalli-Sebille and Chapouthier, 1991). Allopregnanolone reduces spatial learning in the MWM test (Johansson et al., 2002). Most of the animals injected with allopregnanolone eight minutes before the MWM test preferred to swim close to the pool wall (thigmotaxis); this effect is not caused by the impairment of the motor function, as the swimming speed is unaffected (Johansson et al., 2002). Using the parameters escape latency (i.e. time to reach the platform, in seconds), path length (cm), swimming speed (cm/s) and thigmotaxis, the performance of rats in the MWM test was evaluated for all the training days. 6 or 8 mg UC1011/kg injected, together with allopregnanolone (2 mg/kg) eight minutes before the swimming test, did not improve learning compared with allopregnanolone alone (data not shown). However, with an increased dosage of UC1011 (20 mg/kg), it was possible to reduce the negative effects on learning that were induced by allopregnanolone in the MWM test. In rats injected with the mixture of allopregnanolone and UC1011 (2:20 mg/kg, respectively), the parameters latency and thigmotaxis were significantly reduced (both p<0.05) on days 3-6, compared with the allopregnanolone- injected group (Fig. 5). The improved swimming performance was not due to improved motor performance, as there was no significant difference in swimming speed between the groups. In the MWM test, learning that there is an opportunity to escape from the water is a strong incentive factor that motivates the animals to leave the pool wall. This condition is achieved by searching the pool and acquiring the information about the maze. Only after leaving the wall can a true spatial learning of the platform position take place (Morris, 1984). As a result, the allopregnanolone-treated rats are stuck in an early phase of the task.

40

A B 120 100

100 Allopregnanolone 80 80 Allo:UC1011 Vehicle 60 (s) 60 UC1011 (%)

latency 40 40 thigmotaxis 20 20 0 0 0 1 2 3 4 5 6 0 1 2 3 4 5 6 Day Day

Figure 5: MWM test. Allopregnanolone (2 mg/kg), allopregnanolone and UC1011 (2:20 mg/kg respectively), UC1011 (20 mg/kg), or vehicle was injected (i.v) eight minutes before the start of each swimming session. The data shown are mean ± SEM for all four swimming trials each day A) Latency – time to find the hidden platform. B) Thigmotaxis – the fraction of the total path during which the rats swam close to the pool wall.

In the group injected with allopregnanolone and UC1011 (2:20 mg/kg) the concentration of allopregnanolone was lower in the investigated brain regions than the levels in the group only injected with allopregnanolone. The differences were statistically significant in the striatum and cortex. We know that high concentrations of allopregnanolone within the brain are needed to produce severe negative effects on learning. In an earlier study, we found no learning deficit 20 min after the injection, vs. vehicle-injected rats, and the allopregnanolone concentration in different brain regions was then decreased by 30-60% compared with the levels eight minutes after the injection (Johansson et al., 2002). In the present study, we found decreased levels of allopregnanolone within the cortex and striatum, but not in the hippocampus, after injections of allopregnanolone together with UC1011 (20 mg/kg). Changes in the brain concentration of allopregnanolone might be related to the inhibitory effects of UC1011 on allopregnanolone and GABA-A receptor binding and/or changes in the allopregnanolone metabolism that resulting in changes in allopregnanolone concentration in the brain tissue. In spite of this, we know that 3β- are not direct antagonists of 3α- hydroxypregnane steroids, such as allopregnanolone, but are instead non- competitive, probably state-dependent functional inhibitors of GABAA receptor potentiation (Wang et al., 2002). The present study shows that UC1011 can reduce the harmful effect of allopregnanolone on learning. However, to produce a positive effect byUC1011, we needed a ten times higher concentration of the drug than the concentration of allopregnanolone that was infused. It is probable that the

41

allopregnanolone-induced decrease in learning is caused by increased GABA-A activation and that the reduction of the allopregnanolone effect by UC1011depends on reduced GABA-A activation. The effects of allopregnanolone on cognitive functions may be the cause of cognitive disturbances noted by women with premenstrual dysphoric disorder, as their problems occur during the luteal phase of the menstrual cycle when the allopregnanolone concentration is high (Man et al., 1999). It is therefore essential to find a substance that antagonises the adverse effects of allopregnanolone. We also need more studies to determine whether UC1011 can meet the criteria required for a drug that can inhibit the negative effects produced by allopregnanolone.

4.2 Chronic allopregnanolone tolerance, Paper II To study the cognitive effects of repeated administration of allopregnanolone, two experiments with differences in pre-treatment time and injection method were performed. In Experiment 1, the three-days pre- treatment period consisted of i.v. injections (in the tail vein) with allopregnanolone (2 mg/kg) dissolved in 10% 2-hydroxypropyl-β- cyclodextrin (vehicle) two times a day (the 3xAllo+Allo group), with allopregnanolone (2 mg/kg) in the morning and vehicle (1 ml/kg) in the evening (the 3xAllo once+Allo group), or with vehicle (1 ml/kg) twice a day (the 3xVeh+Allo and 3xVeh+Veh groups). In Experiment 2, the rats were injected twice a day during the seven-day pre-treatment period (i.p.) with allopregnanolone (20 mg/kg) dissolved in sesame oil (the 7xAllo+Allo group) or with sesame oil (1.33 ml/kg, the 7xVeh+Allo and 7xVeh+Veh groups). After pre-treatment, the animals were subjected to the MWM test for five consecutive days. In the first experiment, the group pre-treated with allopregnanolone twice a day (3xAllo+Allo) experienced a less negative effect from the allopregnanolone that was injected before swimming. The rats in this allopregnanolone-pretreated group had significantly lower latency and thigmotaxis and shorter path length in comparison with the acute allopregnanolone group (Fig. 6). However, the 3xAllo+Allo group also had significantly higher values for the parameters latency, path length and thigmotaxis than the control group (3xVeh+Veh). The group pre-treated with allopregnanolone once a day (3xAllo once+Allo) did not differ from the acute allopregnanolone-treated group (3xVeh+Allo) and only differed significantly from the control rats.

42

A) B)

Latency Thigmotaxis

120 100 3 x Veh.+Allo 100 3 x Allo once+Allo 75 80 3 x Allo+Allo s

3 x Veh.+Veh. % 60 50

40 25 20

0 0 0 1 2 3 4 5 0 1 2 3 4 5 Day Day

Figure 6: Data from the MWM test after three days of i.v. pretreatment. (A) Latency to escape; (B) Thigmotaxis (path swum close to wall), during the five days of training with a hidden platform. Allopregnanolone 2 mg/kg or vehicle was injected before the four trials each day. The maximum time allowed during one trial was 120 seconds.

Chronic tolerance is defined as a change in the dose-response relationship produced by multiple exposure to a drug (Kalant et al., 1971) and a given dose therefore has less effect as a result of the previous administrations. These findings suggest that pre-treatment once a day for three days was not able to induce tolerance of the acute allopregnanolone effect during the MWM testing. Rats pre-treated with allopregnanolone twice daily (3xAllo+Allo) showed incomplete tolerance and had a swimming performance between that of the 3xVeh+Allo and 3xVeh+Veh groups. The differences in performance might depend on the short treatment duration and the fact that an insufficient steroid concentration was reached in rats injected just once a day. The development of tolerance towards a given drug is dependent on factors like the length of exposure, the dose and the route of delivery of the drug (Kalant et al., 1971). In the next experiment, with longer pre-treatment, the was changed to i.p., since the tail veins are unable to tolerate excessive i.v injections during pre-treatment. However, as in Experiment 1, i.v. injections were given during the MWM test. Sesame oil was used as a vehicle for intra-peritoneal injections. This produces a slow and continuous absorption of the steroid, creating optimal conditions for tolerance development, as the duration of drug exposure is prolonged (Chien, 1981). When allopregnanolone is injected intra-peritoneally, the neurosteroid is taken up into the portal circulation and the metabolism of the drug in the begins before the drug reaches the brain. A larger amount of steroid is therefore required to induce an effect (Holzbauer, 1976). For this reason, the

43

dosage of allopregnanolone was changed accordingly. In Experiment 1, we used a fixed dose of allopregnanolone (2 mg/kg) during both pre-treatment and MWM testing. Previous studies revealed that allopregnanolone at this dosage produces an acute negative effect in the MWM (Johansson et al., 2002). Spatial memory in the water maze has been studied after i.p injections and the dosage of 17 mg of allopregnanolone/kg significantly impaired memory when given i.p. However, memory was significantly poorer when 20 mg of allopregnanolone/kg was used (Matthews et al., 2002). So, in experiment 2, we used the dosage of 20 mg of allopregnanolone /kg i.p. given during pre-treatment. The group that was pre-treated with two i.p. injections of allopregnanolone a day for seven days (the 7xAllo+Allo group) had a significantly better performance than the acute allopregnanolone group (7xVeh+Allo) that was pre-treated with the vehicle. The rats in the 7xAllo+Allo group improved their performance, while the 7xVeh+Allo group did not learn to locate the platform. The 7xAllo+Allo group learnt the task almost as well as the controls (7xVeh+Veh), as they were significantly different only on the first day (p<0.01 for all parameters, Fig. 7). In the MWM test, learning is expected after fewer than 20 trials (in this setting, five days) and the first two days of the test are accepted as acquisition of the escape behaviour (Morris, 1984).

A) Latency B) Thigmotaxis 7 x Veh.+Allo 120 100 7 x Allo+Allo 100 7 x Veh.+Veh. 80 80 ↓ * ↓ ↓ 60 s 60 % ↓ ↓ 40 ↓ 40 * 20 20

0 0 0 1 2 3 4 5 0 1 2 3 4 5 Day Day

Figure 7: Data from the MWM test after seven days of pre-treatment (i.p.). A) Latency to escape and B) Thigmotaxis during the five days of training with a hidden platform. Allopregnanolone 2 mg/kg or vehicle was injected and four trials per rat were performed each day. The maximum time allowed during one trial was 120 seconds.

In this experiment, a probe test was carried out on day 6. The test results confirmed that pre-treatment with allopregnanolone improves the acute negative effects of the neurosteroid in the MWM test. The 7xAllo+Allo group spent an equal amount of time in the quadrant with the former platform position as did the control rats (group 7xVeh+Veh), while the

44

7xVeh+Allo group spent significantly less time in the right quadrant (p<0.05, Fig. 8). Probe test

50 Quad NE (former platform position) 40 Quad SE

30 Quad SW (opposite to former platform) Quad NW 20 % of total time total of % 10

0 7 x Allo+Allo 7 x Veh+Allo 7 x Veh+Veh Groups

Figure 8: Data from the memory test performed in Experiment 2. The rats swam for 60 seconds without any platform being present in the pool.

However, the 7xAllo+Allo group did not recall the platform position as well as the 7xVeh+Veh group of rats. These findings suggest that seven days of pre-treatment induced a clear-cut but no total chronic tolerance to the negative effect of the allopregnanolone given before the MWM test (Fig. 8). In animals, it has previously been shown that tolerance to the anticonvulsant activity of allopregnanolone develops after five days of intra- cerebroventricular repeated administration (Czlonkowska et al., 2001), whereas the effects of allopregnanolone on sleep were not changed after the same treatment time with daily i.p. injections (Damianisch et al., 2001). In fact, it is known that tolerance does not always develop at an identical speed to all the effects of a given drug and the drug exposure length needed for tolerance development can vary depending on the method and dosage (Damianisch et al., 2001; Kokate et al., 1998; LeBlanc et al., 1969; Mendelson, 1964; Mendelson et al., 1964; Wikler, 1956; Wikler et al., 1956). Experience from studies with other substances acting on the GABA- A receptor shows similar time frames for tolerance development as allopregnanolone in the MWM test. For instance, cats and rabbits develop tolerance to the hypnotic effects of some barbiturates after two to seven days’ treatment (Gluckman and Gruber, 1952; Jaffe and Sharpless, 1965), while rats and mice show tolerance after one to five days (Aston, 1965, 1966; Rumke, 1963). In humans, tolerance of the hypnotic action of barbiturates has been reported to occur after three to 8.5 days (Belleville and Fraser, 1957).

45

In this study, we demonstrated for the first time that repeated allopregnanolone treatment, given over several days, can generate tolerance towards the spatial learning impairment produced by acute allopregnanolone administration.

4.3 Acute allopregnanolone tolerance and the GABA-A receptor subunits, Paper III Possible changes in GABA-A receptor subunit mRNA expression levels during the development of acute allopregnanolone tolerance were studied. Allopregnanolone (3 mg/ml dissolved in 10% hydroxypropyl-β- cyclodextrin) was infused (i.v.) during EEG recording. After different time intervals (first SS, 30 min or 90 min of anaesthesia), the last infusion period to SS was followed by decapitation. Samples from blood, different brain regions, muscle and fat tissue were analysed. Half the brain was used for the determination of GABA-A receptor subunit protein and mRNA expression levels.

100 90 30 min 80 90 min 70 60 50

(mg/kg) 40 30

Cumulative dose 20 10 MDR1 MDR2 MDR3 0 0 10 20 30 40 50 60 70 80 90 100 Duration of anaesthesia (min)

Figure 9: Cumulative allopregnanolone doses given to maintain anaesthesia at the silent second level in the 30-min and 90-min groups. A theoretical straight line is shown for the cumulative doses without tolerance development. The time periods used for calculating maintenance dose rates (MDR) are indicated.

The development of acute tolerance to allopregnanolone was detected using the EEG-threshold anaesthesia method (Korkmaz and Wahlstrom, 1997). The initial sensitivity to allopregnanolone (threshold dose) for the first silent second induction did not differ between the three experimental groups (control, 30-min and 90-min groups). In the 90-minute group, the dose rate of allopregnanolone needed to maintain anaesthesia (MDR) was significantly (p<0.001) increased during the time period 65-85 min (MDR3) compared with the period 10-30 min (MDR1) and the period 35-55 min

46

(MDR2, Fig. 9), indicating that acute tolerance had developed in the MDR3 period. The difference between MDR3 and MDR1 (ΔMDR) is a direct in vivo measurement of the magnitude of acute tolerance, since this shows the increase in allopregnanolone dose that is needed to maintain SS anaesthesia. The allopregnanolone concentrations in certain regions of the brain in the 90-min group were significantly higher than in the SS group (Fig. 10). These regions were the hippocampus, MTH (midbrain, thalamus and hypothalamus) and MOP (medulla oblongata and pons). Allopregnanolone concentrations in serum and fat in the 90-min group and in fat in the 30-min group were higher when compared with the SS group (Fig. 10).

160 * † 120 SS (control) mol/l)

μ 30 min 80 * 90 min † * * * † (nmol/g or Allopregnanolone 40 *

0 . m H ex u ala . o fat d b llum scle hipp MT MOP g e serumcort mu striat my a ereb c Figure 10: Allopregnanolone concentrations in serum, different brain areas, muscle and fat tissue in the different groups. * – significant difference versus the controls in group; SS  – significant difference versus 30-min group. hipp = Hippocampus; MTH = Midbrain, thalamus and hypothalamus; MOP = Medulla oblongata and Pons; b.o. = Olfactory bulb.

In a previous study, acute tolerance to produced increased drug concentrations in the cortex and brainstem as early as at 30 min of anaesthesia vs. controls (Bolander and Wahlstrom, 1988). In old rats but not in young ones, brain concentrations increased compared with controls after 60 min of anaesthesia in the hippocampus, striatum, brainstem and cerebellum, but not in the cortex (Larsson and Wahlstrom, 1996). Substance accumulation in fat tissue has been observed in all the studies in which acute tolerance has been detected using the EEG-threshold method. From the allopregnanolone concentration data, the hippocampus, thalamus and hypothalamus were found to be the most important brain regions for the development of acute allopregnanolone tolerance. We therefore performed

47

in situ hybridisation in order to analyse mRNA expression in the areas that are most probably important for tolerance development. GABA-A receptor subunit expression patterns were similar to those that have been found before, with only slight differences from Wisden (Wisden et al., 1992).

A) B)

α4 protein * * 40 50 38 ) 36

4 mRNA4 VPM 34

α 25 / X40 area grains/cell ( 32

GABA-A 30 number of positive cells 28 0 0.0 0.1 0.2 0.3 0.4 0.5 Control 30 min 90 min Allopregnanolone dose difference (ΔMDR)

Figure 11: A) Linear regression plot for the GABA-A receptor α4 mRNA expression in the ventral posteriomedial nucleus of the thalamus and the ΔMDR (increase in allopregnanolone dose per minute needed to maintain the anaesthesia; mg/kg/min) in the rats in the 90-min group. B) The number of GABA-A receptor α4 protein positive cells (mean ± S.E.M in the ventral posteriomedial nucleus of the thalamus, in X40 magnification areas from the light microscope. * p < 0.05 vs. the groups’ SS.

A statistically significant decrease in the GABA-A receptor α4 subunit mRNA and protein level was detected in the thalamic VPM nucleus in the tolerant 90-min group when compared with the SS group (post-hoc, p<0.05) and the 30-min group (post-hoc, p<0.05; Fig. 11). There was also a negative correlation between the ΔMDR and the expression of the GABA-A receptor α4 subunit in the VPM of the thalamus (r = -0.74, b = -20.76, d.f. = 7, p<0.05). This means that greater tolerance developed in the animals with the smallest amount of the GABA-A receptor α4 subunit in the VPM of the thalamus. It has been shown that the α4 subunit containing GABA-A receptors is highly plastic and is rapidly altered in response to changes in neuronal activity (Sur et al., 1999). The change in the α4 subunit mRNA and protein levels and the increase in allopregnanolone levels in the thalamus region indicates that this part of the brain is involved in the development of acute allopregnanolone anaesthesia tolerance. The thalamus has often been called “the switchboard of the brain”. This part of the brain plays an

48

important role in modulating and synchronising the signal traffic between regions and is probably also involved in anaesthesia and tolerance of GABA-A receptor active anaesthetic compounds (Reynolds et al., 2003). In line with our results, there are data which indicate that a four-day application of allopregnanolone to developing neuronal cells reduces the GABA-A receptor α4 subunit in a concentration-dependent manner (Grobin and Morrow, 2000). However, in pregnant rats, when there are high allopregnanolone levels for a long period, no change in α4 subunit expression can be found in the cortex or hippocampus. On the seventh day postpartum (withdrawal from allopregnanolone), an increase in this subunit was detected in the hippocampus, while the thalamus was not investigated (Concas et al., 1999). With cerebellar granule cell cultures, progesterone application for five days does not change the amount of the α4 subunit, but, six hours after withdrawal, an increase was found, followed by a decrease at 12 h to 24 h after withdrawal (Follesa et al., 2000). On the other hand, an increase in GABA-A receptor α4 subunit mRNA expression has been shown in the hippocampus in relation to increased anxiety after three days of progesterone or allopregnanolone treatment and 24 h after progesterone withdrawal (Gulinello et al., 2001). With benzodiazepines, an increase in the GABA-A receptor α4 subunit in the anteroventral thalamus was obtained by treatment for four weeks (Ramsey-Williams and Carter, 1996). Interestingly, GABA-A receptor α4 subunit expression in the cortex varies depending on intermittent (increase) or continuous (decrease) supplementation of diazepam for a two-week period (Arnot et al., 2001).The α4 subunit expression therefore varies depending on the drug used, or the time of application. The GABA-A receptor β2 subunit mRNA expression in the hippocampus of the tolerant rats (group 90 min) did not differ from that in the control group. However, during the course of the development of acute tolerance, the amount of β2 subunit mRNA decreased in the 30-minute group (V-shaped changes), at a time point when the acute tolerance has not yet developed. A study with ethomidate, a selective positive GABA-A , demonstrated that the sedative and a component of the hypnotic action of this anaesthetic is mediated by β2 subunit-containing receptors (Reynolds et al., 2003). A change in β2 mRNA expression is therefore probably involved as part of the anaesthesia but not necessarily in the tolerance mechanism. For the GABA-A receptor α1, α2, α5, α6 and δ subunits, no difference in the amount of mRNA was found between groups in the brain regions that were studied.

49

In the 90-minute group, the increase in allopregnanolone doses needed to maintain anaesthesia for 65-85 min was significantly positively correlated with the α2 mRNA in different hippocampal subregions. The in vivo measurement of acute tolerance (ΔMDR) also correlated with α2 mRNA in the CA1 region (r = 0.71, b = 14.14, d.f. = 7, p<0.05). In vitro GABA-A receptors containing the α2 subunit have been shown to react less to allopregnanolone, compared with receptors including the α1, α3, or α6 subunits (Belelli et al., 2002). The higher amount of the α2 mRNA in the hippocampus of tolerant rats might therefore help to explain the need for a higher allopregnanolone dose to induce anaesthesia for 65-85 min. However, it is also possible that other mechanisms might be involved in the development of tolerance, e.g. phosphorylation which changes the steroid sensitivity of the receptor, independently of any change in subunit composition (Harney et al., 2003; Koksma et al., 2003). Our finding of the involvement of GABA-A receptor subunits in the development of acute allopregnanolone tolerance is of great importance for identifying the mechanism of steroid tolerance.

4.4 Persistence of allopregnanolone tolerance, Paper IV In this study, we investigated whether tolerance persisted after the interruption of the exposure following the induction of acute allopregnanolone tolerance by 90 minutes of allopregnanolone anaesthesia at the EEG-threshold level. All the groups were anaesthetised with allopregnanolone (3 mg/ml dissolved in 10% hydroxypropyl-β - cyclodextrin) to the anaesthesia level of the silent second (SS), but the exposure patterns were different: 1) Control group with anaesthesia to the first silent second (SS1 group), 2) 90 minutes of anaesthesia (LAn group), 3) 90 minutes of anaesthesia + anaesthesia to the silent second level (SS2) one day later (SS2;D1 group), 4) 90 minutes of anaesthesia + SS2 two days later (SS2;D2 group). In the SS2; D1 group and the SS2; D2 group, the 90-minute anaesthesia period was induced in either the morning (am subgroup) or the afternoon (pm subgroup), whereas in the LAn group the onset of anaesthesia took place between the am and pm onset (mid-time). In the SS1 group, anaesthesia was always induced around 12 pm. The data are therefore analysed for anaesthesia onset in the three subgroups, am, pm and mid-time. All the rats were decapitated between 11 am and 12.30 pm. Samples from blood, different brain regions, muscle and fat tissue were analysed for

50

allopregnanolone content. Half the brain was used for the determination of GABA-A receptor subunit mRNA expression levels. Acute tolerance to allopregnanolone developed in all groups with 90 minutes of anaesthesia (LAn, SS2;D1 and SS2;D2 groups). Acute tolerance to allopregnanolone was also established in all subgroups (am, pm and mid-time) and the subgroups individually showed a significant increase in MDR (p<0.01 - 0.001, Fig. 13A). Unexpectedly, we found a significant interaction between the time of onset of anaesthesia and MDR (p<0.05), in which ΔMDR (MDR3-MDR1) was significantly lower in the am subgroup compared with the pm subgroup (p<0.01). However, the MDRs showed no significant difference between the subgroups (Fig. 13B).

A) B)

100 am Subgroup * am Subgroup pm Subgroup * * 1.00 pm Subgroup mid-time Subgroup 75 mid-time 0.75 50

(mg/kg) 0.50

25 Cumulative dose Cumulative

rate (mg/kg/min) rate 0.25 MDR1 MDR2 MDR3 Allopregnanolone dose 0 0 10 20 30 40 50 60 70 80 90 0.00 123 Duration of anaesthesia (min) MDR10-30 MDR40-60 MDR70-90

Figure 13: Allopregnanolone cumulative dose and maintenance dose rate (MDR) during the 90-min threshold testing: A) cumulative dose in the am, pm and mid-time subgroups and B) the maintenance dose rate in the am, pm and mid-time subgroups. The maintenance dose rate in the 70- to 90-minute interval (MDR3) was different compared with the maintenance dose rate in the 40- to 60-minute interval (MDR2) and the maintenance dose rate in the 10- to 30-minute interval (MDR2). * MDR3 differences from MDR1 and MDR2 are significant (p<0.05).

The initial CNS sensitivity to allopregnanolone was determined by the amount of allopregnanolone needed to induce anaesthesia to reach the first SS (SS1). The allopregnanolone SS1 dose did not differ between groups (p>0.05; Fig. 14). To investigate whether tolerance remained one day and two days after the long anaesthesia, a new anaesthesia was induced in the SS2;D1 and SS2;D2 groups of rats. The amount of allopregnanolone needed to induce the first SS during the long anaesthesia (SS1) and the dose needed to induce SS one or two days after the long anaesthesia (SS2) showed a significant difference in the SS2;D1 group (ΔSS, p<0.05, Fig. 14)., Between the SS2;D1 and SS2;D2 groups, the ΔSS value did not differ (p<0.05). These findings suggest that tolerance only persisted for one day. The time of

51

day of the onset of anaesthesia (am, pm and mid-time) had no effect on the persistence of tolerance.

* SS1 LAn 10 SS2;D1 SS2;D2

Dose of Dose (mg/kg) 5 Allopregnanolone

0 SS1 SS2 ΔSS (SS2-SS1)

Figure 14: Persistence of acute tolerance. The dose of allopregnanolone needed for the induction of the first SS (SS1) during the 90-min anaesthesia and for the induction of SS2 in the same individual rats in the SS2;D1 and SS2;D2 groups. The SS2;D1 dose was significantly higher than the SS1 dose in the same animals. The ΔSS (SS2-SS1) value in the SS2;D1 group did not differ significantly from the ΔSS in the SS2;D2 group. * The difference is significant (p<0.05).

The allopregnanolone concentrations in the LAn group were significantly higher in the MTH, brain stem and fat tissue compared with the other groups (p<0.01, 0.01 and 0.001 respectively; Fig. 15). The time of day of the onset of anaesthesia had no influence on the allopregnanolone concentrations in the brain regions. * 80 SS1 LAn * SS2;D1 60 SS2;D2 * 40 nmol/g -nmol/L

[Allopregnanolone] 20

0 x H at T um F M Corte in stem Muscle Serum Striatum a Br Hippocamp. Cerebell Figure 15: Allopregnanolone concentrations in tissues and serum. The LAn group was decapitated after a 90-minute long anaesthesia, while the SS1, SS2;D1 and SS2;D2 groups were decapitated after the induction of anaesthesia up to the SS criterion. The values are presented as the mean ± SEM. * Between groups, there is a significant overall difference, p<0.05.

52

It is suggested that allopregnanolone produces a regional difference in responsiveness, in agreement with the variability in the subunit composition of GABA-A receptors in different brain regions (Jussofie, 1993a; 1993b). The normal level of allopregnanolone in the rat brain cortex is 1.3- 8.6 nmol/kg (Purdy et al., 1991) (Johansson et al., 2002) and the amount of drug producing the loss of the righting reflex is with some drugs sufficient for tolerance development (Dingemanse et al., 1990): With allopregnanolone, this is achieved with an concentration of 3 μmol/kg (Korneyev and Costa, 1996). The SS occurs at a deeper anaesthesia than the loss of righting reflex (Korkmaz and Wahlstrom, 1997) and, in the present study, the allopregnanolone concentration in the cortex of all the groups was at least nine times higher than the concentration required to induce the loss of the righting reflex (> 27 µmol/kg). The allopregnanolone concentration in the SS2;D1 group was not higher than the concentration in the control group (SS1), except for in fat. It is possible that the methods that were used were not sensitive enough to detect localised changes in allopregnanolone concentrations in the grossly dissected brain regions. A detailed study of the rat brain uptake of progesterone and its metabolites pregnanolone and pregnanedione revealed that the higher steroid concentration in the brain was not associated with a stronger anaesthesia effect (Raisinghani et al., 1968). Taken together with our results, it can be concluded that the allopregnanolone concentration in tissue may change after acute tolerance development. There was no significant difference between groups in the expression of different GABA-A receptor subunits in the hippocampus (α2) and VPM (α2, α4, δ, γ2). However, the α4 subunit mRNA in the VPM of the individual animals correlated negatively both with the dose of allopregnanolone needed to induce SS2 in the SS2;D1 group (r=-0.88, b=-0.33, DF=8, p<0.01) and with ΔSS (SS2-SS1, r=-0.70, b=-0.27, DF=8, p<0.05). There were no similar correlations in the non-tolerant SS2;D2 group. These recorded negative correlations are very similar to the correlation already reported in Paper III. In the present study, the change in α4 subunit mRNA expression in the thalamic VPM regions (6% changed) was not significant between groups. The α4 subunit containing receptors in normal rats account for 20% of the total thalamic GABA-A receptors (Sur et al., 1999) and it therefore appears that a small variation in α4 level could have a significant functional role. In Paper III, a 10% change in the amount of the α4 subunit (both mRNA and protein) was, however, found to be significant. Animals anaesthetised in the afternoon (pm subgroup) required a higher amount of allopregnanolone for the maintenance of anaesthesia than the

53

animals anaesthetised in the morning. Moreover, in the pm subgroup, the allopregnanolone penetration to the brain (Q-values) was lower than in the am and mid-time subgroups. One possible explanation is based on the diurnal changes in the glucocorticoid level that may influence tolerance development. It has been shown that high levels of glucocorticoid in the brain change the GABA level (Sherman and Gebhart, 1974; Singh et al., 1979) and can influence the allopregnanolone effect and finally augment the function of the brain GABA system (Schwartz et al., 1987; Stromberg et al., 2005). The circadian rhythm of the in rats shows the highest plasma level at the beginning of the dark period, followed by a decrease to the lowest level in the middle of this period (Atkinson and Waddell, 1997). In the current study, the am rats were anaesthetised at the time of the highest glucocorticoid level, as we used a reversed light-dark period, and the pm rats were anaesthetised at the time of the lowest glucocorticoid secretion. To our knowledge, no detailed study of the time course of the acquisition and subsequent loss of allopregnanolone tolerance has been conducted. The present work shows for the first time that acute allopregnanolone tolerance develops progressively over a period of minutes (< 90 min) and reverts to normal several hours (< 48 h) after the end of the neurosteroid treatment.

54

5 GENERAL DISCUSSION

The present thesis shows that the negative effects induced by allopregnanolone on spatial learning can be reduced by using a 3β- hydroxypregnane steroid (UC1011). Numerous studies have evidenced the existence of alterations in learning and memory processes during the course of female life. Cognitive disturbances are reported in pregnancy and during the luteal phase of the menstrual cycle, when high circulating levels of progesterone and allopregnanolone are found (Brett and Baxendale, 2001; Man et al., 1999). In animals, treatment with the neurosteroid allopregnanolone impairs learning and memory (Frye and Sturgis, 1995; Johansson et al., 2002). It would therefore be of great importance to discover or synthesise a drug that is effective against allopregnanolone-induced cognitive disturbances. UC1011 appears to be a promising substance in this field, as it can reduce the allopregnanolone-induced learning impairment and GABA potentiation.

The sensitivity to a substance can decrease, i.e. tolerance can be induced during a single administration of the substance (acute tolerance) or after multiple exposures to the substance (chronic tolerance) (Kalant et al., 1971). Many of the behavioural and physiological effects of allopregnanolone are similar to those of ethanol and other positive modulators of the GABA-A receptor, such as benzodiazepines and barbiturates (Majewska et al., 1986; Paul and Purdy, 1992; Vanover et al., 1999). In contrast to benzodiazepines and ethanol, less is known about the tolerance that might result from exposure to allopregnanolone. In women, the levels of allopregnanolone or allopregnanolone-like substances are high during the luteal phase of the menstrual cycle, pregnancy, use of oral contraceptives and hormone replacement therapy. Tolerance may therefore develop and may cause the appearance of mood and behavioural symptoms during these periods (see Section 1.3). Our results show that one prolonged exposure to allopregnanolone can cause acute tolerance development in rats following less than 90 minutes’ exposure. Moreover, tolerance persists to one day after the induction of acute allopregnanolone tolerance. The time of day of the onset of anaesthesia may also play an important role in the development of allopregnanolone tolerance by influencing the neurosteroid impact in the

55

CNS. We also succeeded in developing chronic tolerance to allopregnanolone after seven days of repeated exposure. The activity of steroids has been suggested to depend on the subunit composition of the GABA-A receptor. Lambert and co-workers have clearly demonstrated a role for the subunit composition of the receptor in neuroactive steroid/GABA-A receptor interactions (Lambert et al., 2003). However, it is not clear which GABA-A receptor subunits are involved in the development of allopregnanolone tolerance. Accordingly, we analysed the expression of GABA-A receptor subunits in the brain and its correlation with tolerance development after 90 minutes of allopregnanolone anaesthesia. In the tolerant group, both the mRNA expression and protein level of the GABA-A receptor α4 subunit in the ventral posteriomedial nucleus of the thalamus was lower than in non-tolerant rats and the in vivo measurements of tolerance (ΔMDR and ΔSS) showed negative correlations with the actual amount of α4 mRNA in the ventral posteriomedial nucleus of the thalamus. As receptors containing the α4 subunit display higher GABA sensitivity than most other receptors containing α subunit (Knoflach et al., 1996), a decrease in α4 subunit expression level can then be suggested as one of the pathways changing the effect of allopregnanolone during tolerance development and the maintenance of tolerance. Finally, the strength of the present thesis is that it addresses the questions of allopregnanolone tolerance and changes in GABA-A receptor function, as this may be related to the cognitive impairment and emotional changes in certain periods of female life.

56

6. CONCLUSIONS

1. The increase in chloride ion uptake induced by allopregnanolone can be inhibited by UC1011.

2. UC1011 significantly improves the allopregnanolone-induced impairment of learning in the water maze test, probably via an effect on the GABA-A receptor.

3. The negative effects of allopregnanolone in the water maze can be improved after pre-treatment with allopregnanolone for several days.

4. Repeated exposure to allopregnanolone induces the development of tolerance to the effects of allopregnanolone.

5. Acute allopregnanolone tolerance can be induced and the GABA-A receptor α4 subunit in the thalamus might be involved in the development of tolerance.

6. Tolerance persists to one day but not to two days after the end of a long allopregnanolone anaesthesia inducing acute tolerance.

7. The time of day of the onset of anaesthesia with allopregnanolone is of importance for the development of tolerance.

57

7. ACKNOWLEDGEMENTS

I wish to express my deepest gratitude to all the people who have supported me throughout this work and especially to:

Associate Professor Inga-Maj Johansson, my main supervisor, for her excellent academic supervision and for introducing me to the field of molecular biology, tack för din eminenta guidning genom vetenskapens gåta

Professor Torbjörn Bäckström, my co-supervisor, for accepting me as a student and introducing and inspiring me to work in the exciting world of research. The opportunity to work with you was something that really helped me to acquire a view of the way science works and opened my mind and encouraged me to work with enthusiasm

Professor Göran Wahlström, for fruitful collaboration and sharing great research experiences and ideas

Professor Ann Lalos, head of the Department of Obstetrics and , for friendly support

Dr. Per Lundgren, for collaboration and friendly guidance in all areas Monica Isaksson, for exceptional and excellent technical laboratory assistance Elizabeth Zingmark and Agneta Andersson, for all the help in the laboratory with the steroid analysis Tobias Näslund, Birgit Ericsson and Carina Jonsson, for all the administrative help during my research studies at the Department of Obstetrics and Gynaecology My friends and co-workers, Charlotte Lindblad, Magnus Löfgren, Vita Birzniece, Mozibur Rahman, Jessica Strömberg, Magdalena Taube, Gianna Ragagnin, Di Zhu, Ming-de Wang, and all wonderful people at the Department of Obstetrics and Gynaecology To all the people at UKBF 3B, especially Jamshid Pourarzi, for always helping out and sharing their experience

58

To my sister and brother in low, Shahla and Hooshmand, for unfailing help and for trying to make my life easy in Umeå

Finally, to my family, Melek, Atlan and Aral, this thesis would have not been possible to finish without their love and understanding, and to my parents, Kerim and Arazbakht, for their unreserved and endless support throughout the whole of my life.

This study was carried out at the Department of Clinical Science, Obstetrics and Gynaecology, Umeå University, and was supported by the EU Structural Programme, objective 1, the Swedish Research Council, Medicine (project No. 72X-11198), a “Spjutspets” grant from Norrland University Hospital, and by Umeå University Foundations.

59

8. REFERENCES AC, N.W., Sundstrom-Poromaa, I., Backstrom, T. 2005 Action by and sensitivity to neuroactive steroids in menstrual cycle related CNS disorders. Psychopharmacology (Berl), 1-14. Adashi, E.Y. 1992 The ovarian life cycle. In Yen, S.C.C., Jaffe, R.B. Reproductive ; Physiology, pathophysiology and clinical management. Chapter 6. W. B. Saunders Company., Philadelphia, USA.181-237 Allison, C., Pratt, J.A. 2003 Neuroadaptive processes in GABAergic and glutamatergic systems in benzodiazepine dependence. Pharmacol Ther, 98, 171-195. AmericanPsychiatricAssociation. 1994 Diagnostic and Statistical Manual of Mental Disorders. 4th edition. American Psychiatric Press, Washington, DC Andreen, L., Bixo, M., Nyberg, S., Sundstrom-Poromaa, I., Backstrom, T. 2003 Progesterone effects during sequential hormone replacement therapy. Eur J Endocrinol, 148, 571-577. Andreen, L., Spigset, O., Andersson, A., Nyberg, S., Backstrom, T. 2006 Pharmacokinetics of progesterone and its metabolites allopregnanolone and pregnanolone after oral administration of low-dose progesterone. Maturitas. Andreen, L., Sundstrom-Poromaa, I., Bixo, M., Andersson, A., Nyberg, S., Backstrom, T. 2005 Relationship between allopregnanolone and negative mood in postmenopausal women taking sequential hormone replacement therapy with vaginal progesterone. Psychoneuroendocrinology, 30, 212-224. Arafat, E.S., Hargrove, J.T., Maxson, W.S., Desiderio, D.M., Wentz, A.C., Andersen, R.N. 1988 Sedative and hypnotic effects of oral administration of micronized progesterone may be mediated through its metabolites. Am J Obstet Gynecol, 159, 1203-1209. Arnot, M.I., Davies, M., Martin, I.L., Bateson, A.N. 2001 GABA(A) receptor gene expression in rat cortex: differential effects of two chronic diazepam treatment regimes. J Neurosci Res, 64, 617-625. Aston, R. 1965 Quantitative aspects of tolerance and posttolerance hypersensitivity to pentobarbital in the rat. J Pharmacol Exp Ther, 150, 253-258. Aston, R. 1966 Acute tolerance indices for pentobarbital in male and female rats. J Pharmacol Exp Ther, 152, 350-353. Astur, R.S., Ortiz, M.L., Sutherland, R.J. 1998 A characterization of performance by men and women in a virtual Morris water task: a large and reliable sex difference. Behav Brain Res, 93, 185-190. Atkinson, H.C., Waddell, B.J. 1997 Circadian variation in basal plasma corticosterone and adrenocorticotropin in the rat: and changes across the . Endocrinology, 138, 3842-3848. Backstrom, T. 1976 Epileptic seizures in women related to plasma estrogen and progesterone during the menstrual cycle. Acta Neurol Scand, 54, 321-347. Backstrom, T. 1996 Side-efects of contraceptives and gonadal hormones. Bailliere´s Clinical Psychiatry. Bailliere Tindall. pp 713-724 Backstrom, T., Andersson, A., Andree, L., Birzniece, V., Bixo, M., Bjorn, I., Haage, D., Isaksson, M., Johansson, I.M., Lindblad, C., Lundgren, P., Nyberg, S., Odmark, I.S., Stromberg, J., Sundstrom-Poromaa, I., Turkmen, S., Wahlstrom, G., Wang,

60

M., Wihlback, A.C., Zhu, D., Zingmark, E. 2003a Pathogenesis in menstrual cycle-linked CNS disorders. Ann N Y Acad Sci, 1007, 42-53. Backstrom, T., Andreen, L., Birzniece, V., Bjorn, I., Johansson, I.M., Nordenstam- Haghjo, M., Nyberg, S., Sundstrom-Poromaa, I., Wahlstrom, G., Wang, M., Zhu, D. 2003b The role of hormones and hormonal treatments in premenstrual syndrome. CNS Drugs, 17, 325-342. Backstrom, T., Sanders, D., Leask, R., Davidson, D., Warner, P., Bancroft, J. 1983 Mood, sexuality, hormones, and the menstrual cycle. II. Hormone levels and their relationship to the premenstrual syndrome. Psychosom Med, 45, 503-507. Backstrom, T., Wahlstrom, G., Wahlstrom, K., Zhu, D., Wang, M.D. 2005 Isoallopregnanolone; an antagonist to the anaesthetic effect of allopregnanolone in male rats. Eur J Pharmacol, 512, 15-21. Backstrom, T., Zetterlund, B., Blom, S., Romano, M. 1984 Effects of intravenous progesterone infusions on the epileptic discharge frequency in women with partial epilepsy. Acta Neurol Scand, 69, 240-248. Backus, K.H., Arigoni, M., Drescher, U., Scheurer, L., Malherbe, P., Mohler, H., Benson, J.A. 1993 Stoichiometry of a recombinant GABAA receptor deduced from mutation-induced rectification. Neuroreport, 5, 285-288. Barbaccia, M.L., Roscetti, G., Trabucchi, M., Purdy, R.H., Mostallino, M.C., Concas, A., Biggio, G. 1997 The effects of inhibitors of GABAergic transmission and stress on brain and plasma allopregnanolone concentrations. Br J Pharmacol, 120, 1582-1588. Barker, M.J., Greenwood, K.M., Jackson, M., Crowe, S.F. 2004 Persistence of cognitive effects after withdrawal from long-term benzodiazepine use: a meta-analysis. Arch Clin Neuropsychol, 19, 437-454. Barnard, E.A., Skolnick, P., Olsen, R.W., Mohler, H., Sieghart, W., Biggio, G., Braestrup, C., Bateson, A.N., Langer, S.Z. 1998 International Union of . XV. Subtypes of gamma-aminobutyric acidA receptors: classification on the basis of subunit structure and receptor function. Pharmacol Rev, 50, 291-313. Barnes, E.M., Jr. 1996 Use-dependent regulation of GABAA receptors. Int Rev Neurobiol, 39, 53-76. Barnes, E.M., Jr. 2001 Assembly and intracellular trafficking of GABAA receptors. Int Rev Neurobiol, 48, 1-29. Baulieu, E.E., Robel, P. 1990 Neurosteroids: a new brain function? J Steroid Biochem Mol Biol, 37, 395-403. Baulieu, E.E., Robel, P. 1995 Non-genomic mechanisms of action of steroid hormones. Ciba Found Symp, 191, 24-37; discussion 37-42. Beesley, J.E. 1993 Immunocytochemistry. Oxford University Press Inc., New York.1-3 Belelli, D., Casula, A., Ling, A., Lambert, J.J. 2002 The influence of subunit composition on the interaction of neurosteroids with GABA(A) receptors. Neuropharmacology, 43, 651-661. Belelli, D., Lambert, J.J. 2005 Neurosteroids: endogenous regulators of the GABA(A) receptor. Nat Rev Neurosci, 6, 565-575. Belleville, R.E., Fraser, H.F. 1957 Tolerance to some effects of barbiturates. J Pharmacol Exp Ther, 120, 469-474.

61

Benke, D., Fritschy, J.M., Trzeciak, A., Bannwarth, W., Mohler, H. 1994 Distribution, prevalence, and drug binding profile of gamma-aminobutyric acid type A receptor subtypes differing in the beta-subunit variant. J Biol Chem, 269, 27100-27107. Bicikova, M., Dibbelt, L., Hill, M., Hampl, R., Starka, L. 1998 Allopregnanolone in women with premenstrual syndrome. Horm Metab Res, 30, 227-230. Bicikova, M., Klak, J., Hill, M., Zizka, Z., Hampl, R., Calda, P. 2002 Two neuroactive steroids in midpregnancy as measured in maternal and fetal sera and in amniotic fluid. Steroids, 67, 399-402. Biggio, G., Dazzi, L., Biggio, F., Mancuso, L., Talani, G., Busonero, F., Mostallino, M.C., Sanna, E., Follesa, P. 2003 Molecular mechanisms of tolerance to and withdrawal of GABA(A) receptor modulators. Eur Neuropsychopharmacol, 13, 411-423. Bitran, D., Hilvers, R.J., Kellogg, C.K. 1991 Anxiolytic effects of 3 alpha-hydroxy-5 alpha[beta]-pregnan-20-one: endogenous metabolites of progesterone that are active at the GABAA receptor. Brain Res, 561, 157-161. Bitran, D., Purdy, R.H., Kellogg, C.K. 1993 Anxiolytic effect of progesterone is associated with increases in cortical allopregnanolone and GABAA receptor function. Pharmacol Biochem Behav, 45, 423-428. Bixo, M., Andersson, A., Winblad, B., Purdy, R.H., Backstrom, T. 1997 Progesterone, 5alpha-pregnane-3,20-dione and 3alpha-hydroxy-5alpha-pregnane-20-one in specific regions of the human female brain in different endocrine states. Brain Res, 764, 173-178. Bixo, M., Backstrom, T., Winblad, B. 1984 Progesterone distribution in the brain of the PMSG treated female rat. Acta Physiol Scand, 122, 355-359. Bixo, M., Backstrom, T., Winblad, B., Selstam, G., Andersson, A. 1986 Comparison between pre- and postovulatory distributions of oestradiol and progesterone in the brain of the PMSG-treated rat. Acta Physiol Scand, 128, 241-246. Bjorn, I., Bixo, M., Nojd, K.S., Collberg, P., Nyberg, S., Sundstrom-Poromaa, I., Backstrom, T. 2002 The impact of different doses of medroxyprogesterone acetate on mood symptoms in sequential hormonal therapy. Gynecol Endocrinol, 16, 1-8. Bjorn, I., Bixo, M., Nojd, K.S., Nyberg, S., Backstrom, T. 2000 Negative mood changes during hormone replacement therapy: a comparison between two . Am J Obstet Gynecol, 183, 1419-1426. Bjorn, I., Sundstrom-Poromaa, I., Bixo, M., Nyberg, S., Backstrom, G., Backstrom, T. 2003 Increase of estrogen dose deteriorates mood during progestin phase in sequential hormonal therapy. J Clin Endocrinol Metab, 88, 2026-2030. Bloch, M., Rotenberg, N., Koren, D., Klein, E. 2005 Risk factors associated with the development of postpartum mood disorders. J Affect Disord, 88, 9-18. Bolander, H.G., Wahlstrom, G. 1988 Acute tolerance to and distribution of hexobarbital in relation to depth and duration of anaesthesia in rats. Pharmacol Toxicol, 63, 199-204. Bouchard, P. 1999 Progesterone and the progesterone receptor. J Reprod Med, 44, 153- 157. Brett, M., Baxendale, S. 2001 Motherhood and memory: a review. Psychoneuroendocrinology, 26, 339-362.

62

Brussaard, A.B., Koksma, J.J. 2002 Short-term modulation of GABAA receptor function in the adult female rat. Prog Brain Res, 139, 31-42. Brussaard, A.B., Wossink, J., Lodder, J.C., Kits, K.S. 2000 Progesterone-metabolite prevents protein kinase C-dependent modulation of gamma-aminobutyric acid type A receptors in oxytocin neurons. Proc Natl Acad Sci U S A, 97, 3625-3630. Cain, D.P., Saucier, D. 1996 The neuroscience of spatial navigation: focus on behavior yields advances. Rev Neurosci, 7, 215-231. Carl, P., Hogskilde, S., Nielsen, J.W., Sorensen, M.B., Lindholm, M., Karlen, B., Backstrom, T. 1990 Pregnanolone emulsion. A preliminary pharmacokinetic and pharmacodynamic study of a new intravenous anaesthetic agent. Anaesthesia, 45, 189-197. Casey, M.L. 1993 The Human Plasenta. In: Redman, C.W.G. Blackwell Scientific Publications., Oxford. PP 237-272 Chang, Y., Wang, R., Barot, S., Weiss, D.S. 1996 Stoichiometry of a recombinant GABAA receptor. J Neurosci, 16, 5415-5424. Cheney, D.L., Uzunov, D., Costa, E., Guidotti, A. 1995 Gas chromatographic-mass fragmentographic quantitation of 3 alpha-hydroxy-5 alpha-pregnan-20-one (allopregnanolone) and its precursors in blood and brain of adrenalectomized and castrated rats. J Neurosci, 15, 4641-4650. Chien, Y.W. 1981 Long-acting parenteral drug formulations. J Parenter Sci Technol, 35, 106-139. Christ, M., Douwes, K., Eisen, C., Bechtner, G., Theisen, K., Wehling, M. 1995 Rapid effects of on sodium transport in vascular smooth muscle cells. Hypertension, 25, 117-123. Coiret, G., Matifat, F., Hague, F., Ouadid-Ahidouch, H. 2005 17-beta-estradiol activates maxi-K channels through a non-genomic pathway in human cancer cells. FEBS Lett, 579, 2995-3000. Collinson, N., Kuenzi, F.M., Jarolimek, W., Maubach, K.A., Cothliff, R., Sur, C., Smith, A., Otu, F.M., Howell, O., Atack, J.R., McKernan, R.M., Seabrook, G.R., Dawson, G.R., Whiting, P.J., Rosahl, T.W. 2002 Enhanced learning and memory and altered GABAergic synaptic transmission in mice lacking the alpha 5 subunit of the GABAA receptor. J Neurosci, 22, 5572-5580. Compagnone, N.A., Mellon, S.H. 2000 Neurosteroids: biosynthesis and function of these novel neuromodulators. Front Neuroendocrinol, 21, 1-56. Concas, A., Follesa, P., Barbaccia, M.L., Purdy, R.H., Biggio, G. 1999 Physiological modulation of GABA(A) receptor plasticity by progesterone metabolites. Eur J Pharmacol, 375, 225-235. Corpechot, C., Robel, P., Axelson, M., Sjovall, J., Baulieu, E.E. 1981 Characterization and measurement of dehydroepiandrosterone in rat brain. Proc Natl Acad Sci U S A, 78, 4704-4707. Corpechot, C., Young, J., Calvel, M., Wehrey, C., Veltz, J.N., Touyer, G., Mouren, M., Prasad, V.V., Banner, C., Sjovall, J., et al. 1993 Neurosteroids: 3 alpha-hydroxy-5 alpha-pregnan-20-one and its precursors in the brain, plasma, and steroidogenic glands of male and female rats. Endocrinology, 133, 1003-1009.

63

Cullberg, J. 1972 Mood changes and menstrual symptoms with different gestagen/estrogen combinations. A double blind comparison with a placebo. Acta Psychiatr Scand Suppl, 236, 1-86. Czlonkowska, A.I., Krzascik, P., Sienkiewicz-Jarosz, H., Siemiatkowski, M., Szyndler, J., Maciejak, P., Bidzinski, A., Plaznik, A. 2001 Tolerance to the anticonvulsant activity of and allopregnanolone in a model of picrotoxin seizures. Eur J Pharmacol, 425, 121-127. Damianisch, K., Rupprecht, R., Lancel, M. 2001 The influence of subchronic administration of the neurosteroid allopregnanolone on sleep in the rat. Neuropsychopharmacology, 25, 576-584. Darnaudery, M., Koehl, M., Piazza, P.V., Le Moal, M., Mayo, W. 2000 increases hippocampal acetylcholine release and spatial recognition. Brain Res, 852, 173-179. Davey, D.A. 1995 Dewhurst´s Textbook of Obstetrics and Gynecology For Postgraduates. In:Whitfield, C.R. (Ed.). Blackwell Scientific Publications, Oxford. PP 87-108. Dazzi, L., Sanna, A., Cagetti, E., Concas, A., Biggio, G. 1996 Inhibition by the neurosteroid allopregnanolone of basal and stress-induced acetylcholine release in the brain of freely moving rats. Brain Res, 710, 275-280. de Wit, H., Schmitt, L., Purdy, R., Hauger, R. 2001 Effects of acute progesterone administration in healthy postmenopausal women and normally-cycling women. Psychoneuroendocrinology, 26, 697-710. DeLorey, T.M., Olsen, R.W. 1992 Gamma-aminobutyric acidA receptor structure and function. J Biol Chem, 267, 16747-16750. Devan, B.D., McDonald, R.J., White, N.M. 1999 Effects of medial and lateral caudate- putamen lesions on place- and cue-guided behaviors in the water maze: relation to thigmotaxis. Behav Brain Res, 100, 5-14. Devaud, L.L., Fritschy, J.M., Sieghart, W., Morrow, A.L. 1997 Bidirectional alterations of GABA(A) receptor subunit peptide levels in rat cortex during chronic ethanol consumption and withdrawal. J Neurochem, 69, 126-130. Dingemanse, J., Hameter, B.M., Danhof, M. 1990 of tolerance development to the and anticonvulsant effects of in rats. J Pharm Sci, 79, 207-211. Dripps, R.D., Dundee, J.W., Price, H.L. 1956 Acute tolerance to thiopentone in man. Br J Anaesth, 28, 344-352. Duras, M., Mlynarczuk, J., Kotwica, J. 2005 Non-genomic effect of steroids on oxytocin- stimulated intracellular mobilization of and on F2alpha and E2 secretion from bovine endometrial cells. Other Mediat, 76, 105-116. Essrich, C., Lorez, M., Benson, J.A., Fritschy, J.M., Luscher, B. 1998 Postsynaptic clustering of major GABAA receptor subtypes requires the gamma 2 subunit and gephyrin. Nat Neurosci, 1, 563-571. Fancsik, A., Linn, D.M., Tasker, J.G. 2000 Neurosteroid modulation of GABA IPSCs is phosphorylation dependent. J Neurosci, 20, 3067-3075. Follesa, P., Concas, A., Porcu, P., Sanna, E., Serra, M., Mostallino, M.C., Purdy, R.H., Biggio, G. 2001 Role of allopregnanolone in regulation of GABA(A) receptor

64

plasticity during long-term exposure to and withdrawal from progesterone. Brain Res Brain Res Rev, 37, 81-90. Follesa, P., Serra, M., Cagetti, E., Pisu, M.G., Porta, S., Floris, S., Massa, F., Sanna, E., Biggio, G. 2000 Allopregnanolone synthesis in cerebellar granule cells: roles in regulation of GABA(A) receptor expression and function during progesterone treatment and withdrawal. Mol Pharmacol, 57, 1262-1270. Freeman, E.W., Purdy, R.H., Coutifaris, C., Rickels, K., Paul, S.M. 1993 Anxiolytic metabolites of progesterone: correlation with mood and performance measures following oral progesterone administration to healthy female volunteers. Neuroendocrinology, 58, 478-484. Fritschy, J.M., Weinmann, O., Wenzel, A., Benke, D. 1998 Synapse-specific localization of NMDA and GABA(A) receptor subunits revealed by antigen-retrieval immunohistochemistry. J Comp Neurol, 390, 194-210. Frye, C.A., Bayon, L.E. 1999 Cyclic withdrawal from endogenous and exogenous progesterone increases kainic acid and perforant pathway induced seizures. Pharmacol Biochem Behav, 62, 315-321. Frye, C.A., Duncan, J.E. 1994 Progesterone metabolites, effective at the GABAA receptor complex, attenuate pain sensitivity in rats. Brain Res, 643, 194-203. Frye, C.A., Mermelstein, P.G., DeBold, J.F. 1992 Evidence for a non-genomic action of progestins on sexual receptivity in hamster ventral tegmental area but not hypothalamus. Brain Res, 578, 87-93. Frye, C.A., Sturgis, J.D. 1995 Neurosteroids affect spatial/reference, working, and long- term memory of female rats. Neurobiol Learn Mem, 64, 83-96. Frye, C.A., Van Keuren, K.R., Erskine, M.S. 1996 Behavioral effects of 3 alpha- androstanediol. I: Modulation of sexual receptivity and promotion of GABA- stimulated chloride flux. Behav Brain Res, 79, 109-118. Gallagher, M., Burwell, R., Burchinal, M. 1993 Severity of spatial learning impairment in aging: development of a learning index for performance in the Morris water maze. Behav Neurosci, 107, 618-626. Gallo, M.A., Smith, S.S. 1993 Progesterone withdrawal decreases latency to and increases duration of electrified prod burial: a possible rat model of PMS anxiety. Pharmacol Biochem Behav, 46, 897-904. Gangrade, N.K., Boudinot, F.D., Price, J.C. 1992 Pharmacokinetics of progesterone in ovariectomized rats after single dose intravenous administration. Biopharm Drug Dispos, 13, 703-709. Gasior, M., Carter, R.B., Witkin, J.M. 1999 Neuroactive steroids: potential therapeutic use in neurological and psychiatric disorders. Trends Pharmacol Sci, 20, 107-112. Gazzaniga, M.S., Ivry, R.B., Mangun, G.R. 2002 Learning and Memory. In Cognitive Neuroscience.Chapter 8. pp 302-350 Gee, K.W. 1988 Steroid modulation of the GABA/benzodiazepine receptor-linked chloride ionophore. Mol Neurobiol, 2, 291-317. Genazzani, A.R., Petraglia, F., Bernardi, F., Casarosa, E., Salvestroni, C., Tonetti, A., Nappi, R.E., Luisi, S., Palumbo, M., Purdy, R.H., Luisi, M. 1998 Circulating levels of allopregnanolone in humans: , age, and endocrine influences. J Clin Endocrinol Metab, 83, 2099-2103.

65

Genazzani, A.R., Stomati, M., Morittu, A., Bernardi, F., Monteleone, P., Casarosa, E., Gallo, R., Salvestroni, C., Luisi, M. 2000 Progesterone, progestagens and the central nervous system. Hum Reprod, 15 Suppl 1, 14-27. Gerlai, R. 2001 Behavioral tests of hippocampal function: simple paradigms complex problems. Behav Brain Res, 125, 269-277. Ghoneim, M.M., Mewaldt, S.P., Berie, J.L., Hinrichs, J.V. 1981 Memory and performance effects of single and 3-week administration of diazepam. Psychopharmacology (Berl), 73, 147-151. Gluckman, M.I., Gruber, C.M. 1952 Development of tolerance and cross tolerance by mephobarbital. Proc Soc Exp Biol Med, 79, 87-88. Gorrie, G.H., Vallis, Y., Stephenson, A., Whitfield, J., Browning, B., Smart, T.G., Moss, S.J. 1997 Assembly of GABAA receptors composed of alpha1 and beta2 subunits in both cultured neurons and fibroblasts. J Neurosci, 17, 6587-6596. Grady, K.L., Phoenix, C.H., Young, W.C. 1965 Role Of The Developing Rat Testis In Differentiation Of The Neural Tissues Mediating Mating Behavior. J Comp Physiol Psychol, 59, 176-182. Greenblatt, D.J., Shader, R.I. 1978 Dependence, tolerance, and addiction to benzodiazepines: clinical and pharmacokinetic considerations. Drug Metab Rev, 8, 13-28. Grobin, A.C., Matthews, D.B., Devaud, L.L., Morrow, A.L. 1998 The role of GABA(A) receptors in the acute and chronic effects of ethanol. Psychopharmacology (Berl), 139, 2-19. Grobin, A.C., Morrow, A.L. 2000 3alpha-hydroxy-5alpha-pregnan-20-one exposure reduces GABA(A) receptor alpha4 subunit mRNA levels. Eur J Pharmacol, 409, R1-2. Gulinello, M., Gong, Q.H., Li, X., Smith, S.S. 2001 Short-term exposure to a neuroactive steroid increases alpha4 GABA(A) receptor subunit levels in association with increased anxiety in the female rat. Brain Res, 910, 55-66. Haage, D., Backstrom, T., Johansson, S. 2005 Interaction between allopregnanolone and pregnenolone sulfate in modulating GABA-mediated synaptic currents in neurons from the rat medial preoptic nucleus. Brain Res, 1033, 58-67. Halbreich, U. 1990 Gonadal hormones and antihormones, serotonin and mood. Psychopharmacol Bull, 26, 291-295. Hamilton, D.A., Driscoll, I., Sutherland, R.J. 2002 Human place learning in a virtual Morris water task: some important constraints on the flexibility of place navigation. Behav Brain Res, 129, 159-170. Hammarback, S., Backstrom, T., Holst, J., von Schoultz, B., Lyrenas, S. 1985 Cyclical mood changes as in the premenstrual tension syndrome during sequential estrogen-progestagen postmenopausal replacement therapy. Acta Obstet Gynecol Scand, 64, 393-397. Hammarback, S., Ekholm, U.B., Backstrom, T. 1991 Spontaneous anovulation causing disappearance of cyclical symptoms in women with the premenstrual syndrome. Acta Endocrinol (Copenh), 125, 132-137. Harney, S.C., Frenguelli, B.G., Lambert, J.J. 2003 Phosphorylation influences neurosteroid modulation of synaptic GABAA receptors in rat CA1 and dentate gyrus neurones. Neuropharmacology, 45, 873-883.

66

Harris, R.A., Allan, A.M. 1985 Functional coupling of gamma-aminobutyric acid receptors to chloride channels in brain membranes. Science, 228, 1108-1110. Harris, R.A., Brodie, M.S., Dunwiddie, T.V. 1992 Possible substrates of ethanol reinforcement: GABA and dopamine. Ann N Y Acad Sci, 654, 61-69. Harrison, N.L., Majewska, M.D., Harrington, J.W., Barker, J.L. 1987 Structure-activity relationships for steroid interaction with the gamma-aminobutyric acidA receptor complex. J Pharmacol Exp Ther, 241, 346-353. Hering, H., Sheng, M. 2001 Dendritic spines: structure, dynamics and regulation. Nat Rev Neurosci, 2, 880-888. Holt, R.A., Bateson, A.N., Martin, I.L. 1996 Chronic treatment with diazepam or differently affects the expression of GABAA receptor subunit mRNAs in the rat cortex. Neuropharmacology, 35, 1457-1463. Holzbauer, M. 1976 Physiological aspects of steroids with anaesthetic properties. Med Biol, 54, 227-242. Holzbauer, M., Birmingham, M.K., De Nicola, A.F., Oliver, J.T. 1985 In vivo secretion of 3 alpha-hydroxy-5 alpha-pregnan-20-one, a potent anaesthetic steroid, by the of the rat. J Steroid Biochem, 22, 97-102. Hosie, A., Wilkins, ME., da Silva, H., Smart, TG. 2005 Sites action of neurosteroids on GABA-A receptors. Abstract Of Invited Lectures And Free Contributions. International Meeting Steroid And Nervous System. Torino, Italy, Villa Gualino, Feb-2005, Madrid Hughes, L.M., Wasserman, E.A., Hinrichs, J.V. 1984 Chronic diazepam administration and appetitive discrimination learning: acquisition versus steady-state performance in pigeons. Psychopharmacology (Berl), 84, 318-322. Huh, K.H., Endo, S., Olsen, R.W. 1996 Diazepam-insensitive GABAA receptors in rat cerebellum and thalamus. Eur J Pharmacol, 310, 225-233. Ihmsen, H., Albrecht, S., Hering, W., Schuttler, J., Schwilden, H. 2004 Modelling acute tolerance to the EEG effect of two benzodiazepines. Br J Clin Pharmacol, 57, 153-161. Jaffe, J.H., Sharpless, S.K. 1965 The rapid development of physical dependence on barbiturates. J Pharmacol Exp Ther, 150, 140-145. Jarvinen, A., Kainulainen, P., Nissila, M., Nikkanen, H., Kela, M. 2004 Pharmacokinetics of and medroxyprogesterone acetate in different age groups of postmenopausal women. Maturitas, 47, 209-217. Johansson, I.M., Birzniece, V., Lindblad, C., Olsson, T., Backstrom, T. 2002 Allopregnanolone inhibits learning in the Morris water maze. Brain Res, 934, 125-131. Johansson, T., Le Greves, P. 2005 The effect of dehydroepiandrosterone sulfate and allopregnanolone sulfate on the binding of [(3)H] to the N-methyl-d- aspartate receptor in rat frontal cortex membrane. J Steroid Biochem Mol Biol, 94, 263-266. Jung-Testas, I., Renoir, M., Bugnard, H., Greene, G.L., Baulieu, E.E. 1992 Demonstration of receptors and steroid action in primary cultures of rat glial cells. J Steroid Biochem Mol Biol, 41, 621-631.

67

Jussofie, A. 1993a Brain area specific differences in the effects of neuroactive steroids on the GABAA receptor complexes following acute treatment with anaesthetically active steroids. Acta Endocrinol (Copenh), 129, 480-485. Jussofie, A. 1993b Brain region-specific effects of neuroactive steroids on the affinity and density of the GABA-binding site. Biol Chem Hoppe Seyler, 374, 265-270. Kalant, H. 1993 Problems in the search for mechanisms of tolerance. Alcohol Alcohol Suppl, 2, 1-8. Kalant, H., LeBlanc, A.E., Gibbins, R.J. 1971 Tolerance to, and dependence on, some non-opiate psychotropic drugs. Pharmacol Rev, 23, 135-191. Kan, C.C., Hilberink, S.R., Breteler, M.H. 2004 Determination of the main risk factors for benzodiazepine dependence using a multivariate and multidimensional approach. Compr Psychiatry, 45, 88-94. Karavolas, H.J., Hodges, D.R. 1990 Neuroendocrine metabolism of progesterone and related progestins. Ciba Found Symp, 153, 22-44; discussion 44-55. Khan, Z.U., Gutierrez, A., De Blas, A.L. 1996 The alpha 1 and alpha 6 subunits can coexist in the same cerebellar GABAA receptor maintaining their individual benzodiazepine-binding specificities. J Neurochem, 66, 685-691. Klaiber, E.L., Broverman, D.M., Vogel, W., Peterson, L.G., Snyder, M.B. 1997 Relationships of serum estradiol levels, menopausal duration, and mood during hormonal replacement therapy. Psychoneuroendocrinology, 22, 549-558. Knoflach, F., Benke, D., Wang, Y., Scheurer, L., Luddens, H., Hamilton, B.J., Carter, D.B., Mohler, H., Benson, J.A. 1996 Pharmacological modulation of the diazepam-insensitive recombinant gamma-aminobutyric acidA receptors alpha 4 beta 2 gamma 2 and alpha 6 beta 2 gamma 2. Mol Pharmacol, 50, 1253-1261. Kokate, T.G., Yamaguchi, S., Pannell, L.K., Rajamani, U., Carroll, D.M., Grossman, A.B., Rogawski, M.A. 1998 Lack of anticonvulsant tolerance to the neuroactive steroid pregnanolone in mice. J Pharmacol Exp Ther, 287, 553-558. Koksma, J.J., van Kesteren, R.E., Rosahl, T.W., Zwart, R., Smit, A.B., Luddens, H., Brussaard, A.B. 2003 Oxytocin regulates neurosteroid modulation of GABA(A) receptors in supraoptic nucleus around parturition. J Neurosci, 23, 788-797. Korkmaz, S., Wahlstrom, G. 1997 The EEG burst suppression threshold test for the determination of CNS sensitivity to intravenous in rats. Brain Res Brain Res Protoc, 1, 378-384. Korneyev, A., Costa, E. 1996 Allopregnanolone (THP) mediates anesthetic effects of progesterone in rat brain. Horm Behav, 30, 37-43. Korneyev, A., Guidotti, A., Costa, E. 1993 Regional and interspecies differences in brain progesterone metabolism. J Neurochem, 61, 2041-2047. Korpi, E.R., Kleingoor, C., Kettenmann, H., Seeburg, P.H. 1993 Benzodiazepine-induced motor impairment linked to point mutation in cerebellar GABAA receptor. Nature, 361, 356-359. Krieger, N.R., Scott, R.G. 1984 3 alpha-Hydroxysteroid oxidoreductase in rat brain. J Neurochem, 42, 887-890. Kroboth, P.D., Bertz, R.J., Smith, R.B. 1993 Acute tolerance to triazolam during continuous and step infusions: estimation of the effect offset rate constant. J Pharmacol Exp Ther, 264, 1047-1055.

68

Kumar, S., Sieghart, W., Morrow, A.L. 2002 Association of protein kinase C with GABA(A) receptors containing alpha1 and alpha4 subunits in the cerebral cortex: selective effects of chronic ethanol consumption. J Neurochem, 82, 110-117. Ladurelle, N., Eychenne, B., Denton, D., Blair-West, J., Schumacher, M., Robel, P., Baulieu, E. 2000 Prolonged intracerebroventricular infusion of neurosteroids affects cognitive performances in the mouse. Brain Res, 858, 371-379. Lambert, J.J., Belelli, D., Peden, D.R., Vardy, A.W., Peters, J.A. 2003 Neurosteroid modulation of GABAA receptors. Prog Neurobiol, 71, 67-80. Lancel, M., Faulhaber, J., Schiffelholz, T., Romeo, E., Di Michele, F., Holsboer, F., Rupprecht, R. 1997 Allopregnanolone affects sleep in a benzodiazepine-like fashion. J Pharmacol Exp Ther, 282, 1213-1218. Landgren, S., Wang, M.D., Backstrom, T., Johansson, S. 1998 Interaction between 3 alpha-hydroxy-5 alpha-pregnan-20-one and in the control of neuronal excitability in hippocampal slices of female rats in defined phases of the oestrus. Acta Physiol Scand, 162, 77-88. Larsson, J.E., Wahlstrom, G. 1996 Age-dependent development of acute tolerance to propofol and its distribution in a pharmacokinetic compartment-independent rat model. Acta Anaesthesiol Scand, 40, 734-740. LeBlanc, A.E., Kalant, H., Gibbins, R.J., Berman, N.D. 1969 Acquisition and loss of tolerance to ethanol by the rat. J Pharmacol Exp Ther, 168, 244-250. Lindner, M.D. 1997 Reliability, distribution, and validity of age-related cognitive deficits in the Morris water maze. Neurobiol Learn Mem, 68, 203-220. Lister, R.G. 1985 The amnesic action of benzodiazepines in man. Neurosci Biobehav Rev, 9, 87-94. Lovick, T.A., Griffiths, J.L., Dunn, S.M., Martin, I.L. 2005 Changes in GABA(A) receptor subunit expression in the midbrain during the oestrous cycle in Wistar rats. Neuroscience, 131, 397-405. Luisi, S., Petraglia, F., Benedetto, C., Nappi, R.E., Bernardi, F., Fadalti, M., Reis, F.M., Luisi, M., Genazzani, A.R. 2000 Serum allopregnanolone levels in pregnant women: changes during pregnancy, at delivery, and in hypertensive patients. J Clin Endocrinol Metab, 85, 2429-2433. Lundgren, P., Stromberg, J., Backstrom, T., Wang, M. 2003 Allopregnanolone-stimulated GABA-mediated chloride ion flux is inhibited by 3beta-hydroxy-5alpha-pregnan- 20-one (isoallopregnanolone). Brain Res, 982, 45-53. Luscher, B., Keller, C.A. 2004 Regulation of GABAA receptor trafficking, channel activity, and functional plasticity of inhibitory synapses. Pharmacol Ther, 102, 195-221. Macdonald, R., Barker, J.L. 1978 Benzodiazepines specifically modulate GABA- mediated postsynaptic inhibition in cultured mammalian neurones. Nature, 271, 563-564. Macdonald, R.L., Olsen, R.W. 1994 GABAA receptor channels. Annu Rev Neurosci, 17, 569-602. Mahmoudi, M., Kang, M.H., Tillakaratne, N., Tobin, A.J., Olsen, R.W. 1997 Chronic intermittent ethanol treatment in rats increases GABA(A) receptor alpha4-subunit expression: possible relevance to alcohol dependence. J Neurochem, 68, 2485- 2492.

69

Maitra, R., Reynolds, J.N. 1998 Modulation of GABA(A) receptor function by neuroactive steroids: evidence for heterogeneity of steroid sensitivity of recombinant GABA(A) receptor isoforms. Can J Physiol Pharmacol, 76, 909- 920. Maitra, R., Reynolds, J.N. 1999 Subunit dependent modulation of GABAA receptor function by neuroactive steroids. Brain Res, 819, 75-82. Majewska, M.D. 1987 Antagonist-type interaction of glucocorticoids with the GABA receptor-coupled chloride channel. Brain Res, 418, 377-382. Majewska, M.D. 1990 Steroid regulation of the GABAA receptor: ligand binding, chloride transport and behaviour. Ciba Found Symp, 153, 83-97; discussion 97- 106. Majewska, M.D. 1992 Neurosteroids: endogenous bimodal modulators of the GABAA receptor. Mechanism of action and physiological significance. Prog Neurobiol, 38, 379-395. Majewska, M.D., Bisserbe, J.C., Eskay, R.L. 1985 Glucocorticoids are modulators of GABAA receptors in brain. Brain Res, 339, 178-182. Majewska, M.D., Demirgoren, S., Spivak, C.E., London, E.D. 1990 The neurosteroid dehydroepiandrosterone sulfate is an allosteric antagonist of the GABAA receptor. Brain Res, 526, 143-146. Majewska, M.D., Harrison, N.L., Schwartz, R.D., Barker, J.L., Paul, S.M. 1986 Steroid hormone metabolites are -like modulators of the GABA receptor. Science, 232, 1004-1007. Majewska, M.D., Mienville, J.M., Vicini, S. 1988 Neurosteroid pregnenolone sulfate antagonizes electrophysiological responses to GABA in neurons. Neurosci Lett, 90, 279-284. Man, M.S., MacMillan, I., Scott, J., Young, A.H. 1999 Mood, neuropsychological function and cognitions in premenstrual dysphoric disorder. Psychol Med, 29, 727-733. Martini, L., Melcangi, R.C., Maggi, R. 1993 and progesterone metabolism in the central and peripheral nervous system. J Steroid Biochem Mol Biol, 47, 195- 205. Matthews, D.B., Devaud, L.L., Fritschy, J.M., Sieghart, W., Morrow, A.L. 1998 Differential regulation of GABA(A) receptor gene expression by ethanol in the rat hippocampus versus cerebral cortex. J Neurochem, 70, 1160-1166. Matthews, D.B., Morrow, A.L. 2000 Effects of acute and chronic ethanol exposure on spatial cognitive processing and hippocampal function in the rat. Hippocampus, 10, 122-130. Matthews, D.B., Morrow, A.L., Tokunaga, S., McDaniel, J.R. 2002 Acute ethanol administration and acute allopregnanolone administration impair spatial memory in the Morris water task. Alcohol Clin Exp Res, 26, 1747-1751. Mayo, W., Dellu, F., Robel, P., Cherkaoui, J., Le Moal, M., Baulieu, E.E., Simon, H. 1993 Infusion of neurosteroids into the nucleus basalis magnocellularis affects cognitive processes in the rat. Brain Res, 607, 324-328. McEwen, B.S. 1994 Steroid hormone actions on the brain: when is the genome involved? Horm Behav, 28, 396-405.

70

McEwen, B.S., Biegon, A., Davis, P.G., Krey, L.C., Luine, V.N., McGinnis, M.Y., Paden, C.M., Parsons, B., Rainbow, T.C. 1982 Steroid hormones: humoral signals which alter brain cell properties and functions. Recent Prog Horm Res, 38, 41-92. McKernan, R.M., Rosahl, T.W., Reynolds, D.S., Sur, C., Wafford, K.A., Atack, J.R., Farrar, S., Myers, J., Cook, G., Ferris, P., Garrett, L., Bristow, L., Marshall, G., Macaulay, A., Brown, N., Howell, O., Moore, K.W., Carling, R.W., Street, L.J., Castro, J.L., Ragan, C.I., Dawson, G.R., Whiting, P.J. 2000 Sedative but not anxiolytic properties of benzodiazepines are mediated by the GABA(A) receptor alpha1 subtype. Nat Neurosci, 3, 587-592. McKernan, R.M., Whiting, P.J. 1996 Which GABAA-receptor subtypes really occur in the brain? Trends Neurosci, 19, 139-143. McNamara, R.K., dePape, G.E., Skelton, R.W. 1993 Differential effects of benzodiazepine receptor agonists on hippocampal long-term potentiation and spatial learning in the Morris water maze. Brain Res, 626, 63-70. McNamara, R.K., Skelton, R.W. 1993 The neuropharmacological and neurochemical basis of place learning in the Morris water maze. Brain Res Brain Res Rev, 18, 33-49. McNamara, R.K., Skelton, R.W. 1997 Tolerance develops to the spatial learning deficit produced by diazepam in rats. Pharmacol Biochem Behav, 56, 383-389. Mehta, A.K., Shank, R.P. 1995 Interaction of abecarnil, , and RO 19-8022 with diazepam-sensitive and -insensitive benzodiazepine sites in the rat cerebellum and cerebral cortex. Life Sci, 57, 2215-2222. Mehta, A.K., Ticku, M.K. 1999 An update on GABAA receptors. Brain Res Brain Res Rev, 29, 196-217. Mellanby, E. 1919 Alcohol: Its absorbtion into and disappearance from the blood under different conditions. Special report series. No. 31. Medical Research Committee, London Mellon, S.H., Griffin, L.D. 2002 Neurosteroids: biochemistry and clinical significance. Trends Endocrinol Metab, 13, 35-43. Mendelson, J.H. 1964 Alcoholism: some current issues in therapy and research. J Chronic Dis, 17, 645-652. Mendelson, J.H., Ladou, J., Solomon, P. 1964 [Experimentally Induced Chronic Alcoholic Intoxication and Its Suspension. Ii. Psychiatric Results]. Hippokrates, 35, 297-302. Mhatre, M.C., Ticku, M.K. 1992 Chronic ethanol administration alters gamma- aminobutyric acidA receptor gene expression. Mol Pharmacol, 42, 415-422. Miczek, K.A., Fish, E.W., De Bold, J.F. 2003 Neurosteroids, GABAA receptors, and escalated aggressive behavior. Horm Behav, 44, 242-257. Mihalek, R.M., Banerjee, P.K., Korpi, E.R., Quinlan, J.J., Firestone, L.L., Mi, Z.P., Lagenaur, C., Tretter, V., Sieghart, W., Anagnostaras, S.G., Sage, J.R., Fanselow, M.S., Guidotti, A., Spigelman, I., Li, Z., DeLorey, T.M., Olsen, R.W., Homanics, G.E. 1999 Attenuated sensitivity to neuroactive steroids in gamma-aminobutyrate type A receptor delta subunit knockout mice. Proc Natl Acad Sci U S A, 96, 12905-12910.

71

Monteleone, P., Luisi, S., Tonetti, A., Bernardi, F., Genazzani, A.D., Luisi, M., Petraglia, F., Genazzani, A.R. 2000 Allopregnanolone concentrations and premenstrual syndrome. Eur J Endocrinol, 142, 269-273. Montgomery, J.C., Appleby, L., Brincat, M., Versi, E., Tapp, A., Fenwick, P.B., Studd, J.W. 1987 Effect of oestrogen and testosterone implants on psychological disorders in the climacteric. Lancet, 1, 297-299. Morris, R. 1984 Developments of a water-maze procedure for studying spatial learning in the rat. J Neurosci Methods, 11, 47-60. Morrow, A.L., Montpied, P., Lingford-Hughes, A., Paul, S.M. 1990a Chronic ethanol and pentobarbital administration in the rat: effects on GABAA receptor function and expression in brain. Alcohol, 7, 237-244. Morrow, A.L., Pace, J.R., Purdy, R.H., Paul, S.M. 1990b Characterization of steroid interactions with gamma-aminobutyric acid receptor-gated chloride ion channels: evidence for multiple steroid recognition sites. Mol Pharmacol, 37, 263-270. Morrow, A.L., Suzdak, P.D., Paul, S.M. 1987 Steroid hormone metabolites potentiate GABA receptor-mediated chloride ion flux with nanomolar potency. Eur J Pharmacol, 142, 483-485. Muller, R.U., Kubie, J.L., Ranck, J.B., Jr. 1987 Spatial firing patterns of hippocampal complex-spike cells in a fixed environment. J Neurosci, 7, 1935-1950. Myhrer, T. 2003 Neurotransmitter systems involved in learning and memory in the rat: a meta-analysis based on studies of four behavioral tasks. Brain Res Brain Res Rev, 41, 268-287. Norberg, L., Wahlstrom, G., Backstrom, T. 1987 The anaesthetic potency of 3 alpha- hydroxy-5 alpha-pregnan-20-one and 3 alpha-hydroxy-5 beta-pregnan-20-one determined with an intravenous EEG-threshold method in male rats. Pharmacol Toxicol, 61, 42-47. Nyberg, S., Wahlstrom, G., Backstrom, T., Sundstrom Poromaa, I. 2004 Altered sensitivity to alcohol in the late luteal phase among patients with premenstrual dysphoric disorder. Psychoneuroendocrinology, 29, 767-777. O'Brien C, P. 2005 Benzodiazepine use, abuse, and dependence. J Clin Psychiatry, 66 Suppl 2, 28-33. O'Dell, L.E., Purdy, R.H., Covey, D.F., Richardson, H.N., Roberto, M., Koob, G.F. 2005 Epipregnanolone and a novel synthetic neuroactive steroid reduce alcohol self- administration in rats. Pharmacol Biochem Behav, 81, 543-550. O'Keefe, J., Dostrovsky, J. 1971 The hippocampus as a spatial map. Preliminary evidence from unit activity in the freely-moving rat. Brain Res, 34, 171-175. O´Keefe, J., Nadal, L. 1978 The Hippocampus as a cognitive Map. Clarendon Press, Oxford Olsen, R.W., Tobin, A.J. 1990 Molecular biology of GABAA receptors. Faseb J, 4, 1469-1480. Olton, D.S., Papas, B.C. 1979 Spatial memory and hippocampal function. Neuropsychologia, 17, 669-682. Ottander, U., Poromaa, I.S., Bjurulf, E., Skytt, A., Backstrom, T., Olofsson, J.I. 2005 Allopregnanolone and pregnanolone are produced by the human corpus luteum. Mol Cell Endocrinol, 239, 37-44.

72

Palmer, A.A., Moyer, M.R., Crabbe, J.C., Phillips, T.J. 2002 Initial sensitivity, tolerance and cross-tolerance to allopregnanolone- and ethanol-induced hypothermia in selected mouse lines. Psychopharmacology (Berl), 162, 313-322. Panay, N., Studd, J. 1997 intolerance and compliance with hormone replacement therapy in menopausal women. Hum Reprod Update, 3, 159-171. Paul, S.M., Purdy, R.H. 1992 Neuroactive steroids. Faseb J, 6, 2311-2322. Peng, Z., Hauer, B., Mihalek, R.M., Homanics, G.E., Sieghart, W., Olsen, R.W., Houser, C.R. 2002 GABA(A) receptor changes in delta subunit-deficient mice: altered expression of alpha4 and gamma2 subunits in the forebrain. J Comp Neurol, 446, 179-197. Phoenix, C.H., Goy, R.W., Gerall, A.A., Young, W.C. 1959 Organizing action of prenatally administered on the tissues mediating mating behavior in the female guinea pig. Endocrinology, 65, 369-382. Poisbeau, P., Williams, S.R., Mody, I. 1997 Silent GABAA synapses during withdrawal are region-specific in the hippocampal formation. J Neurosci, 17, 3467-3475. Postma, A., Winkel, J., Tuiten, A., van Honk, J. 1999 Sex differences and menstrual cycle effects in human spatial memory. Psychoneuroendocrinology, 24, 175-192. Pregenzer, J.F., Im, W.B., Carter, D.B., Thomsen, D.R. 1993 Comparison of interactions of [3H]muscimol, t-butylbicyclophosphoro[35S]thionate, and [3H]flunitrazepam with cloned gamma-aminobutyric acidA receptors of the alpha 1 beta 2 and alpha 1 beta 2 gamma 2 subtypes. Mol Pharmacol, 43, 801-806. Purdy, R.H., Moore, P.H., Jr., Rao, P.N., Hagino, N., Yamaguchi, T., Schmidt, P., Rubinow, D.R., Morrow, A.L., Paul, S.M. 1990a Radioimmunoassay of 3 alpha- hydroxy-5 alpha-pregnan-20-one in rat and human plasma. Steroids, 55, 290-296. Purdy, R.H., Morrow, A.L., Blinn, J.R., Paul, S.M. 1990b Synthesis, metabolism, and pharmacological activity of 3 alpha-hydroxy steroids which potentiate GABA- receptor-mediated chloride ion uptake in rat cerebral cortical synaptoneurosomes. J Med Chem, 33, 1572-1581. Purdy, R.H., Morrow, A.L., Moore, P.H., Jr., Paul, S.M. 1991 Stress-induced elevations of gamma-aminobutyric acid type A receptor-active steroids in the rat brain. Proc Natl Acad Sci U S A, 88, 4553-4557. Quirk, K., Whiting, P.J., Ragan, C.I., McKernan, R.M. 1995 Characterisation of delta- subunit containing GABAA receptors from rat brain. Eur J Pharmacol, 290, 175- 181. Raffalli-Sebille, M.J., Chapouthier, G. 1991 Similar effects of a beta-carboline and of flumazenil in negatively and positively reinforced learning tasks in mice. Life Sci, 48, 685-692. Raisinghani, K.H., Dorfman, R.I., Forchielli, E., Gyermek, L., Genther, G. 1968 Uptake of intravenously administered progesterone, pregnanedione and pregnanolone by the rat brain. Acta Endocrinol (Copenh), 57, 395-404. Ramsey-Williams, V.A., Carter, D.B. 1996 Chronic triazolam and its withdrawal alters GABAA receptor subunit mRNA levels: an in situ hybridization study. Brain Res Mol Brain Res, 43, 132-140. Rasgon, N.L., Altshuler, L.L., Fairbanks, L.A., Dunkin, J.J., Davtyan, C., Elman, S., Rapkin, A.J. 2002 Estrogen replacement therapy in the treatment of major

73

depressive disorder in perimenopausal women. J Clin Psychiatry, 63 Suppl 7, 45- 48. Reddy, D.S. 2002 The clinical potentials of endogenous neurosteroids. Drugs Today (Barc), 38, 465-485. Reddy, D.S. 2004 Role of neurosteroids in catamenial epilepsy. Epilepsy Res, 62, 99-118. Reddy, D.S., Kulkarni, S.K. 1998 The effects of neurosteroids on acquisition and retention of a modified passive-avoidance learning task in mice. Brain Res, 791, 108-116. Reddy, D.S., Kulkarni, S.K. 1999 Sex and estrous cycle-dependent changes in neurosteroid and benzodiazepine effects on food consumption and plus-maze learning behaviors in rats. Pharmacol Biochem Behav, 62, 53-60. Reddy, D.S., Kulkarni, S.K. 2000 Development of neurosteroid-based novel psychotropic drugs. Prog Med Chem, 37, 135-175. Reynolds, D.S., Rosahl, T.W., Cirone, J., O'Meara, G.F., Haythornthwaite, A., Newman, R.J., Myers, J., Sur, C., Howell, O., Rutter, A.R., Atack, J., Macaulay, A.J., Hadingham, K.L., Hutson, P.H., Belelli, D., Lambert, J.J., Dawson, G.R., McKernan, R., Whiting, P.J., Wafford, K.A. 2003 Sedation and mediated by distinct GABA(A) receptor isoforms. J Neurosci, 23, 8608-8617. Riedel, G. 1996 Function of metabotropic glutamate receptors in learning and memory. Trends Neurosci, 19, 219-224. Rosenberg, M.J., Waugh, M.S. 1998 Oral contraceptive discontinuation: a prospective evaluation of frequency and reasons. Am J Obstet Gynecol, 179, 577-582. Roy-Byrne, P.P. 2005 The GABA-benzodiazepine receptor complex: structure, function, and role in anxiety. J Clin Psychiatry, 66 Suppl 2, 14-20. Rumke, C.L. 1963 The influence of drugs on the duration of hexobarbital and narcosis in mice. Naunyn Schmiedebergs Arch Exp Pathol Pharmakol, 244, 519-530. Rupprecht, R. 2003 Neuroactive steroids: mechanisms of action and neuropsychopharmacological properties. Psychoneuroendocrinology, 28, 139- 168. Rupprecht, R., Holsboer, F. 1999 Neuroactive steroids: mechanisms of action and neuropsychopharmacological perspectives. Trends Neurosci, 22, 410-416. Saunders, P.A., Ito, Y., Baker, M.L., Hume, A.S., Ho, I.K. 1990 Pentobarbital tolerance and withdrawal: correlation with effects on the GABAA receptor. Pharmacol Biochem Behav, 37, 343-348. Schmidt, B.M., Gerdes, D., Feuring, M., Falkenstein, E., Christ, M., Wehling, M. 2000 Rapid, nongenomic steroid actions: A new age? Front Neuroendocrinol, 21, 57- 94. Schofield, P.R., Darlison, M.G., Fujita, N., Burt, D.R., Stephenson, F.A., Rodriguez, H., Rhee, L.M., Ramachandran, J., Reale, V., Glencorse, T.A., et al. 1987 Sequence and functional expression of the GABA A receptor shows a ligand-gated receptor super-family. Nature, 328, 221-227. Schumacher, M., Weill-Engerer, S., Liere, P., Robert, F., Franklin, R.J., Garcia-Segura, L.M., Lambert, J.J., Mayo, W., Melcangi, R.C., Parducz, A., Suter, U., Carelli, C., Baulieu, E.E., Akwa, Y. 2003 Steroid hormones and neurosteroids in normal and pathological aging of the nervous system. Prog Neurobiol, 71, 3-29.

74

Schwabe, K., Gavrilovici, C., McIntyre, D.C., Poulter, M.O. 2005 Neurosteroids exhibit differential effects on mIPSCs recorded from normal and seizure prone rats. J Neurophysiol, 94, 2171-2181. Schwartz, R.D., Wess, M.J., Labarca, R., Skolnick, P., Paul, S.M. 1987 Acute stress enhances the activity of the GABA receptor-gated chloride ion channel in brain. Brain Res, 411, 151-155. Scoville, W.B., Milner, B. 1957 Loss of recent memory after bilateral hippocampal lesions. J Neurol Neurosurg Psychiatry, 20, 11-21. Serra, G.B. 1983 The ovary. Raven Press., New York, USA. Sherman, A., Gebhart, G.F. 1974 Regional levels of GABA and glutamate in mouse brain following exposure to pain. Neuropharmacology, 13, 673-675. Shumaker, S.A., Legault, C., Kuller, L., Rapp, S.R., Thal, L., Lane, D.S., Fillit, H., Stefanick, M.L., Hendrix, S.L., Lewis, C.E., Masaki, K., Coker, L.H. 2004 Conjugated equine and incidence of probable dementia and mild cognitive impairment in postmenopausal women: Women's Health Initiative Memory Study. Jama, 291, 2947-2958. Shumaker, S.A., Legault, C., Rapp, S.R., Thal, L., Wallace, R.B., Ockene, J.K., Hendrix, S.L., Jones, B.N., 3rd, Assaf, A.R., Jackson, R.D., Kotchen, J.M., Wassertheil- Smoller, S., Wactawski-Wende, J. 2003 Estrogen plus progestin and the incidence of dementia and mild cognitive impairment in postmenopausal women: the Women's Health Initiative Memory Study: a randomized controlled trial. Jama, 289, 2651-2662. Simon, P., Dupuis, R., Costentin, J. 1994 Thigmotaxis as an index of anxiety in mice. Influence of dopaminergic transmissions. Behav Brain Res, 61, 59-64. Singh, H.C., Singh, R.H., Udupa, K.N. 1979 Effect of feet electroshock on GABA glutamate content in rat brain. Indian J Exp Biol, 17, 418-419. Slany, A., Zezula, J., Tretter, V., Sieghart, W. 1995 Rat beta 3 subunits expressed in human embryonic 293 cells form high affinity [35S]t- butylbicyclophosphorothionate binding sites modulated by several allosteric ligands of gamma-aminobutyric acid type A receptors. Mol Pharmacol, 48, 385- 391. Song, M., Messing, R.O. 2005 Protein kinase C regulation of GABAA receptors. Cell Mol Life Sci, 62, 119-127. Speroff, L., Glass R.H., Kase N.G. 1999 Clinical gynecologic endocrinology and infertility. Lippincott Williams & Wilkins. Stromberg, J., Backstrom, T., Lundgren, P. 2005 Rapid non-genomic effect of glucocorticoid metabolites and neurosteroids on the gamma-aminobutyric acid-A receptor. Eur J Neurosci, 21, 2083-2088. Sundstrom, I., Ashbrook, D., Backstrom, T. 1997 Reduced benzodiazepine sensitivity in patients with premenstrual syndrome: a pilot study. Psychoneuroendocrinology, 22, 25-38. Sundstrom, I., Backstrom, T. 1998 Patients with premenstrual syndrome have decreased saccadic eye velocity compared to control subjects. Biol Psychiatry, 44, 755-764. Sundstrom, I., Spigset, O., Andersson, A., Appelblad, P., Backstrom, T. 1999 Lack of influence of menstrual cycle and premenstrual syndrome diagnosis on pregnanolone pharmacokinetics. Eur J Clin Pharmacol, 55, 125-130.

75

Sundstrom Poromaa, I., Smith, S., Gulinello, M. 2003 GABA receptors, progesterone and premenstrual dysphoric disorder. Arch Women Ment Health, 6, 23-41. Sur, C., Farrar, S.J., Kerby, J., Whiting, P.J., Atack, J.R., McKernan, R.M. 1999 Preferential coassembly of alpha4 and delta subunits of the gamma-aminobutyric acidA receptor in rat thalamus. Mol Pharmacol, 56, 110-115. Tehrani, M.H., Baumgartner, B.J., Barnes, E.M., Jr. 1997 Clathrin-coated vesicles from bovine brain contain uncoupled GABAA receptors. Brain Res, 776, 195-203. Tia, S., Wang, J.F., Kotchabhakdi, N., Vicini, S. 1996 Distinct deactivation and desensitization kinetics of recombinant GABAA receptors. Neuropharmacology, 35, 1375-1382. Tietz, E.I., Huang, X., Chen, S., Ferencak, W.F., 3rd 1999 Temporal and regional regulation of alpha1, beta2 and beta3, but not alpha2, alpha4, alpha5, alpha6, beta1 or gamma2 GABA(A) receptor subunit messenger RNAs following one- week oral flurazepam administration. Neuroscience, 91, 327-341. Timby, E., Balgard, M., Nyberg, S., Spigset, O., Andersson, A., Porankiewicz-Asplund, J., Purdy, R.H., Zhu, D., Backstrom, T., Poromaa, I.S. 2005 Pharmacokinetic and behavioral effects of allopregnanolone in healthy women. Psychopharmacology (Berl), 1-11. Tokunaga, S., McDaniel, J.R., Morrow, A.L., Matthews, D.B. 2003 Effect of acute ethanol administration and acute allopregnanolone administration on spontaneous hippocampal pyramidal cell neural activity. Brain Res, 967, 273-280. Tomson, T. 1997 Epilepsy and Pregnancy. Wrightson Biomedical Publishing Ltd., Petersfield, UK. pp 113-123 Wahlstrom, G., Norberg, L. 1984 A comparative investigation in the rat of the anesthetic effects of the of two barbiturates. Brain Res, 310, 261-267. van Rijnsoever, C., Tauber, M., Choulli, M.K., Keist, R., Rudolph, U., Mohler, H., Fritschy, J.M., Crestani, F. 2004 Requirement of alpha5-GABAA receptors for the development of tolerance to the sedative action of diazepam in mice. J Neurosci, 24, 6785-6790. Wang, M., He, Y., Eisenman, L.N., Fields, C., Zeng, C.M., Mathews, J., Benz, A., Fu, T., Zorumski, E., Steinbach, J.H., Covey, D.F., Zorumski, C.F., Mennerick, S. 2002 3beta -hydroxypregnane steroids are pregnenolone sulfate-like GABA(A) receptor antagonists. J Neurosci, 22, 3366-3375. Wang, M., Seippel, L., Purdy, R.H., Backstrom, T. 1996 Relationship between symptom severity and steroid variation in women with premenstrual syndrome: study on serum pregnenolone, pregnenolone sulfate, 5 alpha-pregnane-3,20-dione and 3 alpha-hydroxy-5 alpha-pregnan-20-one. J Clin Endocrinol Metab, 81, 1076-1082. Wang, M.D., Backstrom, T., Landgren, S. 2000 The inhibitory effects of allopregnanolone and pregnanolone on the population spike, evoked in the rat hippocampal CA1 stratum pyramidale in vitro, can be blocked selectively by epiallopregnanolone. Acta Physiol Scand, 169, 333-341. Vanover, K.E., Suruki, M., Robledo, S., Huber, M., Wieland, S., Lan, N.C., Gee, K.W., Wood, P.L., Carter, R.B. 1999 Positive allosteric modulators of the GABA(A) receptor: differential interaction of benzodiazepines and neuroactive steroids with ethanol. Psychopharmacology (Berl), 141, 77-82.

76

Weir, C.J., Ling, A.T., Belelli, D., Wildsmith, J.A., Peters, J.A., Lambert, J.J. 2004 The interaction of anaesthetic steroids with recombinant glycine and GABAA receptors. Br J Anaesth, 92, 704-711. Whishaw, I.Q. 1985 Formation of a place learning-set by the rat: a new paradigm for neurobehavioral studies. Physiol Behav, 35, 139-143. Whitfield, G.K., Jurutka, P.W., Haussler, C.A., Haussler, M.R. 1999 Steroid hormone receptors: evolution, ligands, and molecular basis of biologic function. J Cell Biochem, Suppl 32-33, 110-122. Whiting, P.J., Bonnert, T.P., McKernan, R.M., Farrar, S., Le Bourdelles, B., Heavens, R.P., Smith, D.W., Hewson, L., Rigby, M.R., Sirinathsinghji, D.J., Thompson, S.A., Wafford, K.A. 1999 Molecular and functional diversity of the expanding GABA-A receptor gene family. Ann N Y Acad Sci, 868, 645-653. Widmer, H., Ludwig, M., Bancel, F., Leng, G., Dayanithi, G. 2003 Neurosteroid regulation of oxytocin and release from the rat supraoptic nucleus. J Physiol, 548, 233-244. Wikler, A. 1956 The uses of drugs in psychiatric research. Am J Psychiatry, 112, 961- 969. Wikler, A., Pescor, F.T., Fraser, H.F., Isbell, H. 1956 Electroencephalographic changes associated with chronic alcoholic intoxication and the alcohol abstinence syndrome. Am J Psychiatry, 113, 106-114. Wilkinson, D. 1995 In Situ Hybridization; A practical approach. Oxford University Press, London Williams, C.L., Meck, W.H. 1991 The organizational effects of gonadal steroids on sexually dimorphic spatial ability. Psychoneuroendocrinology, 16, 155-176. Wisden, W., Laurie, D.J., Monyer, H., Seeburg, P.H. 1992 The distribution of 13 GABAA receptor subunit mRNAs in the rat brain. I. Telencephalon, diencephalon, mesencephalon. J Neurosci, 12, 1040-1062. Wong, G., Sei, Y., Skolnick, P. 1992 Stable expression of type I gamma-aminobutyric acidA/benzodiazepine receptors in a transfected cell line. Mol Pharmacol, 42, 996-1003. Woodbury, D.M. 1952 Effect of adrenocortical steroids and adrenocorticotrophic hormone on electroshock seizure threshold. J Pharmacol Exp Ther, 105, 27-36. Woolley, C.S., McEwen, B.S. 1993 Roles of estradiol and progesterone in regulation of hippocampal dendritic spine density during the estrous cycle in the rat. J Comp Neurol, 336, 293-306. Wu, Y., Rosenberg, H.C., Chiu, T.H., Zhao, T.J. 1994 Subunit- and brain region-specific reduction of GABAA receptor subunit mRNAs during chronic treatment of rats with diazepam. J Mol Neurosci, 5, 105-120. Yu, R., Follesa, P., Ticku, M.K. 1996a Down-regulation of the GABA receptor subunits mRNA levels in mammalian cultured cortical neurons following chronic neurosteroid treatment. Brain Res Mol Brain Res, 41, 163-168. Yu, R., Hay, M., Ticku, M.K. 1996b Chronic neurosteroid treatment attenuates single cell GABAA response and its potentiation by modulators in cortical neurons. Brain Res, 706, 160-162. Yu, R., Ticku, M.K. 1995a Chronic neurosteroid treatment decreases the efficacy of benzodiazepine ligands and neurosteroids at the gamma-aminobutyric acidA

77

receptor complex in mammalian cortical neurons. J Pharmacol Exp Ther, 275, 784-789. Yu, R., Ticku, M.K. 1995b Chronic neurosteroid treatment produces functional heterologous uncoupling at the gamma-aminobutyric acid type A/benzodiazepine receptor complex in mammalian cortical neurons. Mol Pharmacol, 47, 603-610. Zhu, D., Birzniece, V., Backstrom, T., Wahlstrom, G. 2004 Dynamic aspects of acute tolerance to allopregnanolone evaluated using anaesthesia threshold in male rats. Br J Anaesth, 93, 560-567. Zhu, D., Wang, M.D., Backstrom, T., Wahlstrom, G. 2001 Evaluation and comparison of the pharmacokinetic and pharmacodynamic properties of allopregnanolone and pregnanolone at induction of anaesthesia in the male rat. Br J Anaesth, 86, 403- 412.

78

APPENDIX -Papers I-IV

79

Umeå University Medical Dissertations New Series No 1026 – ISSN 0346-6612 – ISBN 91-7264-073-1

Edited by the Dean of the Faculty of Medicine

Umeå University WWW.UMU.SE 2006

80