<<

View metadata, citation and similar papers at core.ac.uk brought to you by CORE REVIEW ARTICLE published: 04 October 2011 provided by PubMed Central doi: 10.3389/fendo.2011.00044 and GABA-A

Mingde Wang*

Section of Obstetrics and Gynecology, Department of Clinical Science, Umeå Research Center, Umeå University, Umeå, Sweden

Edited by: Neurosteroids represent a class of endogenous that are synthesized in the brain, Hubert Vaudry, University of Rouen, the adrenals, and the gonads and have potent and selective effects on the GABAA- France receptor. 3α-hydroxy A-ring reduced of , deoxycorticosterone, Reviewed by: Valerio Magnaghi, Università Degli and are positive modulators of GABAA-receptor in a non-genomic man- Studi di Milano, Italy ner. (3α-OH-5α-pregnan-20-one), 5α--3α,17α-diol (Adiol), and Giovanni Biggio, University of Cagliari, 3α5α-tetrahydrodeoxycorticosterone (3α5α-THDOC) enhance the GABA-mediated Cl- cur- Italy rents acting on a site (or sites) distinct from the GABA, , , and Remy Schlichter, University α α α α Strasbourg – CNRS, France binding sites. 3 5 -P and 3 5 -THDOC potentiate synaptic GABAA-receptor δ *Correspondence: function and activate -subunit containing extrasynaptic receptors that mediate tonic Mingde Wang, Section of Obstetrics currents. On the contrary, 3β-OH steroids and (PS) are and Gynecology, Department of GABAA-receptor antagonists and induce activation-dependent inhibition of the receptor. Clinical Science, Umeå Neurosteroid The activities of neurosteroid are dependent on brain regions and types of . In Research Center, Umeå University, 901 85 Umeå, Sweden. addition to the slow genomic action of the parent steroids, the non-genomic, and rapid e-mail: [email protected] actions of neurosteroids play a significant role in the GABAA-receptor function and shift in mood and function. This review describes molecular mechanisms underlying neurosteroid action on the GABAA-receptor, mood changes, and cognitive functions.

Keywords: allopregnanolone, THDOC, , GABAA-receptor, premenstrual dysphoric disorder, mood, cognition

NEUROSTEROID in the CNS and alter neuronal excitability (Majewska et al., 1986; Sex hormones act through genomic mechanisms through the Paul and Purdy, 1992; Lambert et al., 1995). intracellular receptors located in the nucleus or cytoplasm. They The term “neurosteroid” was introduced to describe these act as -activated transcription factors in the regulation of metabolites that modulate neuronal activity (Paul gene expression. However, metabolites of progesterone and sev- and Purdy, 1992; Baulieu et al., 2007; Mellon, 2007). The eral stress hormones act on the membrane bound receptor via a 3α-hydroxy A-ring reduced metabolites of progesterone and non-genomic mechanism. The receptor binding to DNA and RNA deoxycorticosterone, allopregnanolone (3α5α-P), and 3α,5α- synthesis is thus not required (Frye et al., 1992; Baulieu and Robel, tetrahydrodeoxycorticosterone (3α5α-THDOC) were first shown 1995; Rupprecht, 2003). While the genomic action of sex hor- to modulate neuronal excitability by interaction with the GABAA- mones requires a time period from minutes to hours and limited receptor (Majewska et al., 1986). Several other neurotransmit- by the rate of (McEwen, 1991), the modula- ters like the NMDA, nicotinic, muscarinic, , adren- tion on the membrane receptor is fast occurring event and requires ergic, and sigma 1 receptors are also targets for neurosteroids only milliseconds to seconds (McEwen, 1991). Today it is known (Klangkalya and Chan, 1988; Wu et al., 1991; Valera et al., 1992; that metabolites of sex and stress hormones act non-genomically Compagnone and Mellon, 2000; Mellon et al., 2001; Parry, 2001; Halbreich, 2003; Mellon, 2007). The functional modulation of the GABAA-receptor by neurosteroids at low is believed to induce moderate to severe adverse mood changes in Abbreviations: 3β5β-P, 3β-OH-5β-pregnan-20-one, ; 3β5α-P, 3β-OH-5α-pregnan-20-one, isoallopregnanolone; 3α5α-P, 3α-OH-5α-pregnan- up to 20% of female individuals (Beauchamp et al., 2000; Fish 20-one, allopregnanolone; 3α5β-P, 3α-OH-5β-pregnan-20-one, ; et al., 2001). The clinical complex of premenstrual dystrophic 3α5α-THDOC, 3α,5α-tetrahydrodeoxycorticosterone; 3α5β-THDOC, 3α,5β- disorders (PMDD; Backstrom et al., 2003; Sundstrom Poromaa tetrahydrodeoxycorticosterone; 3β5β-THDOC, 5β-pregnan-3β, 21-diol-20-one; et al., 2003), petit mal (Grunewald et al., 1992; Baner- α α α α α β β β 3 -adiol/3 5 -ADL, 5 -androstane-3 ,17-diol; 3 -HSD, 3 -hydroxysteroid jee and Snead, 1998), and catamenial epilepsy (Backstrom, 1976) dehydrogenase; 3α-OH, 3α-hydroxy; 3β-OH, 3β-hydroxy; CGC, cerebellar gran- ule cells; CNS, central ; DGC, dentate gyrus granule cells; DHEAS, are among the disorders that may involve neurosteroid action. At sulfate; DOC, deoxycorticosterone; EC50, higher doses, neurosteroids may affect (Johansson et al., of that produces 50% of maximum response; GABA, γ-amino ; 2002), act as , anti-aggressive, /, and LTP, long-term potentiation; MPA, ; MPN, medial preoptic anti-epileptic agents in both animals and humans (Backstrom nucleus; nAChRs, nicotinic receptors; NMDA, N -methyl-d-aspartate; et al., 1990; Paul and Purdy, 1992; Wang et al., 2001; Bjorn et al., P45021, steroid 21 hydroxylase; P450c17, steroid 17 alpha-hydroxylase/17, 20 lyase; PMS/PMDD, /premenstrual dysphoric disorder; PS, preg- 2002). nenolone sulfate; sIPSC,spontaneous inhibitory postsynaptic current; TBPS,t-butyl As shown in Figure 1, neurosteroids are synthesized in glial bicyclophosphorothionate; UC1011, 3β,20β-dihydroxy-5α-pregnane. cells and neurons of the central and peripheral nervous system

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 1 Wang Neurosteroids and GABAA-receptor function

from or steroidal precursors imported from periph- progesterone and this conversion is mediated by P45021 (Edwards eral sources (Schumacher et al., 2000; Baulieu et al., 2007). They et al., 2005). The conversion of DOC to 3α5α-THDOC occurs include 3β-hydroxy-Δ5-compounds such as pregnenolone and both in peripheral tissues and in the brain (Reddy, 2003). The dehydroepiandrosterone, their sulfate esters and reduced metabo- A-ring reduced of testosterone, 5α-androstane-3α,17β- lites of steroid and stress hormones such as the tetrahydroderiv- diol (3α5α-adiol), acts also as a GABAA-receptor (Frye ative of progesterone, 3α-hydroxy-5α-pregnan-20-one (allopreg- et al., 1996). nanolone; Baulieu et al., 2007). Progesterone itself is also a neuros- teroid, and a has been detected in peripheral GABAA-RECEPTOR and central glial cells (Schumacher et al.,2000; Baulieu et al.,2007). GABA mediates most of the inhibitory in the At different sites in the brain, concentrations of neurosteroids vary mammalian brain. GABA-mediated inhibition is crucially in both according to environmental and behavioral circumstances, such as short-term and long-term regulation of neuronal excitability. It stress, sex recognition, and aggressiveness. Allopregnanolone can has been estimated that approximately 33% of the in the accumulate in the brain after adrenalectomy and gonadectomy mammalian are GABAergic (Purvez et al., 2004). (Purdy et al., 1991; Corpechot et al., 1993; Cheney et al., 1995). Two major types of receptors, GABAA- and GABAB-receptors can This indicates that allopregnanolone is synthesized in the brain be identified in the CNS (Sivilotti and Nistri, 1991; Bormann, via A-ring reduction of progesterone (Celotti et al., 1992; Do Rego 2000). It appears that neurosteroids more or less exclusively tar- et al., 2009). Table 1 shows concentrations of major neurosteroids get the GABAA-receptor which is a ligand gated anion-selective in the plasma and brain. channel (Schofield et al., 1987). Pregnenolone sulfate (PS) and dehydroepiandrosterone sul- Various isoforms of the GABAA-receptorhavebeenidentified fate (DHEAS) are naturally occurring neurosteroids that inhibit that comprise α1–6, β1–3, γ1–3, δ, ε, π, θ, and ρ1–3 subunits (Amin the GABAA-receptor (Paul and Purdy, 1992). PS is synthesized and Weiss, 1994; Mehta and Ticku, 1999; Rudolph et al., 2001). In from pregnenolone by the sulfotransferase (Figure 1). general, GABAA-receptors are pentameric (Nayeem et al., Conversion from DHEA to DHEAS is also mediated by sulfotrans- 1994) that are built of five subunits which includes two α-subunits, ferase. DHEA is metabolized from pregnenolone by cytochrome two β-subunits and one subunit of either the γ-,δ-,ε-,π-,or θ-type P450C17 (Mensah-Nyagan et al., 1999). On the other hand, 3β- (Farrar et al., 1999; Knight et al., 2000; Klausberger et al., 2001). HSD is essential for the synthesis of 3β-OH steroids, i.e., 3β-OH- A subgroup of GABAA-receptor channels was named as GABAC- 5α-pregnan-20-one (isoallopregnanolone; 3β5α-P) and 3β-OH- receptor earlier and composed of homo-oligomeric ρ1–3 subunits. 5β-pregnan-20-one (epipregnanolone; 3β5β-P; Stromstedt et al., They are pharmacologically distinct from other GABAA-receptor 1993). PS, DHEAS, and 3β-OH steroids act as antagonists on the channels and this difference is illustrated by the insensitivity of ρ GABAA-receptor (Wang et al., 2002; Eisenman et al., 2003) and receptor channels to many known modulators such as can be measured from human blood samples (Hill et al., 2001). and benzodiazepine (Amin and Weiss, 1994, 1996). As metabolites of stress hormone deoxycorticosterone (DOC), Immunological, pharmacological, and functional analysis give 3α5α-THDOC and 3α5β-THDOC are also potent modulators of the evidence that the α1β2γ2 combination is the most common the GABAA-receptor (Crawley et al., 1986; Majewska et al., 1986; GABAA-receptor within the CNS (∼60%), followed by α2β3γ2 Gasior et al., 1999; Lambert et al., 2001a). Both steroids have sig- (∼15–20%) and α3βnγ2 (∼10–15%, n = 1, 2, or 3; Mohler et al., nificant sedative effects in vivo.3α5α-THDOC is responsible for 2002, 2004; Fritschy and Brunig, 2003; Wallner et al., 2003). Recep- the sedative and anti- activity of DOC in animal mod- tors containing the α4-, α5-, and α6-subnit, as well as the β1-, γ1–3, els (Reddy and Rogawski, 2002). DOC can be metabolized from δ-, π-, and θ-subunit, form a minor receptor population. Each of

Table 1 | Neurosteroid concentrations in human plasma and brain (mean ± SEM).

Steroids Plasma (nM) Brain (nM)

Follicular Luteal Postmenopausal Luteal

Progesterone 5.0 ± 0.50 (Wang et al., 1996) 34.7 ± 2.40 65 (Bixo et al., 1997) 137 (Bixo et al., 1997) 3α-OH-5β-pregnan-20-one 0.6 ± 0.00 (Sundstrom et al., 1998)1.1± 0.50 – 114 (Bixo et al., 1997) 3α-OH-5α-pregnan-20-one 0.2–0.6 (Wang et al., 1996; 2–4 (Wang et al., 1996; 47 (Bixo et al., 1997)66(Bixo et al., 1997) Genazzani et al., 1998) Genazzani et al., 1998) 3β-OH-5α-pregnan-20-one 0.3–0.5 (Wang et al., 1996; 1.5–3.5 (Wang et al., 1996; –– Sundstrom et al., 1998)) Sundstrom et al., 1998) 3β-OH-5β-pregnan-20-one 0.09 ± 0.08 (Havlikova et al., 2006) 0.26 ± 0.13 – – Pregnenolone sulfate 11.2 ± 0.6 (Wang et al., 1996) 15.2 ± 0.8 (Bixo et al., 1997)– 100(Lanthier and Pat- wardhan, 1986) Pregnenolone *2.19 (Kancheva et al., 2007)– – – 3α5α-androstane-3α,17β-diol *0.475 (Kancheva et al., 2007)– – –

Data cited from references are inserted as indicated. *Concentration in adult men.

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 2 Wang Neurosteroids and GABAA-receptor function

FIGURE 1 | Biosynthesis of allopregnanolone and pregnenolone 5α-reductase. Transport of cholesterol across the mitochondrial membrane sulfate (PS) from cholesterol within the or glial . is enhanced by the steroidogenic acute-regulatory (StAR) protein and the involved are P450 side-chain cleavage (P450scc), 3α-hydroxysteroid mitochondrial benzodiazepine receptor (MBR). 5α-DHP represents dehydrogenase (3α-HSD), 3β-hydroxysteroid dehydrogenase (3β-HSD) and 5α-dihydroprogesterone.

the α4βnδ, α4βnγ, and the α6β2/3γ2 receptor accounts less than 5% it is localized to interneurons in the and cortex of all the GABAA-receptor quantity (McKernan andWhiting,1996; (layer I–IV), and cerebral Purkinje cells (McKernan and Whiting, Whiting, 2003a,b). The α6βnδ receptor has a small population in 1996). The α2β3γ2 receptor is present in cerebral cortex (layer I– the cerebellum and the α6β2/3γ2 receptor located exclusively in the IV), hippocampal formation, amygdale, striatum, olfactory bulb, cerebellum (McKernan and Whiting, 1996; Whiting, 2003a,b). , superior colliculi, and motor nuclei (Fritschy and The expression of GABAA-receptor subtypes in the adult brain Brunig, 2003). The α3βnγ2, α3γ2, and α3θ receptors are abun- exhibits a remarkable regional and neuronal specificity which sug- dant in the cerebral cortex (layers V–VI), amygdala, olfactory bulb, gests that individual subtypes are present in distinct neuronal thalamic reticular and intralaminar nuclei, superior colliculus, circuits. The α1β2γ2 receptor is present in most brain areas and brainstem, spinal cord, and locus coeruleus. The α4βnδ (n = 1, 2,

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 3 Wang Neurosteroids and GABAA-receptor function

or 3) receptor is presented in the dentate gyrus and thalamus. The the concept that GABAA-receptor heterogeneity is a major facet α5β3γ2 receptor is widespread in the hippocampus and dentate determining the functional properties of GABAergic inhibitory gyrus (Glykys et al., 2008), deep cortical layers, amygdala, olfactory circuits (Sieghart, 2000; Mohler et al., 2001). In particular, the type bulb, hypothalamus, superior colliculus, superior olivary nucleus, of α-subunit determines the kinetics of receptor deactivation (Ver- spinal trigeminal nucleus, and spinal cord. The α6β2/3γ2, α6β2/3δ, doorn et al., 1990; Hutcheon et al., 2000; Devor et al., 2001), and and α6β2/3γ2 receptors are found mainly in the cerebellum and the presence of the δ-subunit results in markedly increased agonist dorsal cochlear nucleus (Fritschy and Brunig, 2003). affinity and apparent lack of desensitization (Burgard et al., 1996; Fisher and Macdonald, 1997; Adkins et al., 2001). PHARMACOLOGICAL RELEVANCE OF GABAA-RECEPTOR SUBUNIT The GABAA-receptor expresses different subunit compositions COMPOSITIONS in different parts of the brain (Sieghart and Sperk, 2002). The Especially the different α-subunits have been attributed to spe- subunit composition is related to different function of the spe- cific behavioral effects as shown in Table 2. For example, an cific part of the brain (Korpi et al., 2007; Table 2). This anatomical enhancement at the α1 subunit has been associated with seda- diversity constitutes the very basis for the pathogenesis of different tion (Rudolph et al., 1999) and the α2 with anxiolytic action. conditions. Steroids interact differently with the GABAA-receptor Recently, it has been shown that the α5 subunit is important depending on the subunit composition (Belelli et al., 2002). The for sedative tolerance development to and for α5 subunit is localized in high degree in the hippocampus, a key acquisition and expression of associative memory and spatial area for memory and learning, α5 is shown to be important for learning (Collinson et al., 2002; Crestani et al., 2002; van Rijn- learning and memory function (Table 2) since α5-subunit knock- soever et al., 2004; Yee et al., 2004). In addition, the α4 subunit out mice in comparison with wild-type mice show significantly is also implicated in the regulation of (Gulinello et al., better performance in a water maze model of spatial learning 2001). A concentration-dependent decrease of the α4 subunit is (Collinson et al., 2002). In addition, blockade of the GABAA- seen after 4-day application of allopregnanolone to developing receptor subunit α5 increased learning and memory (Casula et al., neuronal cells (Grobin and Morrow, 2000). In the hippocampus 2001; Maubach,2003). GABAA-receptor activation can inhibit LTP and cerebellum, an increase of the α4 subunit can be detected after induction and NMDA receptors, which are involved in the regu- withdrawal from chronic exposure and after short-term treatment lation of hippocampal-dependent spatial memory (Riedel et al., of progesterone and allopregnanolone (Smith et al., 1998; Concas 2003). et al., 1999; Follesa et al., 2001; Gulinello et al., 2001). The α6 sub- Several papers report changes in the GABAA-receptor sub- unit is highly sensitive to (Thompson et al., 1996) unit composition and decreased GABA function after long-term and neurosteroids (Belelli et al., 2002). exposure to GABAA-receptor (Miller et al., 1988; Belelli The γ2 subunit is also involved in anxiety regulation, and it et al., 2002). It is well-known that tolerance develops during long- is changed during hormone treatment and (Essrich term GABAA-receptor exposure. Tolerance development is noted et al., 1998; Follesa et al., 1998; Concas et al., 1999; Crestani already after 90 min exposure to anesthetic dosages of allopreg- et al., 1999; Kittler et al., 2000). The δ-subunit is responsible nanolone. Changes in the α4 subunit of the GABAA-receptor in for tonic conductance and important for neurosteroid modula- thalamus were related to the tolerance development (Birzniece tion on GABAA-receptor (Stell et al., 2003). Interestingly, receptor et al., 2006a). GABA-steroids are positive neurosteroid modula- knockout studies have revealed that the absence of the δ-subunit tors on the GABAA-receptor (Majewska et al., 1986). Long-term decreases the sensitivity to neurosteroids such as pregnanolone treatment with GABA-steroids may induce down-regulation of the and alphaxalone, thereby influencing the duration of GABAA-receptor function in mammalian cultured neurons both and the anxiolytic effect of those steroids (Mihalek et al., 1999). for neuroactive steroids and other GABAA-receptor active Finally, the ε-subunit reduces neurosteroid and anesthetic mod- (Yu and Ticku, 1995a; Yu et al., 1996). During pregnancy when ulation (Davies et al., 1997; Belelli et al., 2002; Thompson et al., allopregnanolone is high there is a decrease in GABAA-receptor 2002). function and changes in subunit composition of the GABAA- Functionally, distinct subunit-specific properties have been receptor. Inhibiting the synthesis of allopregnanolone blocked identified in both recombinant and native receptors, supporting these changes (Concas et al., 1998). Chronically high levels

Table 2 | Relationship between behavior disorders/symptoms, brain areas, and subunits of the GABAA-receptor.

Behavior disorders/symptoms Brain region GABAA-receptor subunit

Memory and learning disruption Hippocampus α5 δ

Anxiety and mood disturbance Hippocampus; amygdale α2 α4 β3 δ

Fatigue, , and exhaustion Wide spread α1

Depression induced by stress α1 α3 β1 β2 δγ2

Eating disorders Hypothalamus α3

Relapse in abuse β6

Balance and mobility disorders Cerebellum α6 β2 δ

Epilepsy and excitability disorders α1 β3

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 4 Wang Neurosteroids and GABAA-receptor function

and GABA-steroids give irreversible cognitive damages. A reduced located opposite to the release site (Mody et al., 1994; Nusser et al., sensitivity to benzodiazepines, alcohol, and GABA-steroid is also 1997; Brickley et al., 1999). seen in women with PMDD during the . Such a change Synaptically released GABA which act on postsynaptic GABAA- in GABAA-receptor sensitivity, as measured by reduced sedation receptors is termed “phasic” inhibition, whereas the term “tonic” and saccadic eye velocity response to GABA active compounds, inhibition refers to a continuous activation of extrasynaptic recep- contributes to symptom severity of PMDD patients (Backstrom tors by ambient GABA (Farrant and Nusser, 2005). The main et al., 2003). In a rat model of PMDD, allopregnanolone upreg- feature of phasic inhibition is the rapid synchronous opening of a ulates the α4 subunit of the GABAA-receptor in hippocampus relatively small number of channels that are clustered within the parallel to the induction of anxiety (Smith et al., 1998; Gulinello synaptic cleft. This enables a resolution both in time (the release et al., 2001). In addition, the anxiety induction was blocked if is triggered by an incoming ) as well as in space the animals were treated with α4-antisense (Smith et al., 1998; (the release is limited and thereby the postsynaptic action, to a Gulinello et al., 2001). With a further developed rat model of specific . In contrast, tonic inhibition results from ambi- PMDD, the authors addressed that the high risk-taking rats are ent GABA and therefore is relatively constant in both time and those react with anxiety in the PMDD model (Lofgren et al., 2006). space. Although both modes of action clearly impact upon neu- ronal information processing at the cellular and network level, the SYNAPTIC AND EXTRASYNAPTIC GABAA-RECEPTORS extent to which each type of inhibition influences brain excitabil- The GABAA-receptors can either be synaptic (located within ity in normal and diseased states is not known. GABA-mediated the synaptic cleft) or extrasynaptic (located outside the synaptic tonic conductance is found in granule cells of the dentate gyrus cleft; Fortin et al., 2004). The synaptic receptors usually contain (Nusser and Mody, 2002; Farrant and Nusser, 2005), thalamocor- γ-subunits and can be rapidly expressed at the neuronal mem- tical relay neurons of the ventral basal complex (Porcello et al., brane and are sensitive to both benzodiazepines and neurosteroids. 2003), layer V pyramidal neurons in the somatosensory cortex Extrasynaptic GABAA-receptors are in a preferred position to be (Yamada et al., 2004), CA1 pyramidal cells (Bai et al., 2001), and activated by the low levels of ambient GABA, due to their high certain inhibitory interneurons in the CA1 region of the hip- GABA affinity in contrast to the lower affinity of synaptic GABAA - pocampus (Semyanov et al., 2003). Unlike receptors mediating receptors (Saxena and Macdonald, 1994; Mody, 2001; Brown et al., phasic current, tonically active GABAA-receptors show unusual 2002; Farrant and Nusser, 2005). The high-affinity extrasynaptic high GABA affinity (Saxena and Macdonald, 1996) and be acti- GABAA-receptors consist of specific subunit combinations dif- vated by the low ambient GABA concentrations (nanomolar to ferentially expressed in various brain regions. These include the few micromolar; Lerma et al., 1986; Tossman et al., 1986). At δ-subunit containing GABAA-receptors of dentate gyrus and cere- sub-micromolar concentrations, several competitive, and non- bellar granule cells, cortical and thala mic neurons (Nusser et al., competitive GABAA-receptor antagonists (, picrotoxin, 1998; Sur et al., 1999; Pirker et al., 2000; Nusser and Mody, 2002; and ) reduced phasic currents, but had no effect on Stell et al., 2003; Sun et al., 2004; Cope et al., 2005; Jia et al., tonic currents. However, the antagonists blocked both phasic and 2005; Drasbek and Jensen, 2006), and the α5-subunit containing tonic currents at high concentrations (Bai et al., 2001; Stell and GABAA-receptors in CA1 and CA3 pyramidal cells (Sperk et al., Mody, 2002; Semyanov et al., 2003; Yeung et al., 2003; Farrant and 1997; Caraiscos et al., 2004; Glykys and Mody, 2006). The cur- Nusser, 2005). rent mediated by these extrasynaptic receptors has been termed tonic inhibition (Brickley et al., 1996; Farrant and Nusser, 2005), THE EFFECT OF NEUROSTEROIDS ON THE GABAA-RECEPTOR which is highly sensitive to the extracellular GABA concentra- The GABAA-receptor can be modulated by a number of therapeu- tion (GABA). It is enhanced when ambient (GABA) is increased tic agents, including benzodiazepines (Sieghart, 1992; Macdonald by blocking GABA transporters, by adding GABA to the aCSF and Olsen, 1994), barbiturates (Smith and Riskin, 1991), anes- to mimic that normally present in the extracellular space, or by thetics, (Harris et al., 1995), (Smart, 1992), and preventing GABA degradation (Nusser and Mody, 2002; Stell and neurosteroids (Puia et al., 1990; Hawkinson et al., 1994a; Lam- Mody, 2002; Wu et al., 2003; Glykys and Mody, 2006). bert et al., 2001b). The effect of neurosteroids on the GABAA- The δ-subunit is highly sensitive to GABA and neurosteroids receptor depends on the type of steroids (agonist or antagonist), but not to benzodiazepines (Brown et al., 2002; Smith et al., 2007). the type of receptors (synaptic of extrasynaptic), the subunit It also appears that specific types of receptor subunit combina- compositions, and the intrinsic structure of the steroid. Recent tions are expressed in the synaptic cleft and extra synaptically. For studies indicate that the existence of at least two neurosteroid example, the α1β2/3γ2, α2β2/3γ2, and α3β2/3γ2 receptors are the actions on the GABAA-receptor, namely an agonistic action and predominant synaptic receptors (Farrant and Nusser, 2005) and an antagonistic action by the sulfated and 3β-OH steroids. The the combinations of α6βnδ, α4βnδ, α1βnδ, and α5βnγ2 receptors, agonistic action can further be divided into an allosteric enhance- are predominantly or exclusively extrasynaptic (Nusser et al., 1998; ment of GABA-evoked Cl− current and a direct activation of the Fritschy and Brunig,2003; Glykys and Mody,2007). In the synapse, GABAA-receptor. each vesicle is thought to release several thousands of GABA mol- The synaptic currents originates from a short exposure of ecules into the synaptic cleft which leads to a high concentration high concentrations of GABA that induces receptor desensitiza- of GABA (0.3–1.0 mM) in a time span of 10–100 ms (Mozrzymas tion (Jones and Westbrook, 1995). As a consequence of this, the et al., 2003; Semyanov et al., 2004). Furthermore, it is suggested mechanism of neurosteroid modulation at the GABAA-receptor that there are rather few receptors, from 10 to a few hundreds, cannot easily be predicted by studies of currents evoked by low

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 5 Wang Neurosteroids and GABAA-receptor function

concentrations of GABA. It has been shown that allopregnanolone sensitivity compare to γ-subunit containing receptor (Davies et al., remarkably increases the decay time of sIPSC (Haage and Johans- 1997; Belelli et al., 2002). Receptors incorporating the ε-subunit son, 1999; Belelli and Lambert, 2005), which likely depends on a are reported to be insensitive to the modulation by pregnane reduced GABA unbinding rate from the receptor. However, a dif- steroids, not the direct GABA-mimetic effect (Lambert et al., ferent mechanism was suggested for the 3α5α-THDOC induced 2001b). decay time prolongation. Here it was instead suggested that the underlying mechanism was altered kinetics of desensitized states SYNAPTIC EFFECT OF NEUROSTEROIDS (Zhu and Vicini, 1997). The above explanations for the mecha- The brief inhibitory response of neurosteroids by activating nisms underlying the effects of neurosteroids is based on the idea the postsynaptic GABAA-receptor is a phasic response. Synaptic that once GABA is bound to the receptor it can enter different GABAA-receptors are ternary complexes that commonly incorpo- states of which some are open and some are closed (desensitized) rate the γ2 subunit in combination with one α (mainly α1/2/3) and which will determine the shape of the evoked current (Jones and one β2/3 subunit. However, these receptor isoforms can also be Westbrook, 1995). located extrasynaptically (Farrant and Nusser, 2005). The kinetic At low-micromolar concentrations of GABA,allopregnanolone of agonist steroids at synaptic GABAA-receptor has been stud- instead increases the activation rate and induces a more prominent ied thoroughly by measuring the sIPSC from neurons in brain desensitization of GABA-evoked currents (Haage and Johans- slice. Neurosteroids have little effect on the onset time and peak son, 1999). Interestingly, there is also evidence that the effect amplitude of the sIPSC. Agonist neurosteroids prolong the decay of 3α5α-THDOC to prolong the deactivation time (which is the time constant of IPSC (Majewska et al., 1986; Zhu and Vicini, same as the decay time but is used for externally applied GABA) 1997; Haage et al., 2005). However, this effect is neuron specific. In depends on the age. The effect on the deactivation of GABA hippocampal CA1 neurons, cerebellar granule cells and Purkinje response in cerebellar neurons is greater in younger rats than neurons, neurosteroids prolong the sIPSC at relatively low concen- adult rats (Zhu and Vicini, 1997). In summary, we believe that tration (in the nanomolar range; Cooper et al., 1999; Vicini et al., the prolongation of deactivation and altered receptor kinetics 2002; Harney et al.,2003). On the other hand, micromolar concen- in terms of entry and exit from desensitized states, are essential trations are required to produce equivalent responses in to the allosteric modulation of the GABAA-receptor by neuros- neurons of hypothalamus (Brussaard et al., 1997; Koksma et al., teroids. Agonistic neurosteroids affect not only the time course 2003). Moreover,in the preoptic cells in the hypothalamus,100 nM of sIPSC but also the frequency of the sIPSCs. Such an increase allopregnanolone prolong the spontaneous current (Haage et al., in frequency is due to the presynaptic effect of neurosteroids 2005; Stromberg et al., 2006). This indicates that the neurons in (Poisbeau et al., 1997; Haage et al., 2002) by a mechanism that the same brain region can show heterogeneity. In addition, the involves altered presynaptic Ca2+ permeability and activation of effect of 3α5α-THDOC on GABA-binding kinetic is more pro- presynaptic GABAA-receptors. found in the hippocampal CA3 and subiculum than that in CA1 The effect of neurosteroids on the GABAA-receptor can be and entorhinal cortex (Nguyen et al., 1995). At higher concen- attributable to variations in the receptor subunit composition. trations (>10 μM) which can occur in the brain during partu- The action of allopregnanolone was not influenced by the α- rition (Stoffel-Wagner, 2003), neurosteroids activate the GABAA- subunit, when co-expressed with β1- and γ2-subunit in Xenopus receptor directly (Majewska et al., 1986) in a similar pattern as oocytes (Belelli et al., 2006). The GABA responses at α1β1γ2 and barbiturates by interacting with different sites on GABAA-receptor α3β1γ2 receptor are enhanced by low concentration of allopreg- (Kerr and Ong, 1992). This “GABA-mimetic” effect of neuros- nanolone (≥3 nM), whereas several folds higher concentrations teroid is sufficient to suppress the excitatory neurotransmission are needed to obtain the equivalent response on α2-,α4-, α5-, (Shu et al., 2004). or α6-subunit containing receptors. Likewise, the subtypes of the β-subunit (β1–3) have little action on the effects of neu- EXTRASYNAPTIC EFFECT OF NEUROSTEROIDS rosteroids (Hadingham et al., 1993; Sanna et al., 1997; Belelli The response of neurosteroids at relatively low concentrations et al., 2002). The presence of a γ-subunit is not a prerequisite is mediated by the activation of extrasynaptic GABAA-receptors for the neurosteroid activity. In fact, the efficacy of allopreg- containing α4, α6, and δ-subunits. Extrasynaptic receptors iden- nanolone action at the binary α1β1 receptor is higher than that tified at the granule cells of the dentate gyrus and cerebellum, at the ternary α1β1γ2 receptors (Maitra and Reynolds, 1999; and the relay neurons of the thalamus, are distinct from the Belelli et al., 2002). Given the γ-subunit have little or no effect synaptic receptors. Extrasynaptic conductance can have a con- on the maximal GABA-modulation effect of allopregnanolone, it siderable influence on neuronal excitability (Leroy et al., 2004). significantly influences the potency of the steroid with “physio- Extrasynaptic receptors exhibit both a high GABA affinity and logical concentrations” (3–30 nM; Belelli et al., 2002). However, reduced receptor desensitization in the continued presence of the inhibition of GABAA-receptor by PS did not vary between the agonist (Fritschy and Brunig, 2003). Such properties ren- binary and ternary receptor (Wang et al., 2006). The poten- der these receptors ideally suited to sense the low ambient con- cies and efficacies of PS to inhibit GABA saturating concen- centrations (∼0.5–1 μM) of the extrasynaptic tration at the α1β2γ2L and α1β2 receptor were identical (Wang (Kennedy et al., 2002). Extrasynaptic receptors containing the et al., 2006). On the other hand, δ-subunit when co-expressed δ-subunit are highly sensitive to neurosteroids in certain brain with α4- and β3-subunits, a receptor thought to be naturally region (Wohlfarth et al., 2002). At low “physiological” concentra- present in the thalamus (Sur et al., 1999) shows high steroid tions (10–100 nM), 3α5α-THDOC selectively enhance the tonic

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 6 Wang Neurosteroids and GABAA-receptor function

conductance, with little or no effect on the phasic conductance in mouse DGCs and CGCs (Stell et al., 2003; Belelli and Lam- bert, 2005; Farrant and Nusser, 2005). Tonic inhibition is reduced in the δ-subunit “knockout” mice, and the residual tonic cur- rent was insensitive to 3α5α-THDOC (Mihalek et al., 1999; Stell et al., 2003). However, the extrasynaptic effects of neurosteroids also reveal regional difference in the CNS. Two hundred fifty nanomolar 3α5α-THDOC has no effect on the tonic inhibition in ventrobasalis complex of the thalamus (Porcello et al., 2003). On the other hand, the tonic conductance of hippocampal CA1 neurons expressing GABAA-receptor with α5 subunit is affected by 3α5α-THDOC (≥100 nM; Stell et al., 2003; Belelli and Lam- bert, 2005; Farrant and Nusser, 2005). It is suggested that the FIGURE2|Featuresofgeneral structure of the neurosteroid. Figure shows intact four ring system that often contains hydroxyl side chains as modulation of tonic currents is influenced by local neurosteroid sterols. Hydroxyl groups are denoted α if they are oriented above the plane . The inhibition of metabolism can greatly enhance (solid line), and α if they are oriented below the plane (dashed line). the response of the tonic current in dentate granule cells to Stereoisomerism and structural modifications were discussed at carbon 3, endogenous neurosteroids, but not on the synthetic metaboli- 5, 6, 7, 17, 20, and 21 positions. cally stable ganaxalone (Belelli and Herd, 2003). Neurosteroid metabolism reveals regional specificity that also contributes to 7-methanoindene derivatives give a compound that showed weak the regional specificity of the tonic GABAA-receptor mediated potentiating activity (Burden et al., 2000). When the carbonitrile currents. In summary, evidence is emerging that the GABAA- receptor mediated tonic conductance present in some neurons is replaced with an acetyl group the derivatives become inactive may have a considerable influence on neuronal signaling and (Hamilton, 2002). network activity (Brickley et al., 2001; Hamann et al., 2002; The configuration of C5-reduction is important to potency. α β Mitchell and Silver, 2003). The high sensitivity of neurosteroids Even if steroids with either 5 or 5 conformations are active, spa- in extrasynaptic receptor may represent an important target for tial difference in this position may affect the of the α neurosteroids. neurosteroids. Allopregnanolone with 5 -reduction is generally more potent than its 5β-isomer,pregnanolone,as GABAA-receptor α α AGONIST NEUROSTEROIDS ON THE GABA -RECEPTOR agonist both in vivo and in vitro. Studies with 3 5 -THDOC and A α β The structure–activity relationship of neurosteroid actions on the its stereoisomer 3 5 -THDOC revealed important differences in GABA -receptor has been summarized in a number of articles potency, efficacy, and regional selectivity at the GABAA-receptor in A α (Laubach et al., 1955; Gyermek et al., 1968; Gyermek and Soyka, favor of 5 -reduction (Gee and Lan, 1991; Mennerick et al., 2004). α β 1975; Sear, 1997). The systemic investigation of several isomers 5 -steroid, but not 5 -steroids, showed a high degree of enantios- electivity/enantiospecificity in their action as modulator of the of GABAA-receptor active neurosteroids revealed a couple crucial structure variations (Figure 2), namely the geometry between ring GABAA-receptor and as (Covey et al., 2000). The effi- β β A/B; a hydrogen-bond donator in C3 position; a hydrogen-bond cacy of 3 5 -P to inhibit GABAA-receptor is significantly different β α β acceptor in C20 position and/or a flexible bond at C17 position from 3 5 -P (Rahman et al., 2006; Wang et al., 2007). 5 -reduced (Purdy et al., 1990; Zorumski et al., 2000; Ragagnin et al., 2007). antagonist neurosteroids cause significant higher inhibition than α The α- and β-configuration refer to constituents below and above the 5 -isomers. the plane of the steroid backbone. Variation of the core and sub- Like pregnan steroids, metabolites of testosterone such as stitution in certain positions of the steroid molecules has crucial are also modulators of GABA. It was found that α consequence on the neurosteroid effects (Figure 2). the systemic 3 -androstanediol administration conditions a place Anesthetic steroids typically have a saturated backbone of four preference more effectively than does systemic administration rings, although this is not an absolute requirement for activity. of or testosterone (Frye, 2001). Additionally, α The four rings form a rigid framework for positioning the hydro- plasma concentration of 3 -androstanediol is increased compare gen bonding groups in three-dimensional spaces (Figure 2). The to dihydrotestosterone and testosterone. However, the potency α presence of hydrogen-bond donor in the α-configuration at C3 and efficacy effect of 3 -androstane steroid is lower than those α and β-configuration at C17 are critical for the sedative action of of 3 -pregnan steroids (Rahman et al., 2006). agonist neurosteroids (Purdy et al., 1990; Hawkinson et al., 1994b; Hogenkamp et al., 1997). These groups are important for the bind- ANTAGONIST NEUROSTEROIDS ON THE GABAA-RECEPTOR ing of neurosteroids to a variety of proteins by means of hydrogen- Neurosteroids may both enhance and inhibit GABAergic neuro- binding with polar or charged residues (Brzozowski et al., 1997; transmission (Wang et al., 2002; Mennerick et al., 2004; Rahman Grishkovskaya et al., 2000). Replacing the hydrogen of hydroxyl et al., 2006; Stromberg et al., 2006). It has been shown earlier that with methyl, thus eliminating the ability of the steroid to donate a 3β-hydroxy A-ring reduced pregnane steroids (3β-OH steroids) hydrogen bond in this region, results in dramatic reduction in and pregnenolone sulfate inhibit GABAA-receptor coupled anion its potency (Upasani et al., 1997). Replacing the steroid skele- channels (Wang et al., 2002, 2007; Lundgren et al., 2003; Birzniece ton with an alicyclic framework like 2-cyclohexylideneperhydro-4, et al., 2006b). 3β-OH steroids and PS, at concentrations that had

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 7 Wang Neurosteroids and GABAA-receptor function

little effect on GABAergic synaptic currents, significantly reversed leave open the possibility that PS may prefer liganded over unli- the potentiating effect of 3α-OH A-ring reduced steroids (Wang ganded receptors, consistent with a model of state-dependence to et al., 2002). Although antagonistic steroids reduced the potentia- PS actions. tion induced by 3α-OH steroids, they still acted non-competitively An earlier study observed that PS inhibits GABA-gated Cl− cur- with respect to the agonistic steroids and inhibited a larger poten- rent by enhancing receptor desensitization and stabilizing desen- tiation more efficiently (Wang et al., 2002). Furthermore, 3β-OH sitized state with prolonged application of low-affinity GABA steroids co-applied with GABA alone significantly inhibited GABA agonists to nucleated membrane patched (Shen et al., 2000). Note responses at concentrations ≥EC50 (Garrett and Gan,1998; Maitra that the promotion of desensitization is selective for prolonged and Reynolds, 1998; Wang et al., 2002; Lundgren et al., 2003). This GABA applications (Shen et al., 2000). It is well-known that the direct, non-competitive effect of 3β-OH steroids and PS on the time course of GABA-mediated IPSCs is influenced strongly by GABA response was sufficient to account for the apparent antago- the kinetics of the GABAA-receptor desensitization (Celentano nism of agonist steroids (Rahman et al., 2006). In summary, both and Wong, 1994; Jones and Westbrook, 1995). Desensitized states PS and 3β-OH steroids inhibited the GABAA-receptor more effec- are thought to buffer receptors in bound conformations that tively under conditions that promoted agonist binding or channel make it possible for channels to re-open before GABA unbinds. opening (Wang et al., 2002). The interaction between antagonist The fast phase of desensitization limits the open probability of and agonist steroids was due to a use-dependent action of antago- the channels, influences peak synaptic currents, and contributes nist steroids at recombinant and synaptic GABAA-receptors (Wang to the fast component of IPSC decay. The slow component of et al., 2002). decay may result from reopening of GABA channels after exit from desensitized states (Jones and Westbrook, 1995). On the MECHANISM OF PREGNENOLONE SULFATE AND NEUROSTEROID other hand, recombinant GABAA-receptors composed of defined SULFATE ESTERS subunit combinations also give rise to currents with complex Sulfated steroids like pregnenolone sulfate and dehydroepiandros- decay kinetics (Lavoie et al., 1997; Haas and Macdonald, 1999). terone sulfate can produce profound effects on behavior. PS as In cultured hippocampal neurons, PS decreases the peak current an abundant neurosteroid enhances learning (Mayo et al., 1993; of the inhibitory autapic currents, enhances the fast and slow Flood et al., 1995), and antagonizes the impairment of learning phases of deactivation after brief application of a few millisec- and memory produced by ethanol and scopolamine (Melchior and onds. After longer application (∼100 ms or more), PS enhances Ritzmann, 1996). PS may play a role in cognition and have been both the fast and slow phases of desensitization of the GABA reported as negative modulators of the GABAA-receptor based on current (Shen et al., 2000; Eisenman et al., 2003). Moreover, electrophysiological studies and GABA-mediated 36Cl− uptake by PS reduces the agonist steroid evoked prolongation of sIPSC rat brain synaptosomes (Majewska et al., 1990; Demirgoren et al., and reduces channel-opening frequency (Mienville and Vicini, 1991). Structure–activity relationships of GABAA-receptor modu- 1989). However, recording the sIPSC of neurons medial preoptic lation are different for sulfated inhibitory steroids . non-sulfated nucleus (MPN) shows that PS neither affects the peak ampli- potentiating steroids (Park-Chung et al., 1999). Potentiation by tude nor the decay of sIPSC (Haage et al., 2005). Instead, PS non-sulfated steroids requires 3α-stereochemistry. Pregnenolone reduces the frequency of the sIPSC at higher concentration (Haage by itself is inactive on the GABAA-receptor. In contrast, both 3α- et al., 2005). The main effect of PS on the sIPSC time course is and 3β-isomers of PS is inhibitory. Although the addition of a explained by a simplified model that this substance reduce the rate negatively charged sulfate group or hemisuccinate group at the of desensitization while the agonist steroids assumed to reduce C3 position converts the neurosteroid from being potentiating to the unbinding rate of GABA from the receptor (Haage et al., inhibitory (Park-Chung et al., 1999), a negatively charged group 2005). at C3 is not absolutely essential for inhibition since the non- sulfated neurosteroid DHEA is also inhibitory. However, DHEA MECHANISM OF 3β-HYDROXYSTEROID AS GABAA-RECEPTOR is less potent than its sulfated derivative DHEAS (Imamura and ANTAGONIST Prasad, 1998; Park-Chung et al., 1999). Steroids such as 11-keto According to earlier studies, PS (Woodward et al., 1992), 3β- derivative of PS are of particular interest, as it behaves as a posi- OH steroids, or carboxylated steroids (Mennerick et al., 2001) tive, negative, or neutral modulator on the GABAA-receptor. It is are more effective against GABA responses gated by high con- suggested that 11-keto pregnenolone sulfate (11-keto PS) exerts centrations of GABA. In electrophysiological studies, 3β5β-P a dual action on distinct positive and negative steroid modula- antagonized the 3α5β-P induced enhancement of GABA cur- tion sites associated with the GABAA-receptor (Park-Chung et al., rent (Maione et al., 1992; Maitra and Reynolds, 1998). On the 1999). other hand, 3β5α-P diminished the inhibitory effects of 3α5α- Consistent with the idea that PS induces activation-dependent P on population spikes evoked in rat hippocampal CA1 stratum inhibition on the GABAA-receptor (Wang et al., 2002), another pyramidal (Wang et al., 2000). Studies on the uptake report revealed that the inhibition of basal inhibitory postsynap- into synaptosomes in rat cerebral cortex, in hippocampus and tic currents (IPSCs) by antagonist steroids was correlated with sIPSC in MPN showed that 3β-OH steroids reduced the 3α5α-P the basal decay time of the IPSCs (Eisenman et al., 2003). Analy- enhanced GABA response (Stromberg et al., 2006). In the pres- sis of single-channel behavior in the presence of GABA and PS ence of low concentration GABA, some of the 3β-OH steroids suggested no difference in the ability of PS to block liganded potentiated GABA-evoked chloride uptake and prolonged closed vs. liganded open receptors (Akk et al., 2001). These results the decay time, whereas the others had little or no effect on

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 8 Wang Neurosteroids and GABAA-receptor function

GABA stimulated current. Therefore, certain 3β-OH steroids, Selective antagonists of ρ receptors can prevent the development namely 5β-pregnane-3β, 20(S)-diol and 3β-OH-5β-pregnane-20- of myopia and enhance learning and memory in rats in the Mor- one have both agonistic and antagonist property (Wanget al.,2002; ris water maze task after intraperitoneal injection (Chebib et al., Stromberg et al., 2006). Moreover, another study with the Cl− 2009). Receptors containing the ρ subunit have been implicated in uptake method has shown that 3β-OH-5α-pregnan-20-one is a apoptosis of hippocampal neurons (Yang et al., 2003) and regula- useful antagonist of 3α5α-P enhanced GABA response (Lundgren tion of hormone release in the pituitary gland (Boue-Grabot et al., et al., 2003). However, in GABAA-receptor expressed in Xeno- 2000). pus oocytes showed that 3β-steroids inhibit GABA response at Several neurosteroids are shown to modulate the ρ1 recep- near-saturating concentrations (Rahman et al., 2006). It is still tor channel. Allopregnanolone, 5α-THDOC, and 5α-pregnane- unclear why the effect of 3β-OH steroids varies with different 3α-ol-11,20-dione (alphaxalone) potentiated the GABA-evoked methods. It is likely that the steroid structure, drug application currents from ρ1 receptor channels and concomitantly time and base line GABA concentration may be responsible for altered the deactivation kinetics by prolonging the decay the discrepancy. time. In contrast, pregnanolone, 5β-pregnane-3,20-dione (5β- It is suggested that a homologous mutation of the residue at dihydroprogesterone), and 5β-THDOC inhibited the GABA-  2 position closest to the cytoplasmic end of the M2 helix to elicited currents of the ρ1 receptor channel. In comparison to on both the β1 and the β2 subunit, α1V256S and β2A252S, reduced GABAA-receptors, the modulation of ρ1 receptor channels by the desensitization rate of GABA-activation at saturating doses neurosteroids occurred with relatively high concentrations and (Wang et al., 2007). In a receptor complex with reduced desensi- was more prominent in the presence of low concentrations of tization components to GABA-activation (e.g., mutant receptors), GABA. Structural comparison of these six neuroactive steroids the PS-inhibition was also greatly reduced (Akk et al., 2001). reveals that the key parameter in determining the mode of On the other hand, PS has been shown earlier to increase the modulation for the ρ1 receptor channel is the position of the deactivation rate of the GABA-evoked sIPSCs recorded by patch- hydrogen atom bound to the fifth carbon, imposing a trans- clamp techniques at the acutely dissociated neurons from the or cis-configuration in the backbone structure (Morris et al., MPN area of the rat brain slice (Haage et al., 2005). Recent 1999). studies confirmed the findings that PS increased the fast offset A study focusing on the electrophysiological effects of rate of GABA-activation (Wang et al., 2007). However, the slow inhibitory steroids on the ρ1 receptor found that steroid inhibitors component of the offset time course was decreased by PS in a could be divided into three major groups based on how mutations dose-dependent manner. The potencies of 5α-pregnan-3β,20α- to residues in the M2 transmembrane domain modified inhibition diol and 5β-pregnan-3β,20β-diol to influence offset time courses (Li et al., 2007). A recent study from the same research group (Li of GABA-activation were significantly lower than PS, in accor- et al., 2010) selected representatives of the three groups (preg- dance with our earlier findings while PS was significantly more nanolone, tetrahydrodeoxycorticosterone, pregnanolone sulfate, potent to inhibit both peak and steady-state GABA currents than allopregnanolone sulfate and β-) to probe how these 3β-OH steroids (Wang et al., 2002, 2006, 2007). PS-inhibition was steroids, as well as the non-steroidal inhibitor picrotoxinin, mod- already seen at low dose of GABA response (≤EC20; Eisenman ify GABA-elicited fluorescence changes from the Alexa 546 C5 et al., 2003), whereas the inhibition by 3β-OH steroids on the cur- maleimide fluorophore attached to residues in the extracellular rent response was first seen at higher end (>EC50) of the GABA region of the GABAA-receptor. The fluorophore responds with dose-response curve (Wang et al., 2002). Obviously, 5α-pregnan- changes in quantum yield to changes in the environment, allow- 3β,20α-diol and 5β-pregnan-3β,20β-diol prolonged the fast offset ing it to probe for structural changes taking place during channel time course of GABA response,suggesting that the inherent associ- activation or modulation. The authors reported that the modula- ation between 3β-OH steroids and receptor has rather high affinity. tors have specific effects on fluorescence changes suggesting that However, it was not fully excluded that a more complicated model distinct conformational changes accompany inhibition (Li et al., of a multivalent interaction between hydrophobic steroids and 2010). Results are consistent with the steroids acting as allosteric receptor is employed. Kinetic properties of 3β-OH steroids have inhibitors of the ρ1 GABA receptor and support the hypothe- also been elucidated in acute dissociate neurons from the MPN sis that divergent mechanisms underlie the action of inhibitory area of the hypothalamus (Stromberg et al., 2006). These steroids steroids on the ρ1 GABA receptor. have no effect on the activation phase or the maximum amplitude of 3α5α-P enhanced IPSCs. They rather affect the slow deactiva- THE SITE OF NEUROSTEROID AGONIST ON THE tion phase. A recent report shows that 3β-OH steroids exert their GABAA-RECEPTOR effect by reducing the 3α5α-P induced prolongation of decay time Neurosteroids bind to GABAA-receptors at a site that is distinct constant (tdecay; Stromberg et al., 2006). from the recognition sites for GABA, benzodiazepines, and bar- biturates. This results in allosteric modulation of GABA bind- DIFFERENTIAL MODULATION OF ρ RECEPTOR CHANNELS BY ing or channel gating. The ability of neurosteroids to potentiate NEUROSTEROIDS GABA-activated currents recorded from a variety of neurons indi- The ρ1 GABA receptor is constitutes a dominant inhibitory force cated that most isoforms of the GABAA-receptor should be capa- in the retina but is expressed at lower levels throughout the ner- ble of binding neurosteroid modulators (Mitchell et al., 2008). vous system. Drugs acting on the ρ1 receptor can be useful in This concept was supported by studies of recombinant GABAA- the treatment of a variety of visual, sleep, and cognitive disorders. receptors with differing subunit composition that revealed a wide

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 9 Wang Neurosteroids and GABAA-receptor function

spectrum of activity for neurosteroids at GABAA-receptors with α-subunit. The importance of this site for neurosteroid binding some minor variations in efficacy (Puia et al., 1993; Lambert was also confirmed by the functional data on mutant α1 subunit et al., 1996; Maitra and Reynolds, 1999; Belelli et al., 2002). Neu- that reduced the potentiation of GABA responses by many syn- rosteroid modulation lacking subunit selectivity suggested that thetic steroids at α1β2γ2L receptors (Akk et al., 2008; Li et al., neurosteroids are binding to a site that is conserved throughout 2009). most members of the GABAA-receptor family. Studies using GABAA/-receptor chimeras suggested an THE SITE OF NEUROSTEROID ANTAGONIST ON THE allosteric action of neuroactive steroids at the N-terminal side GABAA-RECEPTOR of the middle of the second transmembrane domain (M2) of It is generally accepted that the sulfate moiety is critical in pro- the GABAA-receptor β1 and/or α2 subunits (Rick et al., 1998). ducing a steroid that blocks rather than potentiates GABAA- Electrophysiological studies have confirmed that neurosteroid receptors (Park-Chung et al., 1999). The fact that an anionic agonists enhance Cl−-currents by increasing both channel fre- group is critical gave suggestion to us that the sulfate might actu- quency and channel open duration at GABAA-receptor (Callachan ally interact with residues forming the binding site mediating et al., 1987; Puia et al., 1990; Zhu and Vicini, 1997). Neurosteroid antagonism. However, no voltage dependency was observed with agonists do not require direct aqueous access to the receptor, pregnenolone sulfate-inhibition on the GABA response (Majew- and membrane accumulation is required for receptor modula- ska and Schwartz, 1987; Majewska et al., 1988; Akk et al., 2001), tion (Akk et al., 2005). Hosie et al. (2006) identified two dis- which suggests that the charged sulfate moiety does not interact crete binding sites in the receptor’s transmembrane domains that significantly with the membrane field as PS approaches the transi- mediate the potentiating and direct activation effects of neuros- tion state between unbound and unblocked to bound and blocked teroid agonists at the GABAA-receptor. The potentiating effect (Woodhull, 1973). Therefore, the lack of voltage dependence of of 3α5α-THDOC is mediated by a cavity formed by the α- pregnenolone sulfate-inhibition suggests that it is unlikely that subunit transmembrane domains. On the other hand, the direct the sulfate moiety interacts with a site deep within the channel activation of GABAA-receptorby3α5α-THDOC is mediated by (Akk et al., 2001). Results from our earlier report suggest that interfacial residues between α and β-subunits, and is enhanced pregnenolone sulfate was a γ2-subunit independent inhibitor at by steroid binding to the potentiation site (Hosie et al., 2006). GABAA-receptors (Wang et al., 2006). However, residues deep  These profiles indicate that two distinct neuroactive steroid bind- in the channel at the 2 position in the M2 helix of both α1- ing sites may exist; αTHr236 and βTyr284 residues in the trans- and γ2-subunit were critical for pregnenolone sulfate-inhibition membrane domain initiate direct activation whereas αGln241 (Akk et al., 2001; Wang et al., 2002, 2006). Theoretically, a large, and αAsn407 mediate the potentiating response (Hosie et al., rigid molecule such as a steroid would bind end-on rather than 2006). The neurosteroid potentiation site was further identi- with its long axis across the channel. It is thus possible that fied in the α1β2γ2S receptor by mutation of Q241 to methio- either the A-ring (where the sulfate moiety is attached) or the nine or leucine, which reduced the potentiation of GABA cur- D-ring (the other end of the molecule) would penetrate the Cl− rents by the naturally occurring neurosteroids, allopregnanolone, channel most deeply. The lack of voltage dependence indicates or tetrahydrodeoxycorticosterone (THDOC; Hosie et al., 2006, that the A-ring is unlikely to penetrate to the 2 position. The 2007). structure of the D-ring of PS is identical to that for many 3β- In order to address whether the potentiation site for neu- hydroxysteroids. Since the inhibitory properties of 5α-pregnan- rosteroids on GABAA-receptors is conserved amongst different 3β,20α-diol and 5β-pregnan-3β,20β-diol were also reduced GABAA-receptor isoforms, Hosie et al. (2006) used chimera to in the α1Mβ2γ2L receptor, we think it is more likely that the generate 100 models based on an alignment of the transmem- uncharged portions of the steroid molecules interact with the 2 brane domains of the GABAA-receptor α1 subunit and the nACh residue closest to the cytoplasmic end of the M2 helix on the α1 receptor α chain and selecting the 10 best models, revealed that subunit. the potentiation site is associated only with the a subunit and the The site of action of sulfated neuroactive steroids on GABAA- activation site is interfacial between the α- and β-subunit. receptors remains unclear. Based upon the observation that PS 35 By using heterologous expression of GABAA-receptors in HEK reduced the apparent affinity of [ S]-TBPS, Sousa and Ticku cells, in combination with whole-cell patch-clamp recording (1997) suggested that PS and DHEAS might bind at the picro- methods, a relatively consistent potentiation by allopregnanolone /cage site. However, a mutation to the transmem- of GABA-activated currents was evident for receptors composed brane M2 channel domain eliminated picrotoxin sensitivity but of one a subunit isoform (α2–5) assembled with β3 and γ2S sub- the inhibitory effects of PS and DHEAS persisted (Shen et al., units (Hosie et al.,2009). By introducing mutant αβγ receptors,the 1999; Gibbs et al., 2006). The absence of voltage sensitivity or neurosteroid potentiation was dependent on the conserved gluta- alteration of single-channel open time argues against a binding mine residue in M1 of the respective a subunit (Hosie et al., 2009). site within the pore. Akk et al. (2001) identified a valine residue Studying wild-type and mutant receptors composed of α4β3δ sub- in the channel domain of the α1 subunit that slowed the devel- units revealed that the d subunit is unlikely to contribute to the opment of PS-inhibition when mutated to serine, but concluded neurosteroid potentiation binding site and probably affects the that this residue is unlikely to be part of the binding site and efficacy of potentiation (Hosie et al., 2009). likely influences PS action indirectly (Akk et al., 2001). Using In summary, the neurosteroid potentiation site was likely con- homologous mutation of the residue at 2 position closest to the tained entirely within or on the transmembrane domains of the cytoplasmic end of the M2 helix to serine on both α1 and β2

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 10 Wang Neurosteroids and GABAA-receptor function

subunit, namely α1V256S and β2A252S, it was found that effect the follicular phase and up to 4 nmol/L in the luteal phase. of PS is greatly reduced in these two mutant (Akk et al., 2001; In the third trimester of a pregnancy, serum concentrations of Wang et al., 2006, 2007). This suggests that these two amino allopregnanolone can increase to more than 100 nmol/L (Bicikova acids are involved in the PS mediated inhibition. However, it is et al., 1995; Luisi et al., 2000). Pregnanolone displays a similar pat- unclear whether these specific mutations at the cytoplasmic end tern with increase in the luteal phase (Wang et al.,1996; Sundstrom of M2 helix exert their effect by altering the allosteric mecha- et al., 1998). nism or by directly altering a binding site. A recent study in C. GABA-steroids are positive modulators of GABA responses elegans identified multiple residues in transmembrane domain 1 on the GABAA-receptor (Majewska et al., 1986). Chemically (M1), as well as a residue near the extracellular end of the M2 they are 3α-hydroxy-5α/β metabolites of progesterone (preg- helix, that are critical for low-micromolar inhibition of C. ele- nanolone and allopregnanolone), testosterone (3α 5α-androstan- gans GABAA-receptors by PS (Wardell et al., 2006). This latter diol),and desoxycorticosterone (tetrahydro-desoxycorticosterone, residue is of particular interest, as it is a positively charged argi- THDOC). GABA-steroids enhance the GABA effect on chloride nine that could potentially coordinate with the negatively charged flux resulting in increased inhibition of neural activity (Majewska sulfate of PS. The C. elegans receptor exhibits some pharmaco- et al., 1986; Gee et al., 1987; Birzniece et al., 2006b). GABA- logical differences as compared to mammalian GABAA-receptors steroids induce sedation and can be used as anesthetic drugs (for example, pregnanolone is inhibitory). So, these results may in human (Carl et al., 1990; Timby et al., 2006). GABA-steroids or may not be relevant to mammalian receptors; however, it is are anti-epileptic (Backstrom et al., 1984; Landgren et al., 1987) notable that an arginine residue is also found in this region of and in animal experiments they also possess an anxiolytic effect mammalian GABAA-receptor subunits (Gibbs et al., 2006; Wardell similar to benzodiazepines (Wieland et al., 1991). An anxiolytic et al., 2006). effect of allopregnanolone has, however, never been shown in Block of responses to high-efficacy agonists by this sulfated humans. steroid is greater than block of responses to partial agonists at sat- Human and animal studies revealed typical GABAA-receptor urating concentrations. This is called “activation dependant” or agonistic effects of allopregnanolone and pregnanolone at high “state dependant inhibition” (Wang et al., 2002; Eisenman et al., doses, i.e., sedation/anesthesia (Carl et al., 1990; Timby et al., 2003). Picrotoxin, another GABA superficially 2006), anti-epileptic effects (Landgren et al., 1998), and anxiolytic similar to pregnenolone sulfate in its activation dependence but effects in animals (Wieland et al., 1991). However, earlier reports the site of action of PS does not require a functional picrotoxin from human and animal models also indicated that all GABAA- site for inhibition of GABA responses (Shen et al., 1999). The receptor agonists could induce negative symptoms with anxiety, GABA Zn2+ also acts in activation dependant /aggressiveness in certain individuals. Strong irritabil- manner. However, the Zn2+ binding site is located in the interface ity/ was induced in 3–6% of individuals and moderate between the α- and β-subunit and Zn2+ activity is minimal in symptoms are induced in 20–30% (Masia et al., 2000; Weinbroum ternary αβγ receptor (Hosie et al., 2003). et al., 2001). Interestingly, the prevalence of PMDD among women in reproductive age was in the similar range of 3–8%. The corre- GABA-STEROIDS AND THE MOOD CHANGE sponding prevalence of PMS, with milder symptom severity than Mood disorders are common health problems affecting women, PMDD, was 25–35% of women in reproductive age (Sveindottir especially during fertile ages. It is suggested that periods of hor- and Backstrom, 2000). monal variability, i.e., menarche (Angold et al., 1999), premen- It is puzzling why an increase in allopregnanolone during strual periods (Soares et al., 2001), postpartum (Kendell et al., the is related to development of negative mood 1981; Chaudron et al., 2001), and perimenopause (Freeman et al., as allopregnanolone should be anxiolytic agent like benzodi- 2004) increase the risk of mood disorders in certain women. azepines. The answer depends on the fact that all GABAA- There are three obvious examples in interaction between mood, receptor agonists such as benzodiazepines, barbiturates, alco- steroids, and CNS, namely PMDD, side effects of oral contracep- hol, and allopregnanolone have paradoxical effects tives and negative mood symptoms encountered during sequential in certain individuals. At low concentrations or doses they progestagen addition to treatment in postmenopausal give severe adverse emotional reactions in a subset of indi- women. viduals (3–6%) and moderate reactions in up to 20–30% of An obvious relation between sex steroids and mood changes individuals. This paradoxical effect is induced by allopreg- are cyclic symptoms related to the menstrual cycle. Estradiol nanolone (Beauchamp et al., 2000; Miczek et al., 2003)ben- and progesterone display regular predictable changes during the zodiazepines (Ben-Porath and Taylor, 2002), barbiturates (Lee menstrual cycle. In parallel with the increase of progesterone, an et al., 1988; Kurthen et al., 1991; Weinbroum et al., 2001), and increase also occur in serum concentrations of allopregnanolone ethanol (Cherek et al., 1992; Dougherty et al., 1996; Miczek and pregnanolone (Wang et al., 1996). Although allopregnanolone et al., 2003). Symptoms induced by these GABAA-receptor active is synthesized in the brain, the major source to its brain con- drugs are depressive mood, irritability, aggression, and other centration is the corpus luteum of the ovary (Bixo et al., 1997; symptoms known to occur during the luteal phase in women Ottander et al., 2005). Allopregnanolone and progesterone are with PMS/PMDD. A biphasic effect was also observed for produced in parallel at the luteal phase of menstrual cycle (Wang medroxyprogesterone (MPA) and natural progesterone in post- et al., 1996; Genazzani et al., 1998). In fertile women plasma menopausal women taking hormone therapy. These women felt levels of allopregnanolone are approximately 0.2–0.5 nmol/L in worse on a lower dosage of MPA or progesterone than on a

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 11 Wang Neurosteroids and GABAA-receptor function

higher dosage or placebo (Bjorn et al., 2002; Andreen et al., 2005, reported that the change in effect coincide with an higher expres- 2006). sion of the α4β2δ subunit combination (Shen et al., 2007). The Thus allopregnanolone seems to have a biphasic effect on mood GABAA-receptor subunit combination that contains the δ-subunit with an inverted U-shaped relationship between concentration was known to be very sensitive to GABA-steroids, and is activated and effect. In postmenopausal women receiving progesterone, a by lower concentrations of allopregnanolone than other receptor biphasic relation between the negative mood symptoms and the subtypes (Wohlfarth et al., 2002; Liang et al., 2004). Therefore, plasma concentrations of allopregnanolone was observed. The the action of allopregnanolone at the α4β2δ receptor provides a negative mood increased with the elevating serum concentration mechanism for the generation of negative mood at . The of allopregnanolone up to the maximum concentration seen at the proposed hypothesis requires a higher number or proportion of luteal phase. With further increase in allopregnanolone concentra- GABAA-receptors with the type(s) that are inhibited by steroids tion there was a decrease in symptom severity (Andreen et al., also in other conditions when paradoxical steroid effects are seen 2005, 2006). An inverted U-shaped relation between allopreg- (PMDD/PMS). Actually, the α4βδ subunit composition was a key nanolone dosage and irritability/aggression has also been noted factor in the progesterone withdrawal model for PMDD (Gallo in rats (Miczek et al., 2003). and Smith, 1993; Smith et al., 1998). A third hypothesis is that the paradoxical effect arises when POSSIBLE EXPLANATIONS FOR THE PARADOXICAL EFFECT the Cl− gradient across the membrane of critical neurons favors OF GABA-STEROIDS an excitatory rather than inhibitory effect of GABAA-receptor In this section, possible mechanisms of the biphasic response activation. The Cl− gradient across the is subject to curve of allopregnanolone on behavioral parameters are discussed. active regulation and may also vary between different types of neu- The basic idea of so called paradoxical effect where neuros- rons as well as between different compartments within the neuron. teroids show one type of effect at low concentrations and another In adult vertebrate neurons, the intracellular Cl− concentration type at high concentrations is that an enhanced GABAA-receptor was usually relatively low and, the GABAA-receptor activation activity may give an excitatory net effect in certain situations, inhibited neuronal impulse activity. In contrast, during fetal and instead of the usual inhibitory effect (Backstrom et al., 2011). early postnatal development, the intracellular Cl− concentration The following hypotheses suggest several possible mechanisms was comparably high, and thus the GABAA-receptor activation how this can be achieved. In addition, there are no contradic- caused an excitation (Kahle et al., 2008). The intracellular Cl− tions between the different hypothesis and they may very well act concentration is determined not only by passive flux through the − in parallel. GABAA-receptor and other Cl -permeable channels, but also by One explanation is based on that inhibitory neurons are the cation-Cl−-cotransporters, with the major inward transport often most sensitive to GABA-steroids and therefore are the first to be mediated by NKCC1 and the major outward transport by KCC2 inhibited at elevated steroid concentrations. This will give a release (Price et al., 2005, 2009; De Koninck, 2007). In adults, the out- of the inhibitory tone at excitatory neurons, a so called disin- ward transport mediated by KCC2 dominates, which keeps the hibition. The idea emphasizes that specific combination of sub- intracellular Cl− concentration low. units determines the receptor sensitivity to neurosteroids (Belelli However, the intracellular Cl− concentration may be higher et al., 2002; Stromberg et al., 2006). In this context, extrasy- also in the adult brain, which has been demonstrated for the naptic GABAA-receptors containing the δ-subunit are of interest axon initial segment of cortical neurons (Szabadics et al., 2006; because they seem to be more sensitive to GABA-steroids than Khirug et al., 2008) and certain types of neurons in amygdala other types of subunit combinations (Mody, 2008). A possibil- (Martina et al., 2001) and substantia nigra (Gulacsi et al., 2003), as ity exists that one type of inhibitory neurons contains GABAA- well as in gonadotropin-releasing hormone neurons (Moenter and receptors with the δ-subunit, while the next in order contains DeFazio, 2005) and in presynaptic terminals (Haage et al., 2002). another subunit, making them less sensitive to GABA-steroids. Interestingly, estradiol was one factor that increased the activ- It has also been reported that the receptor subunit composition ity of NKCC1 under normal physiological conditions and thus and sensitivity to GABA-modulators can be regulated by environ- increased the intracellular Cl− concentration (Nakamura et al., mental factors such as stress (Biggio et al., 1990; Concas et al., 2004; Galanopoulou, 2008). Intriguingly, estradiol dose depen- 1996). dently increased the mood provoking effect of progestagens in Another idea is based on the finding that GABA-evoked cur- women (Bjorn et al., 2003), and negative mood effects in women rents at receptors with the α4β2δ combination are actually inhib- with PMDD/PMS (Dhar and Murphy, 1990). ited, instead of potentiated, by allopregnanolone (Shen et al., Finally, would it be possible to explain the inverted bell shaped 2007). During stress allopregnanolone and other GABA-steroids relationship between GABAA-receptor active steroid concentra- are produced and give an anxiolytic effect (Bitran et al., 1999). tions and symptoms that has been reported from clinical studies? However, Smith and colleagues showed that female mice reacted ThereportofPrescott et al. (2006) suggested that at a slightly with increased anxiety to stress during puberty (Shen et al., 2007). more positive Cl− equilibrium potential, a moderate increase in In humans, puberty is often a period characterized by mood GABAergic activity gave an increased excitation, whereas a larger swings and anxiety. This research group revealed that allopreg- GABAergic activity gave the opposite effect. However, whether an − nanolone changed from being a positive modulator of the GABAA- altered intracellular Cl concentration in adulthood may explain receptor, at the time before and after puberty, to become a negative the paradoxical effect of GABAA-receptor modulators remains to modulator at the time of puberty (Shen et al., 2007). They also be proven.

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 12 Wang Neurosteroids and GABAA-receptor function

THE ROLE OF NEUROSTEROIDS IN BEHAVIORAL DISORDERS magnocellular nucleus enhanced memory performance, whereas A number of review articles have discussed the important role allopregnanolone disrupted memory (Mayo et al., 1993). Preg- of neurosteroids in treating behavioral disorders by interacting nenolone, DHEA, and DHEAS increased memory when injected with the GABAA-receptor (Majewska, 1992; Reddy and Kulkarni, systemically, centrally or into the amygdala (Flood et al., 1988, 2000; Zorumski et al., 2000; Rupprecht et al., 2001). There are 1992; Wolkowitz et al., 1995). There was evidence that the con- a few obstacles preventing the clinical use of endogenous neu- centrations of DHEA and DHEAS decreased in patients suffering rosteroids. First of all, naturally occurring neurosteroids such as from AD (Sunderland et al., 1989; Nasman et al., 1991; Hillen et al., allopregnanolone have low because they are rapidly 2000). It is promising that certain neurosteroids should be further inactivated and eliminated by glucoronide or sulfate conjuga- explored in the context of prevention and treatment of Alzheimer’s tion at the 3α-hydroxy group. The second obstacle is that the and mild cognitive impairment. 3α-hydroxy group of allopregnanolone may undergo oxidation to a , restoring its activity at nuclear hormone receptors GABA-STEROIDS AND COGNITIVE FUNCTIONS (Rupprecht et al., 1993). Ganaxalone (3α-hydroxy-3β-methyl-5α- , stress-steroids, and sex steroids are linked to the pregnane-20-one), the 3β-methyl analog of allopregnanolone, is development of dementia (Shumaker et al., 2003; Lupien et al., an example of synthetic neurosteroid that overcomes these limi- 2005; Sandstrom et al., 2005). In women taking postmenopausal tations (Carter et al., 1997). Like allopregnanolone, ganaxalone is hormone therapy by administration of medroxyprogesterone, a a positive of GABAA-receptor (Carter et al., GABA-steroid precursor, the risk for dementia, and AD doubled 1997). after 5 years of treatment (Shumaker et al., 2003). Medroxyprog- Neurosteroids modulate anxiety and stress level. After an acute esterone can induce anesthesia by acting on the GABAA-receptor stress stimulus, release of progesterone, pregnenolone, allopreg- (Meyerson, 1967). On the other hand, estrogen has been claimed nanolone, and 3α5α-THDOC occur in the blood circulation to protect against dementia (Henderson et al., 1994). This was, (Purdy et al., 1992; Barbaccia et al., 1998; Serra et al., 2000). Allo- however, not confirmed in the large Women’s Health Initiative pregnanolone and 3α5α-THDOC have potent anxiolytic activity (WHI) study but estrogen by itself did not increase the risk to in several animal anxiety models (Crawley et al., 1986; Bitran develop dementia (Shumaker et al., 2004). Hormone therapy with et al., 1995; Wieland et al., 1995; Reddy and Kulkarni, 2000). medroxyprogesterone resembles the exposure to stress-steroids However, the anxiolytic effect of allopregnanolone has not been during chronic stress. GABAA-receptor agonists are known to shown in human (Wihlback et al., 2006). Recent reports indicated impair memory. Allopregnanolone can inhibit learning in rats that allopregnanolone can induce aggression and anxiety at low tested in Morris water maze, an accepted test model for learning concentrations (Fish et al., 2001; Gulinello et al., 2001; Miczek and memory (Johansson et al., 2002). It is well-known that most et al., 2003). Allopregnanolone and cortisol levels are increased GABAA-receptor agonists, e.g., benzodiazepines (Barker et al., during the examination of PhD students (Droogleever Fortuyn 2004), barbiturates (Mohammed et al., 1987), and alcohol (Saun- et al., 2004). Therefore, it is suggested that allopregnanolone has ders et al., 1991; Vincze et al., 2007) impair memory and learning biphasic effects in certain individuals (Miczek et al., 2003). At low and increase the risk for permanent damages, although the risk doses it has an adverse, anxiogenic effect. This effect decreases with with low and moderate alcohol consumption is under debate increasing doses of allopregnanolone and the beneficial and calm- (Solfrizzi et al., 2007). ing property occurs (Beauchamp et al., 2000; Fish et al., 2001). The brain and serum concentrations of GABA-steroids vary A major concern with potential new is whether they with the production of adrenals, ovaries, and testicles (Purdy suffer the same drawbacks as classical benzodiazepines. Using the et al., 1991; Wang et al., 1996; Bixo et al., 1997; Luisi et al., elevated plus-maze paradigm for assessing the anxiolytic activity, 2000). Interestingly, classical stress hormones, cortisol, and cor- selective effects of neurosteroids have been reported which differ ticosterone, metabolized in a similar way and allotetrahydrocor- from (Rodgers and Johnson, 1998). Synthetic deriv- tisol (Allo-THF, 3α-hydroxy-5α-cortisol) interacts also with the ative Co 2-6749 30, which retains a 3β-trifluoromethyl group GABAA-receptor (Celotti et al., 1992). Allo-THF enhances the that should block metabolism and enhance oral bioavailability, inhibitory effect of allopregnanolone and has synergic effect on is selected for clinical development because there is a large sep- neuronal inhibition together with allopregnanolone (Stromberg aration between anxiolytic effects and side effects (Gasior et al., et al., 2005). That is why neurons are exposed to a stronger 1999; Vanover et al., 2000). Women with PMDD show often dif- inhibition during acute and chronic stress. During acute stress, ficulties in concentration and develop fatigue during the luteal allopregnanolone and THDOC increase in the brain of normal phase of the menstrual cycle, which is associated with high cir- animals. In animals subjected to chronic long-term stress, allo- culating levels of allopregnanolone (Sundstro and Backstrom, pregnanolone concentrations decreased in cortex at rest. However, 1999). a larger increase of allopregnanolone was observed in cortex Enzymes involving the production of allopregnanolone are after acute stress in this group of animals than normal con- present in the hippocampus (Compagnone and Mellon, 2000). trols (Serra et al., 2000). Patients with AD show an increased Using the Morris water maze paradigm, allopregnanolone was production compared to healthy elderly control found to inhibit learning (Johansson et al., 2002). Antagonist subjects. In addition, an increased 5α-reduction was seen in neurosteroid 3β,20β-dihydroxy-5α-pregnane (UC1011), reduced patients with AD. Thus an increased glucocorticoid production the negative effect of allopregnanolone on learning in the Morris can be regarded as an early feature of AD since an enhanced water maze (Turkmen et al., 2004). PS infused into the basal production of 5α-reduced metabolites of cortisol was established

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 13 Wang Neurosteroids and GABAA-receptor function

(Rasmuson et al., 2001). 3α-OH-5α-reduced metabolites of cor- noradrenaline release in response to interleukin-1β (IL-1β) in late tisol interact with the GABAA-receptor and enhance the effect pregnancy (Brunton et al., 2009). Allopregnanolone signals the of GABA-steroids on the GABAA-receptor (Stromberg et al., pregnancy status of the animal to the brain and stimulates opi- 2005). oid production in the brainstem. Allopregnanolone also act via Chronic stress can impair memory (de Quervain et al., 1998). GABAA-receptors to enhance GABA action in the PVN. In rat, Memory impairment is permanent in persons with a chroni- allopregnanolone concentration increase greatly in both the cir- cally elevated adrenal production of GABA-steroids (Lupien et al., culation and the brain in pregnancy and reach a peak on day 2005). Memory impairment was also reported in chronic stress 19–20 of pregnancy (Concas et al., 1998; Brunton et al., 2009). syndromes, so called “burn-out syndrome.”“Burn-out syndrome” Parvocellular corticotropin-releasing hormone (CRH) neurons in occurs frequent in patients with AD. The production of corti- the PVN are under direct inhibitory GABAergic control (Mik- sol and GABA-steroids increased in parallel during stress (Purdy los and Kovacs, 2002). 5α-reductase activity is increased in the et al., 1991; Serra et al., 2003; Droogleever Fortuyn et al., 2004). hypothalamus in late pregnancy (Brunton et al., 2009), suggest- During chronic stress, long-term exposure to GABAA-receptor ing that the capacity to produce allopregnanolone is increased in agonist results in similar changes after prolonged exposure to pregnancy. Allopregnanolone may act locally to enhance GABA benzodiazepine and alcohol. Long-term and enhanced activation action in the PVN or on inputs to the CRH neurons to attenuate of the GABAA-receptor is an important factor in of memory HPA axis responses to stress in late pregnancy. The importance impairment during stress. The response of cortisol and GABA- of allopregnanolone to inhibit HPA axis responses to stress in steroids to adrenal stimulation was similar in patients with AD later pregnancy has been tested by blocking its production using a as chronically stressed animals (Nasman et al., 1991, 1996). 5αR inhibitor, finasteride (FIN). Blocking allopregnanolone gen- Patients with mild Alzheimer’s disease have a high and non- eration with finasteride (FIN) at a dose shown to reduce brain suppressible production of cortisol and GABA-steroids (Nasman allopregnanolone content by up to 90% (Concas et al., 1998), et al., 1995). substantially restores HPA axis responses to systemically adminis- After long-term exposure to GABA-steroids, down-regulation tered IL-1β in late pregnant rats (Brunton et al., 2009). AP appears of the GABAA-receptorisexpected(Yu and Ticku, 1995b; Barnes, to depend upon the actions of endogenous to exert its 1996). A malfunctioned GABAA-receptor can be an important suppressive effects on HPA axis activity (Brunton et al., 2009). factor in the pathogenesis of stress-induced , “burn- Allopregnanolone induces tone (as revealed by out” syndrome and sex-steroid related depression (Drugan et al., treatment) over ACTH responses to immune challenge in vir- 1989; Wihlback et al., 2006). The down-regulation occurs at dif- gin rats (Brunton et al., 2009). Induction of inhibitory opioid ferent levels in a time dependent manner: (i) desensitization; (ii) tone over HPA axis responses to IL-1β by AP treatment occurs receptor internalization; (iii) receptor subunit degradation; (iv) within 20 h, consistent with the rapid increase in pENK-A mRNA altered expression of receptor mRNA (Barnes, 1996). Exposure to expression in the NTS that has been seen after IL-1β treatment an agonist of the GABAA-receptor may cause changes in recep- (Engstrom et al., 2003). The mechanism by which AP upregulates tor mRNA and induce changes of the GABAA-receptor subunit opioid expression in the NTS is not clear. However, interac- composition (Smith et al., 1998). In women with PMDD, reduced tion with GABAA-receptors is a possibility (Blyth et al., 2000) benzodiazepine, ethanol, and GABA-steroid sensitivities occur in as has been described for CRH and gene expres- the luteal phase, but not in the follicular phase of the men- sion in the PVN (Bali and Kovacs, 2003) and for hypothala- strual cycle (Sundstrom et al., 1997, 1998; Nyberg et al., 2004). mic oxytocin gene expression at the end of pregnancy (Blyth In rodents, repeated administrations of GABA-steroids caused et al., 2000). It is possible that allopregnanolone-induced opi- tolerance development (Birzniece et al., 2006a) in terms of GABA- oid tone represents a global mechanism through which HPA steroid-induced inhibition on learning (Turkmen et al., 2006). axis responses to other stressful stimuli are restrained in late It is well-known that prolonged exposure to allopregnanolone pregnancy. This pregnancy adaptation is expected to protect the altered the function of α4 subunit and α1 subunit (Smith et al., from adverse early life programming by maternal glu- 1998). In fact, the α4β2γ2 receptor is less sensitive to steroid cocorticoids, although this protection is obviously incomplete modulation than the α1β2γ2 receptor (Belelli et al., 2002). In (Brunton and Russell, 2010, 2011). Withdrawal of these mech- an earlier study on tolerance development in rats, acute tol- anisms or failure to maintain them appropriately might pre- erance against allopregnanolone-induced anesthesia developed dispose the fetuses to disease in later life (Welberg and Seckl, after 90 min of exposure and change in GABAA-receptor α4- 2001). subunit expression was observed in thalamus (Birzniece et al., 2006a). ACKNOWLEDGMENTS The author thank Torbjörn Bäckström and David Haage for ALLOPREGNANOLONE AND SUPPRESSED HPA AXIS STRESS advices and critical reading of an early version of the manu- RESPONSES IN LATE PREGNANCY OF RAT script. This work was supported by the Swedish Research Council- In late pregnancy, hypothalamo-pituitary–adrenal (HPA) axis medicine (project 73p-15450), Insamlingsstiftelsen för medi- responses to stressful stimuli are suppressed (Brunton and Rus- cinsk forskning vid Umeå Universitet, Svenska läkaresällskapets sell, 2003). Up-regulated opioid peptide produced by nucleus forskare program, ALF medel from västerbottens lans landsting, tractus solitarii (NTS) neurons acts presynaptically on noradren- Magn. Bergvalls Stiftelse and the EU regional fund’s objective 1 ergic terminals in the paraventricular nucleus (PVN) to inhibit program.

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 14 Wang Neurosteroids and GABAA-receptor function

REFERENCES G., Wang, M., and Zhu, D. (2003). and reward: allopregnanolone pro- and lateral septum. Brain Res. 850, Adkins, C. E., Pillai, G. V., Kerby, J., Bon- The role of hormones and hor- duces a conditioned place aversion 217–224. nert, T. P.,Haldon, C., McKernan, R. monal treatments in premenstrual in rats. Pharmacol. Biochem. Behav. Bitran, D., Shiekh, M., and McLeod, M., Gonzalez, J. E., Oades, K., Whit- syndrome. CNS Drugs 17, 325–342. 67, 29–35. M. (1995). Anxiolytic effect of ing, P. J., and Simpson, P. B. (2001). Backstrom, T., Gee, K. W., Lan, Belelli, D., Casula,A., Ling,A., and Lam- progesterone is mediated by the alpha4beta3delta GABA(A) recep- N., Sorensen, M., and Wahlstrom, bert, J. J. (2002). The influence of neurosteroid allopregnanolone at tors characterized by fluorescence G. (1990). Steroids in relation subunit composition on the interac- brain GABAA receptors. J. Neuroen- resonance energy transfer-derived to epilepsy and anaesthesia. Ciba tion of neurosteroids with GABA(A) docrinol. 7, 171–177. measurements of membrane poten- Found. Symp. 153, 225–230; discus- receptors. Neuropharmacology 43, Bixo, M., Andersson, A., Winblad, tial. J. Biol. Chem. 276,38934–38939. sion 230–239. 651–661. B., Purdy, R. H., and Back- Akk, G., Bracamontes, J., and Stein- Backstrom, T., Haage, D., Lofgren, M., Belelli, D., and Herd, M. B. strom, T. (1997). Progesterone, bach,J. H. (2001). Pregnenolone sul- Johansson, I. M., Stromberg, J., (2003). The contraceptive agent 5alpha-pregnane-3,20-dione and fate block of GABA(A) receptors: Nyberg, S., Andreen, L., Ossewaarde, Provera enhances GABA(A) 3alpha-hydroxy-5alpha-pregnane- mechanism and involvement of a L., Wingen, G. A., Turkmen, S., and receptor-mediated inhibitory 20-one in specific regions of the residue in the M2 region of the Bengtsson, S. K. (2011). Paradoxical neurotransmission in the rat human female brain in different alpha subunit. J. Physiol. (Lond.) 532, effects of GABA-A modulators may hippocampus: evidence for endoge- endocrine states. Brain Res. 764, 673–684. explain sex steroid induced negative nous neurosteroids? J. Neurosci. 23, 173–178. Akk, G., Li, P., Bracamontes, J., Reichert, mood symptoms in some persons. 10013–10020. Bjorn, I., Bixo, M., Nojd, K. S., Coll- D. E., Covey, D. F., and Steinbach, J. Neuroscience 191, 46–54. Belelli, D., Herd, M. B., Mitchell, E. A., berg, P., Nyberg, S., Sundstrom- H. (2008). Mutations of the GABA-A Backstrom, T., Zetterlund, B., Blom, S., Peden, D. R., Vardy, A. W., Gen- Poromaa, I., and Backstrom, T. receptor alpha1 subunit M1 domain and Romano, M. (1984). Effects of tet, L., and Lambert, J. J. (2006). (2002). The impact of different doses reveal unexpected complexity for intravenous progesterone infusions Neuroactive steroids and inhibitory of medroxyprogesterone acetate on modulation by neuroactive steroids. on the epileptic discharge frequency neurotransmission: mechanisms of mood symptoms in sequential hor- Mol. Pharmacol. 74, 614–627. in women with partial epilepsy. Acta action and physiological relevance. monal therapy. Gynecol. Endocrinol. Akk, G., Shu, H. J., Wang, C., Steinbach, Neurol. Scand. 69, 240–248. Neuroscience 138, 821–829. 16, 1–8. J. H., Zorumski, C. F., Covey, D. F., Bai, D., Zhu, G., Pennefather, P., Jackson, Belelli, D., and Lambert, J. J. (2005). Bjorn, I., Sundstrom-Poromaa, I., Bixo, and Mennerick, S. (2005). Neuros- M. F., MacDonald, J. F., and Orser, Neurosteroids: endogenous regula- M., Nyberg, S., Backstrom, G., and teroid access to the GABAA receptor. B. A. (2001). Distinct functional and tors of the GABA(A) receptor. Nat. Backstrom, T. (2003). Increase of J. Neurosci. 25, 11605–11613. pharmacological properties of tonic Rev. Neurosci. 6, 565–575. estrogen dose deteriorates mood Amin,J.,and Weiss,D. S. (1994). Homo- and quantal inhibitory postsynap- Ben-Porath, D. D., and Taylor, S. P. during progestin phase in sequen- meric rho 1 GABA channels: activa- tic currents mediated by gamma- (2002). The effects of diazepam (val- tial hormonal therapy. J. Clin. tion properties and domains. Recept. aminobutyric acid(A) receptors in ium) and aggressive disposition on Endocrinol. Metab. 88, 2026–2030. Channels 2, 227–236. hippocampal neurons. Mol. Phar- human aggression: an experimen- Blyth, B. J., Hauger, R. L., Purdy, R. H., Amin, J., and Weiss, D. S. (1996). macol. 59, 814–824. tal investigation. Addict. Behav. 27, and Amico, J. A. (2000). The neuros- Insights into the activation mech- Bali, B., and Kovacs, K. J. (2003). 167–177. teroid allopregnanolone modulates anism of rho1 GABA receptors GABAergic control of neuropep- Bicikova, M., Lapcik, O., Hampl, R., oxytocin expression in the hypothal- obtained by coexpression of wild tide gene expression in parvocellular Starka, L., Knuppen, R., Haupt, amic paraventricular nucleus. Am. J. type and activation-impaired sub- neurons of the hypothalamic par- O., and Dibbelt, L. (1995). A Physiol. Regul. Integr. Comp. Physiol. units. Proc. Biol. Sci. 263, 273–282. aventricular nucleus. Eur. J. Neu- novel radioimmunoassay of allo- 278 R684–R691. Andreen, L., Sundstrom-Poromaa, I., rosci. 18, 1518–1526. pregnanolone. Steroids 60, 210–213. Bormann, J. (2000). The‘ABC’of GABA Bixo, M., Andersson, A., Nyberg, S., Banerjee, P. K., and Snead, O. C. III. Biggio, G., Concas, A., Corda, M. G., receptors. Trends Pharmacol. Sci. 21, and Backstrom, T. (2005). Relation- (1998). Neuroactive steroids exac- Giorgi, O., Sanna, E., and Serra, M. 16–19. ship between allopregnanolone and erbate gamma-hydroxybutyric acid- (1990). GABAergic and dopaminer- Boue-Grabot, E., Taupignon, A., negative mood in postmenopausal induced absence in rats. Eur. gic transmission in the rat cerebral Tramu, G., and Garret, M. (2000). women taking sequential hormone J. Pharmacol. 359, 41–48. cortex: effect of stress, anxiolytic and Molecular and electrophysiological replacement therapy with vagi- Barbaccia, M. L., Concas, A., Serra, M., anxiogenic drugs. Pharmacol. Ther. evidence for a GABAc receptor nal progesterone. Psychoneuroen- and Biggio, G. (1998). Stress and 48, 121–142. in thyrotropin-secreting cells. docrinology 30, 212–224. neurosteroids in adult and aged rats. Birzniece, V., Turkmen, S., Lindblad, Endocrinology 141, 1627–1632. Andreen, L., Sundstrom-Poromaa, I., Exp. Gerontol. 33, 697–712. C., Zhu, D., Johansson, I. M., Brickley, S. G., Cull-Candy, S. G., and Bixo, M., Nyberg, S., and Backstrom, Barker, M. J., Greenwood, K. M., Jack- Backstrom, T., and Wahlstrom, G. Farrant, M. (1996). Development of T. (2006). Allopregnanolone con- son, M., and Crowe, S. F. (2004). (2006a). GABA(A) receptor changes a tonic form of synaptic inhibition centrationandmood–abimodal Persistence of cognitive effects after in acute allopregnanolone tolerance. in rat cerebellar granule cells result- association in postmenopausal withdrawal from long-term benzo- Eur. J. Pharmacol. 535, 125–134. ing from persistent activation of women treated with oral proges- diazepine use: a meta-analysis. Arch. Birzniece, V., Backstrom, T., Johansson, GABAA receptors. J. Physiol. (Lond.) terone. Psychopharmacology (Berl.) Clin. Neuropsychol. 19, 437–454. I. M.,Lindblad,C.,Lundgren,P.,Lof- 497(Pt 3), 753–759. 187, 209–221. Barnes, E. M. Jr. (1996). Use-dependent gren, M., Olsson, T., Ragagnin, G., Brickley, S. G., Cull-Candy, S. G., and Angold, A., Costello, E. J., Erkanli, regulation of GABAA receptors. Int. Taube, M., Turkmen, S., Wahlstrom, Farrant, M. (1999). Single-channel A., and Worthman, C. M. (1999). Rev. Neurobiol. 39, 53–76. G.,Wang,M. D.,Wihlback,A. C.,and properties of synaptic and extrasy- Pubertal changes in hormone levels Baulieu, E. E., and Robel, P. (1995). Zhu,D. (2006b). Neuroactive steroid naptic GABAA receptors suggest dif- and depression in girls. Psychol. Med. Non-genomic mechanisms of action effects on cognitive functions with ferential targeting of receptor sub- 29, 1043–1053. of steroid hormones. Ciba Found. a focus on the and GABA types. J. Neurosci. 19, 2960–2973. Backstrom, T. (1976). Epilepsy in Symp. 191, 24–37; discussion 37–42. systems. Brain Res. Brain Res. Rev. 51, Brickley, S. G., Farrant, M., Swan- women. Oestrogen and proges- Baulieu, E. E., Robel, P., and Schu- 212–239. son, G. T., and Cull-Candy, S. G. terone plasma levels. Experientia 32, macher, M. (2007). Neurosteroids: Bitran, D., Dugan, M., Renda, P., Ellis, (2001). CNQX increases GABA- 248–249. beginning of the story. Int. Rev. R., and Foley, M. (1999). Anxiolytic mediated synaptic transmission in Backstrom, T., Andreen, L., Birzniece, Neurobiol. 46, 1–32. effects of the neuroactive steroid the cerebellum by an AMPA/kainate V., Bjorn, I., Johansson, I. M., Beauchamp, M. H., Ormerod, B. K., pregnanolone (3 alpha-OH-5 beta- receptor-independent mecha- Nordenstam-Haghjo, M., Nyberg, S., Jhamandas, K., Boegman, R. J., and pregnan-20-one) after microinjec- nism. Neuropharmacology 41, Sundstrom-Poromaa, I., Wahlstrom, Beninger, R. J. (2000). Neurosteroids tion in the dorsal hippocampus 730–736.

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 15 Wang Neurosteroids and GABAA-receptor function

Brown, N., Kerby, J., Bonnert, T. P., Caraiscos, V. B., Elliott, E. M., You- fragmentographic quantitation of in the brain, plasma, and steroido- Whiting, P. J., and Wafford, K. A. Ten, K. E., Cheng, V. Y., Belelli, 3 alpha-hydroxy-5 alpha-pregnan- genic glands of male and female rats. (2002). Pharmacological character- D., Newell, J. G., Jackson, M. F., 20-one (allopregnanolone) and its Endocrinology 133, 1003–1009. ization of a novel cell line express- Lambert, J. J., Rosahl, T. W., Waf- precursors in blood and brain of Covey, D. F., Han, M., Kumar, A. S., ing human alpha(4)beta(3)delta ford, K. A., MacDonald, J. F., and adrenalectomized and castrated de La Cruz, M. A., Meadows, E. GABA(A) receptors. Br. J. Pharma- Orser, B. A. (2004). Tonic inhibi- rats. J. Neurosci. 15, 4641–4650. S., Hu, Y., Tonnies, A., Nathan, col. 136, 965–974. tion in mouse hippocampal CA1 Cherek, D. R., Spiga, R., and Egli, M. D., Coleman, M., Benz, A., Evers, Brunton, P. J., McKay, A. J., Ochedal- pyramidal neurons is mediated by (1992). Effects of response require- A. S., Zorumski, C. F., and Men- ski, T., Piastowska, A., Rebas, E., alpha5 subunit-containing gamma- ment and alcohol on human aggres- nerick, S. (2000). Neurosteroid Lachowicz, A., and Russell, J. A. aminobutyric acid type A receptors. sive responding. J. Exp. Anal. Behav. analogues. 8. Structure-activity (2009). Central opioid inhibition of Proc. Natl. Acad. Sci. U.S.A. 101, 58, 577–587. studies of N-acylated 17a-aza- neuroendocrine stress responses in 3662–3667. Collinson, N., Kuenzi, F. M., Jarolimek, D-homosteroid analogues of pregnancy in the rat is induced by Carl, P., Hogskilde, S., Nielsen, J. W., W., Maubach, K. A., Cothliff, R., the anesthetic steroids (3alpha, the neurosteroid allopregnanolone. Sorensen, M. B., Lindholm, M., Sur, C., Smith, A., Out, F. M., 5alpha)- and (3alpha,5beta)-3- J. Neurosci. 29, 6449–6460. Karlen, B., and Backstrom, T. (1990). Howell, O., Atack, J. R., McKernan, hydroxypregnan-20-one. J. Med. Brunton, P. J., and Russell, J. A. (2003). Pregnanolone emulsion. A prelimi- R. M., Seabrook, G. R., Dawson, Chem. 43, 3201–3204. Hypothalamic-pituitary-adrenal nary pharmacokinetic and pharma- G. R., Whiting, P. J., and Rosahl, Crawley, J. N., Glowa, J. R., Majew- responses to centrally administered codynamic study of a new intra- T. W. (2002). Enhanced learning ska, M. D., and Paul, S. M. (1986). orexin-A are suppressed in preg- venous anaesthetic agent. Anaesthe- and memory and altered GABAer- Anxiolytic activity of an endoge- nant rats. J. Neuroendocrinol. 15, sia 45, 189–197. gic synaptic transmission in mice nous . Brain Res. 398, 633–637. Carter, R. B., Wood, P. L., Wieland, S., lacking the alpha 5 subunit of the 382–385. Brunton, P. J., and Russell, J. A. (2010). Hawkinson, J. E., Belelli, D., Lam- GABAA receptor. J. Neurosci. 22, Crestani, F., Keist, R., Fritschy, J. Prenatal in the rat bert, J. J., White, H. S., Wolf, H. H., 5572–5580. M., Benke, D., Vogt, K., Prut, L., programmes neuroendocrine and Mirsadeghi, S., Tahir, S. H., Bolger, Compagnone, N. A., and Mellon, S. H. Bluthmann, H., Mohler, H., and behavioural responses to stress in the M. B., Lan, N. C., and Gee, K. W. (2000). Neurosteroids: biosynthesis Rudolph, U. (2002). Trace fear adult offspring: sex-specific effects. J. (1997). Characterization of the anti- and function of these novel neuro- conditioning involves hippocam- Neuroendocrinol. 22, 258–271. convulsant properties of modulators. Front. Neuroendocrinol. pal alpha5 GABA(A) receptors. Brunton, P. J., and Russell, J. A. (2011). (CCD 1042; 3alpha-hydroxy-3beta- 21, 1–56. Proc. Natl. Acad. Sci. U.S.A. 99, Allopregnanolone and suppressed methyl-5alpha-pregnan-20-one), a Concas, A., Follesa, P., Barbaccia, M. 8980–8985. hypothalamo-pituitary-adrenal axis selective, high-affinity, steroid mod- L., Purdy, R. H., and Biggio, G. Crestani, F., Lorez, M., Baer, K., Ess- stress responses in late pregnancy in ulator of the gamma-aminobutyric (1999). Physiological modulation rich, C., Benke, D., Laurent, J. P., the rat. Stress 14, 6–12. acid(A) receptor. J. Pharmacol. Exp. of GABA(A) receptor plasticity by Belzung, C., Fritschy, J. M., Luscher, Brussaard, A. B., Kits, K. S., Baker, R. E., Ther. 280, 1284–1295. progesterone metabolites. Eur. J. B.,and Mohler,H. (1999). Decreased Willems, W. P., Leyting-Vermeulen, Casula, M. A., Bromidge, F. A., Pillai, Pharmacol. 375, 225–235. GABAA-receptor clustering results J. W., Voorn, P., Smit, A. B., Bick- G. V., Wingrove, P. B., Martin, K., Concas, A., Mostallino, M. C., Perra, in enhanced anxiety and a bias nell,R. J.,and Herbison,A. E. (1997). Maubach, K., Seabrook, G. R., Whit- C., Lener, R., Roscetti, G., Bar- for threat cues. Nat. Neurosci. 2, Plasticity in fast synaptic inhibition ing, P. J., and Hadingham, K. L. baccia, M. L., Purdy, R. H., and 833–839. of adult oxytocin neurons caused by (2001). Identification of amino acid Biggio, G. (1996). Functional cor- Davies, P.A., Hanna, M. C., Hales, T. G., switch in GABA(A) receptor subunit residues responsible for the alpha5 relation between allopregnanolone and Kirkness, E. F. (1997). Insensi- expression. Neuron 19, 1103–1114. subunit binding selectivity of L- and [35S]-TBPS binding in the tivity to anaesthetic agents conferred Brzozowski, A. M., Pike, A. C., Dauter, 655,708, a benzodiazepine binding brain of rats exposed to , by a class of GABA(A) receptor sub- Z., Hubbard, R. E., Bonn, T., site ligand at the GABA(A) receptor. or stress. Br. J. unit. Nature 385, 820–823. Engstrom, O., Ohman, L., Greene, G. J. Neurochem. 77, 445–451. Pharmacol. 118, 839–846. De Koninck, Y. (2007). Altered chlo- L., Gustafsson, J. A., and Carlquist, Celentano, J. J., and Wong, R. K. (1994). Concas, A., Mostallino, M. C., Porcu, ride in neurological dis- M. (1997). Molecular basis of ago- Multiphasic desensitization of the P., Follesa, P., Barbaccia, M. L., Tra- orders: a new target. Curr. Opin. nism and antagonism in the oestro- GABAA receptor in outside-out bucchi, M., Purdy, R. H., Grisenti, P., Pharmacol. 7, 93–99. gen receptor. Nature 389, 753–758. patches. Biophys. J. 66, 1039–1050. and Biggio, G. (1998). Role of brain de Quervain, D. J., Roozendaal, B., and Burden, P. M., Ai, T. H., Lin, H. Celotti, F., Melcangi, R. C., and Mar- allopregnanolone in the plasticity of McGaugh, J. L. (1998). Stress and Q., Akinci, M., Costandi, M., tini,L. (1992). The 5 alpha-reductase gamma-aminobutyric acid type A impair retrieval of Hambley, T. M., and Johnston, in the brain: molecular aspects and receptor in rat brain during preg- long-term spatial memory. Nature G. A. (2000). Chiral derivatives relation to brain function. Front. nancy and after delivery. Proc. Natl. 394, 787–790. of 2-cyclohexylideneperhydro-4,7- Neuroendocrinol. 13, 163–215. Acad. Sci. U.S.A. 95, 13284–13289. Demirgoren, S., Majewska, M. D., Spi- methanoindenes, a novel class of Chaudron, L. H., Klein, M. H., Rem- Cooper, E. J., Johnston, G. A., and vak, C. E., and London, E. D. (1991). receptor lig- ington, P., Palta, M., Allen, C., Edwards, F. A. (1999). Effects of Receptor binding and electrophysio- and: synthesis, X-ray analysis, and and Essex, M. J. (2001). Predictors, a naturally occurring neurosteroid logical effects of dehydroepiandros- biological activity. J. Med. Chem. 43, prodromes and incidence of post- on GABAA IPSCs during develop- terone sulfate, an antagonist of the 4629-4635. partum depression. J. Psychosom. ment in rat hippocampal or cerebel- GABAA receptor. Neuroscience 45, Burgard, E. C., Tietz, E. I., Neelands, Obstet. Gynaecol. 22, 103–112. lar slices. J. Physiol. (Lond.) 521(Pt 127–135. T. R., and Macdonald, R. L. (1996). Chebib, M., Hinton, T., Schmid, K. L., 2), 437–449. Devor, A., Fritschy, J. M., and Yarom, Properties of recombinant gamma- Brinkworth, D., Qian, H., Matos, S., Cope, D. W., Hughes, S. W., and Y. (2001). Spatial distribution and aminobutyric acid A receptor iso- Kim, H. L., Abdel-Halim, H., Kumar, Crunelli, V. (2005). GABAA subunit composition of GABA(A) forms containing the alpha 5 sub- R. J., Johnston, G. A., and Hanra- receptor-mediated tonic inhibition receptors in the inferior oli- unit subtype. Mol. Pharmacol. 50, han, J. R. (2009). Novel, potent, and in thalamic neurons. J. Neurosci. 25, vary nucleus. J. Neurophysiol. 85, 119–127. selective GABAC antagonists inhibit 11553–11563. 1686–1696. Callachan, H., Cottrell, G. A., Hather, myopia development and facilitate Corpechot, C., Young, J., Calvel, M., Dhar, V., and Murphy, B. E. (1990). N. Y., Lambert, J. J., Nooney, J. M., learning and memory. J. Pharmacol. Wehrey, C., Veltz, J. N., Touyer, G., Double-blind randomized crossover and Peters, J. A. (1987). Modulation Exp. Ther. 328, 448–457. Mouren, M., Prasad, V. V., Banner, trial of luteal phase of the GABAA receptor by prog- Cheney, D. L., Uzunov, D., Costa, C., and Sjovall, J. (1993). Neuros- (Premarin) in the premenstrual esterone metabolites. Proc. R. Soc. E., and Guidotti, A. (1995). teroids: 3 alpha-hydroxy-5 alpha- syndrome (PMS). Psychoneuroen- Lond. B Biol. Sci. 231, 359–369. chromatographic-mass pregnan-20-one and its precursors docrinology 15, 489–493.

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 16 Wang Neurosteroids and GABAA-receptor function

Do Rego, J. L., Seong, J. Y., Burel, Stoichiometry of a ligand-gatedion Frye, C. A. (2001). The role of neuros- after selective reduction of tonic D., Leprince, J., Luu-The, V., Tsut- channel determined by fluorescence teroids and non-genomic effects of inhibition in GABA A receptor sui, K., Tonon, M. C., Pelletier, G., energy transfer. J. Biol. Chem. 274, progestins and in medi- alpha5 subunit-deficient mice. J. and Vaudry, H. (2009). Neuros- 10100–10104. ating sexual receptivity of rodents. Neurophysiol. 95, 2796–2807. teroid biosynthesis: enzymatic path- Fish, E. W., Faccidomo, S., DeBold, J. Brain Res. Brain Res. Rev. 37, Glykys, J., and Mody, I. (2007). ways and neuroendocrine regula- F., and Miczek, K. A. (2001). Alco- 201–222. The main source of ambient tion by neurotransmitters and neu- hol, allopregnanolone and aggres- Frye, C. A., Mermelstein, P. G., and GABA responsible for tonic ropeptides. Front. Neuroendocrinol. sion in mice. Psychopharmacology DeBold, J. F. (1992). Evidence for inhibition in the mouse hip- 30, 259–301. (Berl.) 153, 473–483. a non-genomic action of progestins pocampus. J. Physiol. (Lond.) 582, Dougherty, D. M., Cherek, D. R., and Fisher, J. L., and Macdonald, R. L. on sexual receptivity in hamster ven- 1163–1178. Bennett, R. H. (1996). The effects of (1997). Single channel properties of tral tegmental area but not hypothal- Grishkovskaya, I., Avvakumov, G. V., alcohol on the aggressive respond- recombinant GABAA receptors con- amus. Brain Res. 578, 87–93. Sklenar, G., Dales, D., Hammond, G. ing of women. J. Stud. Alcohol 57, taining gamma 2 or delta subtypes Frye, C. A., Van Keuren, K. R., and L., and Muller, Y. A. (2000). Crystal 178–186. expressed with alpha 1 and beta Erskine, M. S. (1996). Behavioral structure of human - Drasbek, K. R., and Jensen, K. (2006). 3 subtypes in mouse L929 cells. J. effects of 3 alpha-androstanediol. binding globulin: steroid transport THIP, a and antinocicep- Physiol. 505(Pt 2), 283–297. I: modulation of sexual receptivity by a laminin G-like domain. EMBO tive drug, enhances an extrasynap- Flood, J. F., Morley, J. E., and Roberts, E. and promotion of GABA-stimulated J. 19, 504–512. tic GABAA receptor-mediated con- (1992). Memory-enhancing effects chloride flux. Behav. Brain Res. 79, Grobin, A. C., and Morrow, A. ductance in mouse neocortex. Cereb. in male mice of pregnenolone and 109–118. L. (2000). 3alpha-hydroxy-5alpha- Cortex 16, 1134–1141. steroids metabolically derived from Galanopoulou, A. S. (2008). Sexually pregnan-20-one exposure reduces Droogleever Fortuyn, H. A., van it. Proc. Natl. Acad. Sci. U.S.A. 89, dimorphic expression of KCC2 and GABA(A) receptor alpha4 subunit Broekhoven, F., Span, P. N., Back- 1567–1571. GABA function. Epilepsy Res. 80, mRNA levels. Eur. J. Pharmacol. 409, strom, T., Zitman, F. G., and Verkes, Flood, J. F., Morley, J. E., and 99–113. R1–R2. R. J. (2004). Effects of PhD examina- Roberts, E. (1995). Pregnenolone Gallo, M. A., and Smith, S. S. (1993). Grunewald, R. A., Aliberti, V., and tion stress on allopregnanolone and sulfate enhances post-training mem- Progesterone withdrawal decreases Panayiotopoulos, C. P. (1992). Exac- cortisol plasma levels and periph- ory processes when injected in very latency to and increases duration of erbation of typical absence seizures eral benzodiazepine receptor den- low doses into limbic system struc- electrified prod burial: a possible rat by progesterone. Seizure 1, 137–138. sity. Psychoneuroendocrinology 29, tures: the amygdala is by far the most model of PMS anxiety. Pharmacol. Gulacsi, A., Lee, C. R., Sik, A., 1341–1344. sensitive. Proc. Natl. Acad. Sci. U.S.A. Biochem. Behav. 46, 897–904. Viitanen, T., Kaila, K., Tepper, Drugan, R. C., Morrow, A. L., Weiz- 92, 10806–10810. Garrett, K. M., and Gan, J. (1998). J. M., and Freund, T. F. (2003). man, R., Weizman, A., Deutsch, S. Flood, J. F., Smith, G. E., and Roberts, Enhancement of gamma- Cell type-specific differences in I., Crawley, J. N., and Paul, S. M. E. (1988). Dehydroepiandrosterone aminobutyric acidA receptor chloride-regulatory mechanisms (1989). Stress-induced behavioral and its sulfate enhance memory activity by alpha-. J. and GABA(A) receptor-mediated depression in the rat is associated retention in mice. Brain Res. 447, Pharmacol. Exp. Ther. 285, 680–686. inhibition in rat substantia nigra. J. with a decrease in GABA receptor- 269–278. Gasior, M., Carter, R. B., and Witkin, Neurosci. 23, 8237–8246. mediated chloride ion flux and brain Follesa, P., Concas, A., Porcu, P., Sanna, J. M. (1999). Neuroactive steroids: Gulinello, M., Gong, Q. H., Li, X., benzodiazepine receptor occupancy. E., Serra, M., Mostallino, M. C., potential therapeutic use in neu- and Smith, S. S. (2001). Short-term Brain Res. 487, 45–51. Purdy, R. H., and Biggio, G. (2001). rological and psychiatric disorders. exposure to a neuroactive steroid Edwards, H. E.,Vimal, S., and Burnham, Role of allopregnanolone in regu- Trends Pharmacol. Sci. 20, 107–112. increases alpha4 GABA(A) receptor W. M. (2005). The acute anticonvul- lation of GABA(A) receptor plas- Gee, K. W., Chang, W. C., Brinton, R. E., subunit levels in association with sant effects of deoxycorticosterone ticity during long-term exposure and McEwen, B. S. (1987). GABA- increased anxiety in the female rat. in developing rats: role of metabo- to and withdrawal from proges- dependent modulation of the Cl- Brain Res. 910, 55–66. lites and -receptor terone. Brain Res. Brain Res. Rev. 37, ionophore by steroids in rat brain. Gyermek, L., Iriarte, J., and Crabbe, P. responses. Epilepsia 46, 1888–1897. 81–90. Eur. J. Pharmacol. 136, 419–423. (1968). Steroids. CCCX. Structure- Eisenman, L. N., He, Y., Fields, C., Follesa, P., Floris, S., Tuligi, G., Gee, K. W., and Lan, N. C. (1991). activity relationship of some Zorumski, C. F., and Mennerick, S. Mostallino, M. C., Concas, A., Gamma-aminobutyric acidA recep- steroidal hypnotic agents. J. Med. (2003). Activation dependent prop- and Biggio, G. (1998). Molecular tor complexes in rat frontal cor- Chem. 11, 117–125. erties of pregnenolone sulfate inhi- and functional adaptation of the tex and spinal cord show differen- Gyermek, L., and Soyka, L. F. (1975). bition of GABAA receptor-mediated GABA(A) receptor complex dur- tial responses to steroid modulation. Steroid anesthetics. Anesthesiology current. J. Physiol. 550, 679–691. ing pregnancy and after delivery in Mol. Pharmacol. 40, 995–999. 42, 331–344. Engstrom, L., Engblom, D., and the rat brain. Eur. J. Neurosci. 10, Genazzani, A. R., Petraglia, F., Bernardi, Haage, D., Backstrom, T., and Johans- Blomqvist, A. (2003). Systemic 2905–2912. F., Casarosa, E., Salvestroni, C., son, S. (2005). Interaction between immune challenge induces pre- Fortin, D., Adams, R., and Gallez, Tonetti, A., Nappi, R. E., Luisi, S., allopregnanolone and preg- proenkephalin gene transcription A. (2004). A blood-brain barrier Palumbo, M., Purdy, R. H., and nenolone sulfate in modulating in distinct autonomic structures of disruption model eliminating the Luisi, M. (1998). Circulating lev- GABA-mediated synaptic currents the rat brain. J. Comp. Neurol. 462, hemodynamic effect of . els of allopregnanolone in humans: in neurons from the rat medial 450–461. Can. J. Neurol. Sci. 31, 248–253. gender, age, and endocrine influ- preoptic nucleus. Brain Res. 1033, Essrich, C., Lorez, M., Benson, J. A., Freeman, E. W., Sammel, M. D., Liu, ences. J. Clin. Endocrinol. Metab. 83, 58–67. Fritschy, J. M., and Luscher, B. L., Gracia, C. R., Nelson, D. B., and 2099–2103. Haage, D., Druzin, M., and Johansson, (1998). Postsynaptic clustering of Hollander, L. (2004). Hormones and Gibbs, T. T., Russek, S. J., and Farb, D. H. S. (2002). Allopregnanolone modu- major GABAA receptor subtypes menopausal status as predictors of (2006). Sulfated steroids as endoge- lates spontaneous GABA release via requires the gamma 2 subunit and depression in women in transition nous neuromodulators. Pharmacol. presynaptic Cl- permeability in rat . Nat. Neurosci. 1, 563–571. to . Arch. Gen. Psychiatry Biochem. Behav. 84, 555–567. preoptic nerve terminals. Brain Res. Farrant, M., and Nusser, Z. (2005). Vari- 61, 62–70. Glykys, J., Mann, E. O., and Mody, 958, 405–413. ations on an inhibitory theme: pha- Fritschy, J. M., and Brunig, I. (2003). I. (2008). Which GABA(A) recep- Haage, D., and Johansson, S. (1999). sic and tonic activation of GABA(A) Formation and plasticity of tor subunits are necessary for tonic Neurosteroid modulation of synap- receptors. Nat. Rev. Neurosci. 6, GABAergic synapses: physiological inhibition in the hippocampus? J. tic and GABA-evoked currents in 215–229. mechanisms and pathophysiological Neurosci. 28, 1421–1426. neurons from the rat medial pre- Farrar, S. J., Whiting, P. J., Bonnert, implications. Pharmacol. Ther. 98, Glykys, J., and Mody, I. (2006). optic nucleus. J. Neurophysiol. 82, T. P., and McKernan, R. M. (1999). 299–323. Hippocampal network hyperactivity 143–151. www.frontiersin.org October 2011 | Volume 2 | Article 44 | 17 Wang Neurosteroids and GABAA-receptor function

Haas,K. F.,and Macdonald,R. L. (1999). and Buckwalter, J. G. (1994). GABAA receptor mediates tonic Koksma, J. J., van Kesteren, R. E., GABAA receptor subunit gamma2 Estrogen replacement therapy in inhibition in thalamic VB neurons. Rosahl, T. W., Zwart, R., Smit, A. and delta subtypes confer unique older women. Comparisons between J. Neurophysiol. 94, 4491–4501. B., Luddens, H., and Brussaard, A. kinetic properties on recombinant Alzheimer’s disease cases and non- Johansson, I. M., Birzniece, V., Lind- B. (2003). Oxytocin regulates neu- GABAA receptor currents in mouse demented control subjects. Arch. blad, C., Olsson, T., and Backstrom, rosteroid modulation of GABA(A) fibroblasts. J. Physiol. (Lond.) 514(Pt Neurol. 51, 896–900. T. (2002). Allopregnanolone inhibits receptors in supraoptic nucleus 1), 27–45. Hill, M., Bicikova, M., Parizek, A., Hav- learning in the Morris water maze. around parturition. J. Neurosci. 23, Hadingham, K. L., Wingrove, P.B., Waf- likova, H., Klak, J., Fajt, T., Meloun, Brain Res. 934, 125–131. 788–797. ford, K. A., Bain, C., Kemp, J. A., M., Cibula, D., Cegan, A., Sul- Jones, M. V., and Westbrook, G. L. Korpi, E. R., Debus, F., Linden, A. Palmer, K. J., Wilson, A. W., Wilcox, cova, J., Hampl, R., and Starka, L. (1995). Desensitized states prolong M., Malecot, C., Leppa, E., Vekovis- A. S., Sikela, J. M., and Ragan, C. (2001). Neuroactive steroids, their GABAA channel responses to brief cheva, O., Rabe, H., Bohme, I., Aller, I. (1993). Role of the beta subunit precursors and polar conjugates dur- agonist pulses. Neuron 15, 181–191. M. I., Wisden, W., and Luddens, in determining the pharmacology of ing parturition and postpartum in Kahle, K. T., Staley, K. J., Nahed, B. V., H. (2007). Does ethanol act pref- human gamma-aminobutyric acid maternal blood: 2. Time profiles Gamba, G., Hebert, S. C., Lifton, R. erentially via selected brain GABAA typeAreceptors.Mol. Pharmacol. 44, of pregnanolone isomers. J. Steroid P., and Mount, D. B. (2008). Roles of receptor subtypes? The current evi- 1211–1218. Biochem. Mol. Biol. 78, 51–57. the cation-chloride cotransporters dence is ambiguous. Alcohol 41, Halbreich, U. (2003). The etiology, biol- Hillen, T., Lun, A., Reischies, F. M., in neurological disease. Nat. Clin. 163–176. ogy, and evolving pathology of pre- Borchelt, M., Steinhagen-Thiessen, Pract. Neurol. 4, 490–503. Kurthen, M., Linke, D. B., Reuter, B. menstrual syndromes. Psychoneu- E., and Schaub, R. T. (2000). DHEA- Kancheva, L., Havlikova, H., Hill, M., M., Hufnagel, A., and Elger, C. E. roendocrinology 28(Suppl. 3), 55–99. S plasma levels and incidence of Vrbikova, J., and Starka, L. (2007). (1991). Severe negative emotional Hamann, M., Rossi, D. J., and Attwell, Alzheimer’s disease. Biol. Psychiatry “Neuroactive steroids in adult men,” reactions in intracarotid sodium D. (2002). Tonic and spillover inhi- 47, 161–163. in 4th International Meeting – amytal procedures: further evidence bition of granule cells control infor- Hogenkamp, D. J., Tahir, S. H., Hawkin- Steroids and Nervous system, Torino, for hemispheric asymmetries? Cor- mation flow through cerebellar cor- son, J. E., Upasani, R. B., Alaud- 152. tex 27, 333–337. tex. Neuron 33, 625–633. din, M., Kimbrough, C. L., Acosta- Kendell,R. E.,McGuire,R. J.,Connor,Y., Lambert, J. J., Belelli, D., Hill-Venning, Hamilton, N. M. (2002). Interaction of Burruel, M., Whittemore, E. R., and Cox, J. L. (1981). Mood changes C., Callachan, H., and Peters, J. steroids with the GABA(A) receptor. Woodward, R. M., Lan, N. C., in the first three weeks after child- A. (1996). Neurosteroid modulation Curr. Top. Med. Chem. 2, 887–902. Gee, K. W., and Bolger, M. B. birth. J. Affect. Disord. 3, 317–326. of native and recombinant GABAA Harney, S. C., Frenguelli, B. G., and (1997). Synthesis and in vitro activ- Kennedy, R. T., Thompson, J. E., and receptors. Cell.Mol.Neurobiol.16, Lambert, J. J. (2003). Phosphoryla- ity of 3 beta-substituted-3 alpha- Vickroy, T. W. (2002). In vivo mon- 155–174. tion influences neurosteroid mod- hydroxypregnan-20-ones: allosteric itoring of amino acids by direct Lambert, J. J., Belelli, D., Hill-Venning, ulation of synaptic GABAA recep- modulators of the GABAA receptor. sampling of brain extracellular fluid C., and Peters, J. A. (1995). Neuros- tors in rat CA1 and dentate gyrus J. Med. Chem. 40, 61–72. at ultralow flow rates and capillary teroids and GABAA receptor func- neurones. Neuropharmacology 45, Hosie, A. M., Clarke, L., da Silva, H., electrophoresis. J. Neurosci. Methods tion. Trends Pharmacol. Sci. 16, 873–883. and Smart, T. G. (2009). Conserved 114, 39–49. 295–303. Harris, R. A., Proctor, W. R., McQuilkin, site for neurosteroid modulation of Kerr, D. I., and Ong, J. (1992). GABA Lambert, J. J., Harney, S. C., Belelli, S. J.,Klein,R. L.,Mascia,M. P.,What- GABA A receptors. Neuropharma- agonists and antagonists. Med. Res. D., and Peters, J. A. (2001a). ley, V., Whiting, P. J., and Dunwid- cology 56, 149–154. Rev. 12, 593–636. Neurosteroid modulation of die, T. V. (1995). Ethanol increases Hosie, A. M., Dunne, E. L., Harvey, R. Khirug, S., Yamada, J., Afzalov, R., recombinant and synaptic GABAA GABAA responses in cells stably J., and Smart, T. G. (2003). Zinc- Voipio, J., Khiroug, L., and Kaila, K. receptors. Int. Rev. Neurobiol. 46, transfected with receptor subunits. mediated inhibition of GABA(A) (2008). GABAergic 177–205. Alcohol. Clin. Exp. Res. 19, 226–232. receptors: discrete binding sites of the axon initial segment in corti- Lambert, J. J., Belelli, D., Harney, S. C., Havlikova, H., Hill, M., Kancheva, L., underlie subtype specificity. Nat. cal principal neurons is caused by the Peters, J. A., and Frenguelli, B. G. Vrbikova, J., Pouzar, V., Cerny, I., Neurosci. 6, 362–369. Na-K-2Cl cotransporter NKCC1. J. (2001b). Modulation of native and Kancheva, R., and Starka, L. (2006). Hosie, A. M., Wilkins, M. E., da Silva, Neurosci. 28, 4635–4639. recombinant GABA(A) receptors by Serum profiles of free and conju- H. M., and Smart, T. G. (2006). Kittler, J. T., Wang, J., Connolly, C. N., endogenous and synthetic neuroac- gated neuroactive pregnanolone iso- Endogenous neurosteroids regulate Vicini, S., Smart, T. G., and Moss, tive steroids. Brain Res. Brain Res. mers in nonpregnant women of fer- GABAA receptors through two dis- S. J. (2000). Analysis of GABAA Rev. 37, 68–80. tile age. J. Clin. Endocrinol. Metab. crete transmembrane sites. Nature receptor assembly in mammalian Landgren, S., Aasly, J., Backstrom, T., 91, 3092–3099. 444, 486–489. cell lines and hippocampal neurons Dubrovsky, B., and Danielsson, E. Hawkinson, J. E., Kimbrough, C. Hosie, A. M., Wilkins, M. E., and Smart, using gamma 2 subunit green fluo- (1987). The effect of progesterone L., Belelli, D., Lambert, J. J., T. G. (2007). Neurosteroid binding rescent protein chimeras. Mol. Cell. and its metabolites on the interic- Purdy, R. H., and Lan, N. C. sites on GABA(A) receptors. Phar- Neurosci. 16, 440–452. tal epileptiform discharge in the cat’s (1994a). Correlation of neuroactive macol. Ther. 116, 7–19. Klangkalya, B., and Chan, A. (1988). cerebral cortex. Acta Physiol. Scand. steroid modulation of [35S]t- Hutcheon, B., Morley, P., and Poul- Structure-activity relationships of 131, 33–42. butylbicyclophosphorothionate ter, M. O. (2000). Developmental steroid hormones on muscarinic Landgren, S., Wang, M. D., Backstrom, and [3H]flunitrazepam binding change in GABAA receptor desen- receptor binding. J. Steroid Biochem. T., and Johansson, S. (1998). Inter- and gamma-aminobutyric acidA sitization kinetics and its role in 29, 111–118. action between 3 alpha-hydroxy- receptor function. Mol. Pharmacol. synapse function in rat cortical neu- Klausberger, T., Ehya, N., Fuchs, K., 5 alpha-pregnan-20-one and car- 46, 977–985. rons. J. Physiol. (Lond.) 522(Pt 1), Fuchs, T., Ebert, V., Sarto, I., and bachol in the control of neuronal Hawkinson, J. E., Kimbrough, C. L., 3–17. Sieghart, W. (2001). Detection and excitability in hippocampal slices of McCauley, L. D., Bolger, M. B., Lan, Imamura, M., and Prasad, C. (1998). binding properties of GABA(A) female rats in defined phases of the N. C., and Gee, K. W. (1994b). Modulation of GABA-gated chloride receptor assembly intermediates. J. oestrus. Acta Physiol. Scand. 162, The neuroactive steroid 3 alpha- ion influx in the brain by dehy- Biol. Chem. 276, 16024–16032. 77–88. hydroxy-5 beta-pregnan-20-one is a droepiandrosterone and its metabo- Knight, A. R., Stephenson, F. A., Tall- Lanthier, A., and Patwardhan, V. V. two-component modulator of lig- lites. Biochem. Biophys. Res. Com- man, J. F., and Ramabahdran, T. (1986). Sex steroids and 5-en-3 beta- and binding to the GABAA receptor. mun. 243, 771–775. V. (2000). Monospecific antibod- hydroxysteroids in specific regions Eur. J. Pharmacol. 269, 157–163. Jia, F., Pignataro, L., Schofield, C. M., ies as probes for the stoichiometry of the human brain and cra- Henderson, V. W., Paganini-Hill, A., Yue, M., Harrison, N. L., and Gold- of recombinant GABA(A) receptors. nial nerves. J. Steroid Biochem. 25, Emanuel, C. K., Dunn, M. E., stein, P. A. (2005). An extrasynaptic Recept. Channels 7, 213–226. 445–449.

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 18 Wang Neurosteroids and GABAA-receptor function

Laubach,G. D.,P’An,S.Y.,and Rudel,H. Genazzani, A. R. (2000). Serum complex in brain? Brain Res. 404, G., and Vaudry, H. (1999). Neuros- W. (1955). Steroid anesthetic agent. allopregnanolone levels in pregnant 355–360. teroids: expression of steroidogenic Science 122, 78. women: changes during pregnancy, Martina, M., Royer, S., and Pare, enzymes and regulation of steroid Lavoie, A. M., Tingey, J. J., Harrison, N. at delivery, and in hypertensive D. (2001). Cell-type-specific GABA biosynthesis in the central nervous L., Pritchett, D. B., and Twyman, R. patients. J. Clin. Endocrinol. Metab. responses and chloride homeosta- system. Pharmacol. Rev. 51, 63–81. E. (1997). Activation and deactiva- 85, 2429–2433. sis in the cortex and amygdala. J. Meyerson, B. J. (1967). Relation- tion rates of recombinant GABA(A) Lundgren, P., Stromberg, J., Back- Neurophysiol. 86, 2887–2895. ship between the anesthetic and receptor channels are dependent on strom, T., and Wang, M. (2003). Masia, S. L., Perrine, K., West- gestagenic action and estrous alpha-subunit isoform. Biophys. J. Allopregnanolone-stimulated brook, L., Alper, K., and Devin- behavior-inducing activity of dif- 73, 2518–2526. GABA-mediated chloride ion sky, O. (2000). Emotional out- ferent progestins. Endocrinology 81, Lee, G. P., Loring, D. W., Meador, K. flux is inhibited by 3beta- bursts and post-traumatic stress 369–374. J., Flanigin, H. F., and Brooks, B. hydroxy-5alpha-pregnan-20-one disorder during intracarotid amo- Miczek, K. A., Fish, E. W., and De Bold, S. (1988). Severe behavioral com- (isoallopregnanolone). Brain Res. procedure. Neurology 54, J. F. (2003). Neurosteroids, GABAA plications following intracarotid 982, 45–53. 1691–1693. receptors, and escalated aggres- sodium injection: Lupien, S. J., Fiocco, A., Wan, N., Maubach, K. (2003). GABA(A) recep- sive behavior. Horm. Behav. 44, implications for hemispheric asym- Maheu, F., Lord, C., Schramek, tor subtype selective cognition 242–257. metry of emotion. Neurology 38, T., and Tu, M. T. (2005). Stress enhancers. Curr. Drug Targets CNS Mienville, J. M., and Vicini, S. (1989). 1233–1236. hormones and human mem- Neurol. Disord. 2, 233–239. Pregnenolone sulfate antagonizes Lerma, J., Herranz, A. S., Herreras, O., ory function across the lifespan. Mayo,W.,Dellu,F.,Robel,P.,Cherkaoui, GABAA receptor-mediated currents Abraira, V., and Martin del Rio, R. Psychoneuroendocrinology 30, J., Le Moal, M., Baulieu, E. E., via a reduction of channel opening (1986). In vivo determination of 225–242. and Simon, H. (1993). Infusion frequency. Brain Res. 489, 190–194. extracellular concentration of amino Macdonald, R. L., and Olsen, R. W. of neurosteroids into the nucleus Mihalek, R. M., Banerjee, P. K., Korpi, acids in the rat hippocampus. A (1994). GABAA receptor channels. basalis magnocellularis affects cog- E. R., Quinlan, J. J., Firestone, L. method based on brain dialysis and Annu.Rev.Neurosci.17, 569–602. nitive processes in the rat. Brain Res. L., Mi, Z. P., Lagenaur, C., Tretter, computerized analysis. Brain Res. Maione, S., Berrino, L., Vitagliano, S., 607, 324–328. V., Sieghart, W., Anagnostaras, S. G., 384, 145–155. Leyva, J., and Rossi, F. (1992). Preg- McEwen, B. S. (1991). Non-genomic Sage, J. R., Fanselow, M. S., Guidotti, Leroy, C., Poisbeau, P., Keller, A. F., and nenolone sulfate increases the con- and genomic effects of steroids on A., Spigelman, I., Li, Z., DeLorey, Nehlig, A. (2004). Pharmacological vulsant potency of N-methyl-D- neural activity. Trends Pharmacol. T. M., Olsen, R. W., and Homanics, plasticity of GABA(A) receptors at aspartate in mice. Eur. J. Pharmacol. Sci. 12, 141–147. G. E. (1999). Attenuated sensitivity dentate gyrus synapses in a rat model 219, 477–479. McKernan, R. M., and Whiting, P. to neuroactive steroids in gamma- of temporal lobe epilepsy. J. Physiol. Maitra, R., and Reynolds, J. N. J. (1996). Which GABAA-receptor aminobutyrate type A receptor delta (Lond.) 557, 473–487. (1998). Modulation of GABA(A) subtypes really occur in the brain? subunit knockout mice. Proc. Natl. Li, G. D., Chiara, D. C., Cohen, J. B., and receptor function by neuroactive Trends Neurosci. 19, 139–143. Acad. Sci. U.S.A. 96, 12905–12910. Olsen, R. W. (2009). Neurosteroids steroids: evidence for heterogene- Mehta, A. K., and Ticku, M. K. Miklos, I. H., and Kovacs, K. J. allosterically modulate binding of ity of steroid sensitivity of recom- (1999). An update on GABAA recep- (2002). GABAergic innervation of the anesthetic to gamma- binant GABA(A) receptor isoforms. tors. Brain Res. Brain Res. Rev. 29, corticotropin-releasing hormone aminobutyric acid type A receptors. Can. J. Physiol. Pharmacol. 76, 196–217. (CRH)-secreting parvocellular J. Biol. Chem. 284, 11771–11775. 909–920. Melchior, C. L., and Ritzmann, R. neurons and its plasticity as Li, P., Khatri, A., Bracamontes, J., Maitra, R., and Reynolds, J. N. (1999). F. (1996). Neurosteroids block demonstrated by quantitative Weiss, D. S., Steinbach, J. H., and Subunit dependent modulation of the memory-impairing effects immunoelectron microscopy. Akk, G. (2010). Site-specific fluo- GABAA receptor function by neu- of ethanol in mice. Pharmacol. Neuroscience 113, 581–592. rescence reveals distinct structural roactive steroids. Brain Res. 819, Biochem. Behav. 53, 51–56. Miller, L. G., Greenblatt, D. J., Barn- changes induced in the human rho 75–82. Mellon, S. H. (2007). Neurosteroid reg- hill, J. G., and Shader, R. I. (1988). 1 GABA receptor by inhibitory Majewska, M. D. (1992). Neurosteroids: ulation of Chronic benzodiazepine adminis- neurosteroids. Mol. Pharmacol. 77, endogenous bimodal modulators of development. Pharmacol. Ther. 116, tration. I. Tolerance is associated 539–546. the GABAA receptor. Mechanism 107–124. with benzodiazepine receptor down- Li, W., Jin, X., Covey, D. F., and of action and physiological signifi- Mellon, S. H., Griffin, L. D., and Com- regulation and decreased gamma- Steinbach, J. H. (2007). Neuroac- cance. Prog. Neurobiol. 38, 379–395. pagnone, N. A. (2001). Biosynthesis aminobutyric acidA receptor func- tive steroids and human recombi- Majewska, M. D., Demirgoren, S., Spi- and action of neurosteroids. Brain tion. J. Pharmacol. Exp. Ther. 246, nant rho1 GABAC receptors. J. Phar- vak, C. E., and London, E. D. Res. Brain Res. Rev. 37, 3–12. 170–176. macol. Exp. Ther. 323, 236–247. (1990). The neurosteroid dehy- Mennerick, S., He, Y., Jiang, X., Manion, Mitchell, E. A., Herd, M. B., Gunn, Liang, J., Cagetti, E., Olsen, R. W., and droepiandrosterone sulfate is an B. D., Wang, M., Shute, A., Benz, A., B. G., Lambert, J. J., and Belelli, Spigelman, I. (2004). Altered phar- allosteric antagonist of the GABAA Evers,A. S., Covey, D. F., and Zorum- D. (2008). Neurosteroid modula- macology of synaptic and extrasy- receptor. Brain Res. 526, 143–146. ski, C. F. (2004). Selective antag- tion of GABAA receptors: molecu- naptic GABAA receptors on CA1 Majewska, M. D., Harrison, N. L., onism of 5alpha-reduced neuros- lar determinants and significance in hippocampal neurons is consistent Schwartz, R. D., Barker, J. L., and teroid effects at GABA(A) receptors. health and disease. Neurochem. Int. with subunit changes in a model Paul, S. M. (1986). Steroid hor- Mol. Pharmacol. 65, 1191–1197. 52, 588–595. of alcohol withdrawal and depen- mone metabolites are barbiturate- Mennerick, S., Zeng, C. M., Benz, Mitchell, S. J., and Silver, R. A. (2003). dence. J. Pharmacol. Exp. Ther. 310, like modulators of the GABA recep- A., Shen, W., Izumi, Y., Evers, A. Shunting inhibition modulates neu- 1234–1245. tor. Science 232, 1004–1007. S., Covey, D. F., and Zorumski, ronal gain during synaptic excita- Lofgren, M., Johansson, I. M., Mey- Majewska, M. D., Mienville, J. M., C. F. (2001). Effects on gamma- tion. Neuron 38, 433–445. erson, B., Lundgren, P., and Back- and Vicini, S. (1988). Neuros- aminobutyric acid (GABA)(A) Mody, I. (2001). Distinguishing strom, T. (2006). Progesterone with- teroid pregnenolone sulfate antago- receptors of a neuroactive steroid between GABA(A) receptors drawal effects in the open field test nizes electrophysiological responses that negatively modulates glu- responsible for tonic and phasic can be predicted by elevated plus to GABA in neurons. Neurosci. Lett. tamate neurotransmission and conductances. Neurochem. Res. 26, maze performance. Horm. Behav. 50, 90, 279–284. augments GABA neurotrans- 907–913. 208–215. Majewska, M. D., and Schwartz, R. mission. Mol. Pharmacol. 60, Mody, I. (2008). Extrasynaptic GABAA Luisi, S., Petraglia, F., Benedetto, C., D. (1987). Pregnenolone-sulfate: 732–741. receptors in the crosshairs of hor- Nappi, R. E., Bernardi, F., Fadalti, an endogenous antagonist of the Mensah-Nyagan, A. G., Do-Rego, J. L., mones and ethanol. Neurochem. Int. M., Reis, F. M., Luisi, M., and gamma-aminobutyric acid receptor Beaujean, D., Luu-The, V., Pelletier, 52, 60–64. www.frontiersin.org October 2011 | Volume 2 | Article 44 | 19 Wang Neurosteroids and GABAA-receptor function

Mody, I., De Koninck, Y., Otis, T. S., disturbance in the feedback regula- hypothalamo-hypophyseal cocul- GABAA-steroids antagonists,”in 4th and Soltesz, I. (1994). Bridging the tion of the hypothalamic-pituitary- ture model. J. Physiol. (Lond.) International Meeting – Steroids and cleft at GABA synapses in the brain. adrenal axis in the early phase of 500(Pt 2), 475–485. Nervous system, Torino, 197. Trends Neurosci. 17, 517–525. Alzheimer’s disease. Psychoneuroen- Porcello, D. M., Huntsman, M. M., Rahman, M., Lindblad, C., Johans- Moenter, S. M., and DeFazio, R. docrinology 20, 211–220. Mihalek, R. M., Homanics, G. E., son, I. M., Backstrom, T., and A. (2005). Endogenous gamma- Nayeem, N., Green, T. P., Martin, I. L., and Huguenard, J. R. (2003). Intact Wang, M. D. (2006). Neuros- aminobutyric acid can excite and Barnard, E. A. (1994). Quater- synaptic GABAergic inhibition and teroid modulation of recombinant gonadotropin-releasing hormone nary structure of the native GABAA altered neurosteroid modulation of rat alpha(5)beta(2)gamma(2L) neurons. Endocrinology 146, receptor determined by electron thalamic relay neurons in mice lack- and alpha(1)beta(2)gamma(2L) 5374–5379. microscopic image analysis. J. Neu- ing delta subunit. J. Neurophysiol. 89, GABA(A) receptors in Xenopus Mohammed, A. K., Wahlstrom, G., rochem. 62, 815–818. 1378–1386. oocyte. Eur. J. Pharmacol. 547, Tiger, G., Bjorklund, P. E., Sten- Nguyen, Q., Sapp, D. W., Van Ness, P. Prescott, S. A., Sejnowski, T. J., and 37–44. strom, A., Magnusson, O., Archer, C., and Olsen, R. W. (1995). Mod- De Koninck, Y. (2006). Reduction Rasmuson, S., Andrew, R., Nasman, T., Fowler, C. J., and Nordberg, A. ulation of GABAA receptor bind- of anion subverts B., Seckl, J. R., Walker, B. R., and (1987). Impaired performance of ing in human brain by neuroactive the inhibitory control of firing rate Olsson, T. (2001). Increased glu- rats in the Morris swim-maize test steroids: species and brain regional in spinal lamina I neurons: towards cocorticoid production and altered late in abstinence following long- differences. Synapse 19, 77–87. a biophysical basis for neuropathic cortisol metabolism in women with term sodium barbital treatment. Nusser, Z., Cull-Candy, S., and Farrant, pain. Mol. Pain 2, 32. mild to moderate Alzheimer’s Drug Alcohol Depend. 20, 203–212. M. (1997). Differences in synaptic Price, T. J., Cervero, F., and de Koninck, disease. Biol. Psychiatry 49, Mohler, H., Crestani, F., and Rudolph, GABA(A) receptor number underlie Y. (2005). Role of cation-chloride- 547–552. U. (2001). GABA(A)-receptor sub- variation in GABA mini amplitude. cotransporters (CCC) in pain and Reddy, D. S. (2003). Is there a phys- types: a new pharmacology. Curr. Neuron 19, 697–709. hyperalgesia. Curr. Top. Med. Chem. iological role for the neurosteroid Opin. Pharmacol. 1, 22–25. Nusser, Z., and Mody, I. (2002). Selec- 5, 547–555. THDOC in stress-sensitive condi- Mohler, H., Fritschy, J. M., Crestani, F., tive modulation of tonic and phasic Price, T. J., Cervero, F., Gold, M. S., tions? Trends Pharmacol. Sci. 24, Hensch, T., and Rudolph, U. (2004). inhibitions in dentate gyrus granule Hammond, D. L., and Prescott, S. 103–106. Specific GABA(A) circuits in brain cells. J. Neurophysiol. 87, 2624–2628. A. (2009). Chloride regulation in the Reddy, D. S., and Kulkarni, S. K. (2000). development and therapy. Biochem. Nusser, Z., Sieghart, W., and Somogyi, pain pathway. Brain Res. Rev. 60, Development of neurosteroid-based Pharmacol. 68, 1685–1690. P. (1998). Segregation of different 149–170. novel psychotropic drugs. Prog. Med. Mohler, H., Fritschy, J. M., and Rudolph, GABAA receptors to synaptic and Puia, G., Ducic, I., Vicini, S., and Chem. 37, 135–175. U. (2002). A new benzodiazepine extrasynaptic membranes of cere- Costa, E. (1993). Does neuros- Reddy, D. S., and Rogawski, M. pharmacology. J. Pharmacol. Exp. bellar granule cells. J. Neurosci. 18, teroid modulatory efficacy depend A. (2002). Stress-induced Ther. 300, 2–8. 1693–1703. on GABAA receptor subunit compo- deoxycorticosterone-derived neu- Morris, K. D., Moorefield, C. N., and Nyberg, S., Wahlstrom, G., Back- sition? Recept. Channels 1, 135–142. rosteroids modulate GABA(A) Amin, J. (1999). Differential modu- strom, T., and Sundstrom Poro- Puia, G., Santi, M. R., Vicini, S., Pritch- receptor function and seizure lation of the gamma-aminobutyric maa, I. (2004). Altered sensitivity ett, D. B., Purdy, R. H., Paul, S. M., susceptibility. J. Neurosci. 22, acid type C receptor by neuroac- to alcohol in the late luteal phase Seeburg, P. H., and Costa, E. (1990). 3795–3805. tive steroids. Mol. Pharmacol. 56, among patients with premenstrual Neurosteroids act on recombinant Rick, C. E., Ye, Q., Finn, S. E., and 752–759. dysphoric disorder. Psychoneuroen- human GABAA receptors. Neuron 4, Harrison, N. L. (1998). Neuros- Mozrzymas, J. W., Zarnowska, E. D., docrinology 29, 767–777. 759–765. teroids act on the GABA(A) receptor Pytel, M., and Mercik, K. (2003). Ottander, U., Poromaa, I. S., Bjurulf, E., Purdy, R. H., Moore, P. H. Jr., Mor- at sites on the N-terminal side of Modulation of GABA(A) receptors Skytt, A., Backstrom, T., and Olof- row, A. L., and Paul, S. M. (1992). the middle of TM2. Neuroreport 9, by hydrogen reveals synaptic sson, J. I. (2005). Allopregnanolone Neurosteroids and GABAA receptor 379–383. GABA transient and a crucial role of and pregnanolone are produced by function. Adv. Biochem. Psychophar- Riedel, G., Platt, B., and Micheau, the desensitization process. J. Neu- the human corpus luteum. Mol. Cell. macol. 47, 87–92. J. (2003). Glutamate recep- rosci. 23, 7981–7992. Endocrinol. 239, 37–44. Purdy, R. H., Morrow, A. L., Blinn, J. R., tor function in learning and Nakamura, N. H., Rosell, D. R., Akama, Park-Chung, M., Malayev, A., Purdy, and Paul, S. M. (1990). Synthesis, memory. Behav. Brain Res. 140, K. T., and McEwen, B. S. (2004). R. H., Gibbs, T. T., and Farb, metabolism, and pharmacological 1–47. Estrogen and ovariectomy regulate D. H. (1999). Sulfated and unsul- activity of 3 alpha-hydroxy steroids Rodgers, R. J., and Johnson, N. J. (1998). mRNA and protein of fated steroids modulate gamma- which potentiate GABA-receptor- Behaviorally selective effects of neu- decarboxylases and cation-chloride aminobutyric acidA receptor func- mediated chloride ion uptake roactive steroids on plus-maze anx- cotransporters in the adult rat hip- tion through distinct sites. Brain Res. in rat cerebral cortical synap- iety in mice. Pharmacol. Biochem. pocampus. Neuroendocrinology 80, 830, 72–87. toneurosomes. J. Med. Chem. 33, Behav. 59, 221–232. 308–323. Parry, B. L. (2001). The role of cen- 1572–1581. Rudolph, U., Crestani, F., Benke, D., Nasman, B., Olsson, T., Backstrom, T., tral serotonergic dysfunction in the Purdy, R. H., Morrow, A. L., Moore, P. Brunig, I., Benson, J. A., Fritschy, Eriksson, S., Grankvist, K., Viitanen, aetiology of premenstrual dysphoric H. Jr., and Paul, S. M. (1991). Stress- J. M., Martin, J. R., Bluethmann, M., and Bucht, G. (1991). Serum disorder: therapeutic implications. induced elevations of gamma- H., and Mohler, H. (1999). Benzo- dehydroepiandrosterone sulfate in CNS Drugs 15, 277–285. aminobutyric acid type A receptor- diazepine actions mediated by spe- Alzheimer’s disease and in multi- Paul, S. M., and Purdy, R. H. (1992). active steroids in the rat brain. cific gamma-aminobutyric acid(A) infarct dementia. Biol. Psychiatry 30, Neuroactive steroids. FASEB J. 6, Proc. Natl. Acad. Sci. U.S.A. 88, receptor subtypes. Nature 401, 684–690. 2311–2322. 4553–4557. 796–800. Nasman, B., Olsson, T., Fagerlund, Pirker,S.,Schwarzer,C.,Wieselthaler,A., Purvez, D., Augustine, G. J., Fitzpatrick, Rudolph, U., Crestani, F., and Mohler, M., Eriksson, S., Viitanen, M., Sieghart, W., and Sperk, G. (2000). D., Hall, W. C., LaMantia, A. S., H. (2001). GABA(A) receptor sub- and Carlstrom, K. (1996). Blunted GABA(A) receptors: immunocyto- McNamara, J. O., and Williams, S. types: dissecting their pharmacolog- adrenocorticotropin and increased chemical distribution of 13 subunits M. (2004). Neuroscience, 3rd Edn. ical functions. Trends Pharmacol. Sci. adrenal steroid response to human in the adult rat brain. Neuroscience Sunderland: Sinauer Associates Inc., 22, 188–194. corticotropin-releasing hormone in 101, 815–850. 143. Rupprecht, R. (2003). Neuroactive Alzheimer’s disease. Biol. Psychiatry Poisbeau, P., Feltz, P., and Schlichter, Ragagnin, G., Rahman, M., Zingmark, steroids: mechanisms of action and 39, 311–318. R. (1997). Modulation of GABAA E., Stromberg, J., Lundgren, P., neuropsychopharmacological prop- Nasman, B., Olsson, T., Viitanen, M., receptor-mediated IPSCs by Wang, M., and Backstrom, T. (2007). erties. Psychoneuroendocrinology 28, and Carlstrom, K. (1995). A subtle neuroactive steroids in a rat “Structure actvity relationship of 139–168.

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 20 Wang Neurosteroids and GABAA-receptor function

Rupprecht, R., di Michele, F., Hermann, type-specific tonic inhibition. Nat. women: the Women’s Health Ini- in the rat hippocampus I: immuno- B., Strohle, A., Lancel, M., Romeo, Neurosci. 6, 484–490. tiative Memory Study: a random- cytochemical distribution of 13 sub- E.,and Holsboer,F.(2001). Neuroac- Semyanov, A., Walker, M. C., Kullmann, ized controlled trial. JAMA 289, units. Neuroscience 80, 987–1000. tive steroids: molecular mechanisms D. M., and Silver, R. A. (2004). Toni- 2651–2662. Stell, B. M., Brickley, S. G., Tang, of action and implications for neu- cally active GABA A receptors: mod- Sieghart, W. (1992). GABAA recep- C. Y., Farrant, M., and Mody, I. ropsychopharmacology. Brain Res. ulating gain and maintaining the tors: ligand-gated Cl- ion chan- (2003). Neuroactive steroids reduce Brain Res. Rev. 37, 59–67. tone. Trends Neurosci. 27, 262–269. nels modulated by multiple drug- neuronal excitability by selectively Rupprecht, R., Reul, J. M., Trapp, T., van Serra, M., Pisu, M. G., Floris, I., Cara,V., binding sites. Trends Pharmacol. Sci. enhancing tonic inhibition medi- Steensel, B., Wetzel, C., Damm, K., Purdy, R. H., and Biggio, G. (2003). 13, 446–450. ated by delta subunit-containing Zieglgansberger, W., and Holsboer, Social isolation-induced increase in Sieghart, W. (2000). Unraveling the GABAA receptors. Proc. Natl. Acad. F. (1993). Progesterone receptor- the sensitivity of rats to the steroido- function of GABA(A) receptor sub- Sci. U.S.A. 100, 14439–14444. mediated effects of neuroactive genic effect of ethanol. J. Neurochem. types. Trends Pharmacol. Sci. 21, Stell, B. M., and Mody, I. (2002). Recep- steroids. Neuron 11, 523–530. 85, 257–263. 411–413. tors with different affinities mediate Sandstrom, A., Rhodin, I. N., Lund- Serra, M., Pisu, M. G., Littera, M., Papi, Sieghart, W., and Sperk, G. (2002). Sub- phasic and tonic GABA(A) conduc- berg, M., Olsson, T., and Nyberg, G., Sanna, E., Tuveri, F., Usala, L., unit composition, distribution and tances in hippocampal neurons. J. L. (2005). Impaired cognitive per- Purdy, R. H., and Biggio, G. (2000). function of GABA(A) receptor sub- Neurosci. 22, RC223. formance in patients with chronic Social isolation-induced decreases in types. Curr. Top. Med. Chem. 2, Stoffel-Wagner, B. (2003). Neurosteroid burnout syndrome. Biol. Psychol. 69, both the abundance of neuroac- 795–816. biosynthesis in the human brain and 271–279. tive steroids and GABA(A) receptor Sivilotti, L., and Nistri,A. (1991). GABA its clinical implications. Ann. N. Y. Sanna, E., Murgia, A., Casula, A., and function in rat brain. J. Neurochem. receptor mechanisms in the central Acad. Sci. 1007, 64–78. Biggio, G. (1997). Differential sub- 75, 732–740. nervous system. Prog. Neurobiol. 36, Stromberg, J., Backstrom, T., and Lund- unit dependence of the actions of the Shen, H., Gong, Q. H., Aoki, C., 35–92. gren, P. (2005). Rapid non-genomic general anesthetics alphaxalone and Yuan, M., Ruderman, Y., Dattilo, M., Smart,T. G. (1992). A novel modulatory effect of glucocorticoid metabolites etomidate at gamma-aminobutyric Williams, K., and Smith, S. S. (2007). binding site for zinc on the GABAA and neurosteroids on the gamma- acid type A receptors expressed in Reversal of neurosteroid effects at receptor complex in cultured rat aminobutyric acid-A receptor. Eur. Xenopus laevis oocytes. Mol. Phar- alpha4beta2delta GABAA receptors neurones. J. Physiol. (Lond.) 447, J. Neurosci. 21, 2083–2088. macol. 51, 484–490. triggers anxiety at puberty. Nat. Neu- 587–625. Stromberg, J., Haage, D., Taube, M., Saunders, P. A., Copeland, J. R., Dewey, rosci. 10, 469–477. Smith, M. C., and Riskin, B. J. (1991). Backstrom, T., and Lundgren, P. M. E., Davidson, I. A., McWilliam, Shen, W., Mennerick, S., Covey, D. The clinical use of barbiturates in (2006). Neurosteroid modulation of C., Sharma, V., and Sullivan, C. F., and Zorumski, C. F. (2000). neurological disorders. Drugs 42, allopregnanolone and GABA effect (1991). Heavy drinking as a risk Pregnenolone sulfate modulates 365–378. on the GABA-A receptor. Neuro- factor for depression and dementia inhibitory synaptic transmission Smith, S. S., Gong, Q. H., Hsu, F. C., science 143, 73–81. in elderly men. Findings from the by enhancing GABA(A) receptor Markowitz, R. S., ffrench-Mullen, J. Stromstedt, M., Warner, M., Banner, Liverpool longitudinal community desensitization. J. Neurosci. 20, M., and Li, X. (1998). GABA(A) C. D., MacDonald, P. C., and study. Br. J. Psychiatry 159, 213–216. 3571–3579. receptor alpha4 subunit suppression Gustafsson, J. A. (1993). Role of Saxena, N. C., and Macdonald, R. L. Shen, W., Mennerick, S., Zorumski, E. prevents withdrawal properties of brain in regula- (1994). Assembly of GABAA recep- C., Covey, D. F., and Zorumski, an endogenous steroid. Nature 392, tion of the level of anesthetic steroids tor subunits: role of the delta sub- C. F. (1999). Pregnenolone sulfate 926–930. in the brain. Mol. Pharmacol. 44, unit. J. Neurosci. 14, 7077–7086. and dehydroepiandrosterone sulfate Smith, S. S., Shen, H., Gong, Q. H., and 1077–1083. Saxena, N. C., and Macdonald, R. L. inhibit GABA-gated chloride cur- Zhou, X. (2007). Neurosteroid regu- Sun, C., Sieghart, W., and Kapur, (1996). Properties of putative cere- rents in Xenopus oocytes express- lation of GABA(A) receptors: focus J. (2004). Distribution of alpha1, bellar gamma-aminobutyric acid A ing picrotoxin-insensitive GABA(A) on the alpha4 and delta subunits. alpha4, gamma2, and delta subunits receptor isoforms. Mol. Pharmacol. receptors. Neuropharmacology 38, Pharmacol. Ther. 116, 58–76. of GABAA receptors in hippocam- 49, 567–579. 267–271. Soares, C. N., Cohen, L. S., Otto, M. pal granule cells. Brain Res. 1029, Schofield, P. R., Darlison, M. G., Fujita, Shu, H. J., Eisenman, L. N., Jinadasa, W., and Harlow, B. L. (2001). Char- 207–216. N., Burt, D. R., Stephenson, F. D., Covey, D. F., Zorumski, C. F., acteristics of women with premen- Sunderland, T., Merril, C. R., Harring- A., Rodriguez, H., Rhee, L. M., and Mennerick, S. (2004). Slow strual dysphoric disorder (PMDD) ton, M. G., Lawlor, B. A., Molchan, Ramachandran, J., Reale, V., Glen- actions of neuroactive steroids at who did or did not report history S. E., Martinez, R., and Murphy, D. corse, T. A., Seeburg, P. H., and GABAA receptors. J. Neurosci. 24, of depression: a preliminary report L. (1989). Reduced plasma dehy- Barnard, E. A. (1987). Sequence and 6667–6675. from the Harvard Study of Moods droepiandrosterone concentrations functional expression of the GABA Shumaker, S. A., Legault, C., Kuller, and Cycles. J. Womens Health Gend. in Alzheimer’s disease. Lancet 2, 570. A receptor shows a ligand-gated L., Rapp, S. R., Thal, L., Lane, D. Based Med. 10, 873–878. Sundstro,I. I.,and Backstrom,T.(1999). receptor super-family. Nature 328, S., Fillit, H., Stefanick, M. L., Hen- Solfrizzi, V., D’Introno, A., Colacicco, A. Citalopram increases pregnanolone 221–227. drix, S. L., Lewis, C. E., Masaki, K., M., Capurso, C., Del Parigi, A., Bal- sensitivity in patients with premen- Schumacher, M., Akwa, Y., Guennoun, and Coker, L. H. (2004). Conjugated dassarre, G., Scapicchio, P., Scafato, strual dysphoric disorder. Acta Phys- R., Robert, F., Labombarda, F., equine estrogens and incidence of E., Amodio, M., Capurso, A., and iol. Scand. 167 A6–A7. Desarnaud, F., Robel, P., De Nicola, probable dementia and mild cogni- Panza, F. (2007). Alcohol consump- Sundstrom, I., Andersson, A., Nyberg, A. F., and Baulieu, E. E. (2000). tive impairment in postmenopausal tion, mild cognitive impairment, S., Ashbrook, D., Purdy, R. H., and Steroid synthesis and metabolism women: Women’s Health Initia- and progression to dementia. Neu- Backstrom, T. (1998). Patients with in the nervous system: trophic and tive Memory Study. JAMA 291, rology 68, 1790–1799. premenstrual syndrome have a dif- protective effects. J. Neurocytol. 29, 2947–2958. Sousa, A., and Ticku, M. K. (1997). ferent sensitivity to a neuroactive 307–326. Shumaker, S. A., Legault, C., Rapp, S. Interactions of the neurosteroid steroid during the menstrual cycle Sear, J. W. (1997). ORG 21465, a R., Thal, L., Wallace, R. B., Ock- dehydroepiandrosterone sulfate compared to control subjects. Neu- new water-soluble steroid hyp- ene, J. K., Hendrix, S. L., Jones, B. with the GABA(A) receptor com- roendocrinology 67, 126–138. notic: more of the same or some- N. III, Assaf, A. R., Jackson, R. D., plex reveals that it may act via the Sundstrom, I., Ashbrook, D., and Back- thing different? Br. J. Anaesth. 79, Kotchen, J. M., Wassertheil-Smoller, picrotoxin site. J. Pharmacol. Exp. strom, T. (1997). Reduced ben- 417–419. S., and Wactawski-Wende, J. (2003). Ther. 282, 827–833. zodiazepine sensitivity in patients Semyanov, A., Walker, M. C., and Kull- Estrogen plus progestin and the inci- Sperk, G., Schwarzer, C., Tsunashima, with premenstrual syndrome: a pilot mann, D. M. (2003). GABA uptake dence of dementia and mild cogni- K., Fuchs, K., and Sieghart, W. study. Psychoneuroendocrinology 22, regulates cortical excitability via cell tive impairment in postmenopausal (1997). GABA(A) receptor subunits 25–38. www.frontiersin.org October 2011 | Volume 2 | Article 44 | 21 Wang Neurosteroids and GABAA-receptor function

Sundstrom Poromaa, I., Smith, S., and C., Carter, R. B., and Hawkinson, J. Benz, A., Fu, T., Zorumski, E., Stein- Wieland, S., Belluzzi, J. D., Stein, L., Gulinello, M. (2003). GABA recep- E. (1997). 3 alpha-Hydroxy-3 beta- bach, J. H., Covey, D. F., Zorum- and Lan, N. C. (1995). Compara- tors, progesterone and premenstrual (phenylethynyl)-5 beta-pregnan-20- ski, C. F., and Mennerick, S. (2002). tive behavioral characterization of dysphoric disorder. Arch. Womens ones: synthesis and pharmacological 3beta -hydroxypregnane steroids are the neuroactive steroids 3 alpha- Ment. Health 6, 23–41. activity of neuroactive steroids with pregnenolone sulfate-like GABA(A) OH,5 alpha-pregnan-20-one and 3 Sur, C., Farrar, S. J., Kerby, J., Whiting, high affinity for GABAA receptors. J. receptor antagonists. J. Neurosci. 22, alpha-OH,5 beta-pregnan-20-one in P. J., Atack, J. R., and McKernan, R. Med. Chem. 40, 73–84. 3366–3375. rodents. Psychopharmacology (Berl.) M. (1999). Preferential coassembly Valera, S., Ballivet, M., and Bertrand, Wang, M., Rahman, M., Zhu, D., and 118, 65–71. of alpha4 and delta subunits of the D. (1992). Progesterone modulates Bäckström, T. (2006). Pregnenolone Wieland, S., Lan, N. C., Mirasedeghi, S., gamma-aminobutyric acidA recep- a neuronal nicotinic acetylcholine sulfate and zinc inhibit recombinant and Gee, K. W. (1991). Anxiolytic tor in rat thalamus. Mol. Pharmacol. receptor. Proc. Natl. Acad. Sci. U.S.A. rat GABAA receptor through differ- activity of the progesterone metabo- 56, 110–115. 89, 9949–9953. ent channel property. Acta physiol. lite 5 alpha-pregnan-3 alpha-o1-20- Sveindottir, H., and Backstrom, T. van Rijnsoever, C., Tauber, M., Choulli, (Oxf.)188, 153–163. one. Brain Res. 565, 263–268. (2000). Prevalence of menstrual M. K., Keist, R., Rudolph, U., Mohler, Wang, M., Seippel, L., Purdy, R. Wihlback, A. C., Sundstrom-Poromaa, cycle symptom cyclicity and pre- H., Fritschy, J. M., and Crestani, H., and Backstrom, T. (1996). I., and Backstrom, T. (2006). Action menstrual dysphoric disorder in a F. (2004). Requirement of alpha5- Relationship between symptom by and sensitivity to neuroactive random sample of women using and GABAA receptors for the develop- severity and steroid variation steroids in menstrual cycle related not using oral contraceptives. Acta ment of tolerance to the sedative in women with premenstrual CNS disorders. Psychopharmacology Obstet. Gynecol. Scand. 79, 405–413. action of diazepam in mice. J. Neu- syndrome: study on serum preg- (Berl.) 186, 388–401. Szabadics,J.,Varga,C.,Molnar,G.,Olah, rosci. 24, 6785–6790. nenolone, pregnenolone sulfate, 5 Wohlfarth, K. M., Bianchi, M. T., S., Barzo, P., and Tamas, G. (2006). Vanover, K. E., Rosenzweig-Lipson, alpha-pregnane-3,20-dione and 3 and Macdonald, R. L. (2002). Excitatory effect of GABAergic axo- S., Hawkinson, J. E., Lan, N. C., alpha-hydroxy-5 alpha-pregnan-20- Enhanced neurosteroid potentiation axonic cells in cortical microcircuits. Belluzzi, J. D., Stein, L., Barrett, J. E., one. J. Clin. Endocrinol. Metab. 81, of ternary GABA(A) receptors con- Science 311, 233–235. Wood,P.L., and Carter, R. B. (2000). 1076–1082. taining the delta subunit. J. Neurosci. Thompson, S. A., Bonnert, T. P., Characterization of the anxiolytic Wang, M. D., Backstrom, T., and Land- 22, 1541–1549. Cagetti, E., Whiting, P. J., and properties of a novel neuroactive gren, S. (2000). The inhibitory Wolkowitz, O. M., Reus, V. I., Roberts, Wafford, K. A. (2002). Overex- steroid, Co 2-6749 (GMA-839; effects of allopregnanolone and E., Manfredi, F., Chan, T., Ormis- pression of the GABA(A) receptor WAY-141839; 3alpha, 21-dihydroxy- pregnanolone on the population ton, S., Johnson, R., Canick, J., epsilon subunit results in insensitiv- 3beta-trifluoromethyl-19-nor- spike, evoked in the rat hippocam- Brizendine, L., and Weingartner, ity to anaesthetics. Neuropharmacol- 5beta-pregnan-20-one), a selective pal CA1 stratum pyramidale in vitro, H. (1995). and ogy 43, 662–668. modulator of gamma-aminobutyric can be blocked selectively by epial- cognition-enhancing effects of Thompson, S. A., Whiting, P. J., and acid(A) receptors. J. Pharmacol. lopregnanolone. Acta Physiol. Scand. DHEA in major depression. Ann. N. Wafford, K. A. (1996). Barbiturate Exp. Ther. 295, 337–345. 169, 333–341. Y. Acad. Sci. 774, 337–339. interactions at the human GABAA Verdoorn, T. A., Draguhn, A., Ymer, Wang, M. D., Rahman, M., Zhu, D., Woodhull, A. M. (1973). Ionic blockage receptor: dependence on receptor S., Seeburg, P. H., and Sakmann, Johansson, I. M., and Backstrom, of sodium channels in nerve. J. Gen. subunit combination. Br. J. Pharma- B. (1990). Functional properties of T. (2007). 3Beta-hydroxysteroids Physiol. 61, 687–708. col. 117, 521–527. recombinant rat GABAA receptors and pregnenolone sulfate inhibit Woodward, R. M., Polenzani, L., Timby, E., Balgard, M., Nyberg, depend upon subunit composition. recombinant rat GABA(A) receptor and Miledi, R. (1992). Effects of S., Spigset, O., Andersson, A., Neuron 4, 919–928. through different channel property. steroids on gamma-aminobutyric Porankiewicz-Asplund, J., Purdy, Vicini, S., Losi, G., and Homanics, Eur. J. Pharmacol. 557, 124–131. acid receptors expressed in Xeno- R. H., Zhu, D., Backstrom, T., and G. E. (2002). GABA(A) recep- Wardell, B., Marik, P. S., Piper, D., pus oocytes by poly(A)+ RNA from Poromaa, I. S. (2006). Pharma- tor delta subunit deletion pre- Rutar, T., Jorgensen, E. M., and mammalian brain and retina. Mol. cokinetic and behavioral effects vents neurosteroid modulation of Bamber, B. A. (2006). Residues in Pharmacol. 41, 89–103. of allopregnanolone in healthy inhibitory synaptic currents in cere- the first transmembrane domain Wu, F. S., Gibbs, T. T., and Farb, D. H. women. Psychopharmacology (Berl.) bellar neurons. Neuropharmacology of the Caenorhabditis elegans (1991). Pregnenolone sulfate: a pos- 186, 414–424. 43, 646–650. GABA(A) receptor confer sensitivity itive allosteric modulator at the N- Tossman, U., Jonsson, G., and Unger- Vincze,G.,Almos,P.,Boda,K.,Dome,P., to the neurosteroid pregnenolone methyl-D-aspartate receptor. Mol. stedt, U. (1986). Regional distribu- Bodi, N., Szlavik, G., Magloczki, E., sulfate. Br. J. Pharmacol. 148, Pharmacol. 40, 333–336. tion and extracellular levels of amino Pakaski, M., Janka, Z., and Kalman, 162–172. Wu, Y., Wang, W., and Richerson, G. acids in rat central nervous system. J. (2007). Risk factors of cognitive Weinbroum, A. A., Szold, O., Ogorek, B. (2003). induces tonic Acta Physiol. Scand. 127, 533–545. decline in residential care in Hun- D., and Flaishon, R. (2001). The inhibition via GABA transporter Turkmen, S., Lofgren, M., Birzniece, V., gary. Int. J. Geriatr. Psychiatry 22, -induced paradox phe- reversal without increasing vesicu- Backstrom, T., and Johansson, I. M. 1208–1216. nomenon is reversible by flumaze- lar GABA release. J. Neurophysiol. 89, (2006). Tolerance development to Wallner, M., Hanchar, H. J., and Olsen, nil. Epidemiology, patient character- 2021–2034. Morris water maze test impairments R. W. (2003). Ethanol enhances istics and review of the literature. Yamada, J., Okabe, A., Toyoda, H., Kilb, induced by acute allopregnanolone. alpha 4 beta 3 delta and alpha 6 Eur. J. Anaesthesiol. 18, 789–797. W., Luhmann, H. J., and Fukuda, A. Neuroscience 139, 651–659. beta 3 delta gamma-aminobutyric Welberg, L. A., and Seckl, J. R. (2001). (2004). Cl- uptake promoting depo- Turkmen, S., Lundgren, P., Birzniece, acid type A receptors at low con- Prenatal stress, glucocorticoids and larizing GABA actions in immature V., Zingmark, E., Backstrom, T., centrations known to affect humans. the programming of the brain. J. rat neocortical neurones is mediated and Johansson, I. M. (2004). 3beta- Proc. Natl. Acad. Sci. U.S.A. 100, Neuroendocrinol. 13, 113–128. by NKCC1. J. Physiol. (Lond.) 557, 20beta-dihydroxy-5alpha-pregnane 15218–15223. Whiting, P.J. (2003a). GABA-A receptor 829–841. (UC1011). Antagonism of the Wang, M., Backstrom, T., Sundstrom, I., subtypes in the brain: a paradigm for Yang, L., Omori, K., Otani, H., GABA potentiation and the learning Wahlstrom, G., Olsson, T., Zhu, D., CNS drug discovery? Drug Discov. Suzukawa, J., and Inagaki, C. (2003). impairment induced in rats by Johansson, I. M., Bjorn, I., and Bixo, Today 8, 445–450. GABAC receptor agonist suppressed allopregnanolone. Eur. J. Neurosci. M. (2001). Neuroactive steroids and Whiting, P. J. (2003b). The GABAA ammonia-induced apoptosis in 20, 1604–1612. central nervous system disorders. receptor gene family: new oppor- cultured rat hippocampal neu- Upasani, R. B., Yang, K. C., Acosta- Int. Rev. Neurobiol. 46, 421–459. tunities for drug development. rons by restoring phosphorylated Burruel, M., Konkoy, C. S., McLel- Wang, M., He, Y., Eisenman, L. N., Curr. Opin. Drug Discov. Devel. 6, BAD level. J. Neurochem. 87, lan, J. A., Woodward, R. M., Lan, N. Fields, C., Zeng, C. M., Mathews, J., 648–657. 791–800.

Frontiers in Endocrinology | Neuroendocrine Science October 2011 | Volume 2 | Article 44 | 22 Wang Neurosteroids and GABAA-receptor function

Yee, B. K., Hauser, J., Dolgov, V. V., the GABA receptor subunits J. Pharmacol. Exp. Ther. 275, Received: 05 July 2011; paper pending Keist, R., Mohler, H., Rudolph, mRNA levels in mammalian cul- 784–789. published: 25 July 2011; accepted: 14 U., and Feldon, J. (2004). GABA tured cortical neurons following Zhu, W. J., and Vicini, S. (1997). Neu- September 2011; published online: 04 receptors containing the alpha5 chronic neurosteroid treatment. rosteroid prolongs GABAA channel October 2011. subunit mediate the trace effect Brain Res. Mol. Brain Res. 41, deactivation by altering kinetics of Citation: Wang M (2011) Neu- in aversive and appetitive condi- 163–168. desensitized states. J. Neurosci. 17, rosteroids and GABA-A receptor tioning and extinction of condi- Yu, R., and Ticku, M. K. (1995a). Effects 4022–4031. function. Front. Endocrin. 2:44. doi: tioned fear. Eur. J. Neurosci. 20, of chronic pentobarbital treatment Zorumski, C. F., Mennerick, S., Isen- 10.3389/fendo.2011.00044 1928–1936. on the GABAA receptor com- berg, K. E., and Covey, D. F. (2000). This article was submitted to Frontiers Yeung, J. Y., Canning, K. J., Zhu, G., plex in mammalian cortical neu- Potential clinical uses of neuroactive in Neuroendocrine Science, a specialty of Pennefather, P., MacDonald, J. F., rons. J. Pharmacol. Exp. Ther. 275, steroids. Curr. Opin. Investig. Drugs Frontiers in Endocrinology. and Orser, B. A. (2003). Tonically 1442–1446. 1, 360–369. Copyright © 2011 Wang. This is an open- activated GABAA receptors in hip- Yu, R., and Ticku, M. K. (1995b). access article subject to a non-exclusive pocampal neurons are high-affinity, Chronic neurosteroid treatment Conflict of Interest Statement: The license between the authors and Frontiers low-conductance sensors for extra- decreases the efficacy of benzodi- author declares that the research was Media SA, which permits use, distribu- cellular GABA. Mol. Pharmacol. 63, azepine ligands and neurosteroids conducted in the absence of any tion and reproduction in other forums, 2–8. at the gamma-aminobutyric commercial or financial relationships provided the original authors and source Yu, R., Follesa, P., and Ticku, M. acidA receptor complex in that could be construed as a potential are credited and other Frontiers condi- K. (1996). Down-regulation of mammalian cortical neurons. conflict of interest. tions are complied with.

www.frontiersin.org October 2011 | Volume 2 | Article 44 | 23