Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins

Journal of Thermodynamics & Catalysis, 2012
...Read more
Volume 3 • Issue 4 • 1000116 J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal Research Article Open Access Arias, J Thermodynam Cat 2012, 3:4 DOI: 10.4172/2157-7544.1000116 Review Article Open Access Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins Hugo R. Arias* Department of Medical Education, College of Medicine, California Northstate University, CA, USA *Corresponding author: Hugo R. Arias, Department of Medical Education, College of Medicine, California Northstate University, USA, Tel: 916-686-7300; Fax: 916-686- 7310; E-mail: hugo.arias@cnucom.org Received October 12, 2012; Accepted October 18, 2012; Published October 22, 2012 Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Copyright: © 2012 Arias HR. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. Keywords: Cys-loop ligand-gated ion channels; Nicotinic acetylcholine receptors; Acetylcholine binding proteins; Loop C; Crystallography Abbreviations: AChR: Nicotinic acetylcholine receptor; AChBP: Acetylcholine bound protein; 5-HT: 5-Hydroxytryptamine (serotonin); ACh: Acetylcholine; GABA: γ-aminobutyric acid; DMXBA: 3-(2,4 dimethoxybenzylidene)-anabaseine; α-BTx: α-bungarotoxin; κ-BTx: κ-bungarotoxin; α-CbTx: α-cobrotoxin; MLA: Methyllycaconitine; d-TC: d-tubocurarine; DHβE: Dihydro-β-erythroidine; SPX: 13-desmethyl spirolide C; GYM: Gymnodimine A; α-CTx: α-conotoxin; MTSET + : 2-(trimethylammonium) ethylmethanethiosulfonate; MMTS: methylmethanethiosulfonate; Ls: Lymnaeastagnalis; Ac: Aplysia californica; Bt: Bulinus truncates; Bg: Biomphalaria glabrata; Ct: Capitella teleta; GLIC: Gloebacter violaceus; ELIC: Erwinia chrysanthemi Introduction A vast amount of evidence indicates that nicotinic acetylcholine receptors (AChRs) are important for the homeostasis and function of our body. AChRs are cation channels members of the Cys-loop ligand gated ion channel superfamily, including type 3 serotonin (5-hydroxytryptamine; 5-HT) cation channels as well as anion channels such as type A and C γ-aminobutyric acid (GABA) and glycine receptors [1-3]. A large number of subunits have been cloned for all members of the Cys-loop super family from vertebrates and invertebrates (Ligand-gated ion channel database, http://www.ebi. ac.uk/compneur-srv/LGICdb/cys-loop.php) [4]. In vertebrates, AChR subunits are classifed in two types, α (i.e., α1-α10) and non-α (i.e., β1- β4, γ, δ, and ε), where α subunits contain a disulphide bridge in the binding site, whereas non-α subunits do not. AChRs can be formed by the same subunit comprising homomeric receptors [e.g., α7, α8 (expressed only in chicks), and α9] or by diferent subunits comprising heteromeric receptors (e.g., α4β2, α3β4) [5,6]. However, not all possible subunit arrangements are functional. For example, the neuronal AChR subunits α2-α4, β2, and β4 can co-assemble in pair wise combinations forming functional AChRs (e.g., α4β2, α3β4). Although the α9 subunit is actually expressed in the cochlea, the functional receptor in hear cells is the α9α10 AChR subtype. Although the homomeric α7 AChR is functional, α7 AChRs containing other subunits (e.g., α7β2) have been characterized endogenously and probably these are the native AChRs. Each α5, α6, and β3 subunit does not form functional AChRs by binary combinations with another subunit, instead they prefer ternary combinations [e.g., (α4)2α5(β2) 2 ]. AChRs can even be formed by assembling four diferent subunits [e.g., α4α6α5(β2) 2 ], but there is no evidence of a receptor subtype formed by fve diferent subunits. In addition, two diferent stoichiometries of the α4β2 AChR (and probably for the α3β4 AChR as well), (α4) 3 (β2) 2 and (α4) 2 (β2) 3 , have been found in heterologous [7-10] and endogenous [11] cells. Te (α4) 2 (β2) 3 stoichiometry is more sensitive to the action of agonists including agonist induced up-regulation, desensitizes less rapidly, and has lower Ca 2+ permeability compared to that for the (α4) 3 (β2) 2 stoichiometry. Te use of subunit concatemers might be useful in determining the subunit composition and stoichiometry of native AChRs [5]. Vertebrate AChRs are expressed in neuronal and non-neuronal tissues. In the peripheral and central nervous systems, presynaptic AChRs modulate the release of several neurotransmitters, including ACh, 5-HT, GABA, dopamine, norepinephrine, and glutamate, whereas postsynaptic AChRs mediate rapid transmission by converting a chemical signal into membrane depolarization (i.e., electrical signal), or trigger cytoplasmic cascades. Tese combined actions modulate several important functions in our body, including cognition, memory, pain perception, auditory response, and muscle contraction. In non- neuronal tissues, AChRs are involved in angiogenesis and immune responses [12,13]. Since AChRs are involved in such important physiological functions, their improper activities (e.g., decreased number, mutations, or hypo/hyperactivity) can produce several diseases, Abstract Nicotinic acetylcholine receptors (AChRs) are the best characterized ion channels representing the Cys-loop ligand-gated ion channel superfamily. Studies using Torpedo AChRs in the closed and open states and acetylcholine binding proteins (AChBPs) from different origins have elucidated the most important structural and functional features of the agonist/competitive antagonist binding sites. The frst step in recognizing the neurotransmitter ACh and other agonists is fundamental in the process of agonist-induced activation, including the opening of the intrinsic cation channel. The AChBP studies demonstrated that Loop C is an important structural feature that is modifed by ligand binding. These studies defned important pharmacologic features of AChR ligands, including the differences between full and partial agonists, agonists and competitive antagonists, peptidic and non-peptidic ligands, and between high affnity and high selectivity. The studies showing the structural mechanisms by which specifc ligands can activate, inhibit, and potentiate different AChR subtypes could be of therapeutic importance. Journal of Thermodynamics & Catalysis J o u r n a l o f T h e r m o d y n a m i c s & C a t a l y s i s ISSN: 2157-7544
Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Page 2 of 7 Volume 3 • Issue 4 • 1000116 J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal including Alzheimer’s disease, Parkinson’s disease, schizophrenia, nocturnal frontal lobeepilepsy, attention defcit hyperactivity disorder, Tourette’s syndrome, drug and nicotine addictions, depression and anxiety, myasthenia gravis, myasthenic syndromes, tumor growth, and decreased immune response [2,3,12-15]. Several AChR subunits have been also cloned in insects and invertebrates [16]. In invertebrates, AChRs have become key elements for the development of neuroactive pesticides. For example, several neonicotinoids (e.g., imidacloprid, clothianidin, and thiacloprid) have insecticide activities but extremely low mammalian toxicities due to their high afnity and specifcity for insect AChRs [16]. Breakthroughs in the Study of AChRs Tree main breakthroughs in the last 10 years have helped in elucidating the functionally relevant structural features of AChRs and their cousins [1,17,18]: (1) the Torpedo AChR structures in the closed [19] and open [20] states, showing the main features at the extracellular, transmembrane, and intracellular domains (Figure 1), (2) the crystal structures of several acetylcholine binding proteins (AChBPs) showing details of the binding sites for agonists and competitive antagonists [21], and (3) the recently elucidated prokaryotic cation channels showing subtle diferences between the open and closed conformations, as well as between activated (i.e., with several primary amines [22]), blocked (i.e., with several cations and open-channel blockers [23]), and inhibited (i.e., with several allosteric modulators [24]) states. Table 1 summarizes the diferent three dimensional structures obtained so far for Torpedo AChRs and diferent AChBPs bound to a variety of ligands. Structurally, AChRs are pentameric proteins with an extracellular domain that carries the binding sites for agonists (e.g., the neurotransmitter ACh and nicotine) and competitive antagonists [e.g., methyllycaconitine (MLA)]. In addition to these orthosteric sites, several additional sites have been characterized in this domain for allosteric modulators [25,26]. Te transmembrane domain, specifcally the M2 transmembrane segments from each subunit, forms the ion channel that is essential for cation fux (i.e., Na + and Ca 2+ infux and K + efux), fnally producing membrane depolarization, and in the case of Ca 2+ , triggering diferent intracellular pathways. Te ion channel is the maindomain for the interaction with a very broad group of compounds called noncompetitive antagonists [14,15,27]. In addition, several binding sites for negative allosteric modulators have been characterized between the four M1-M4 transmembrane segments [25,26]. Agonist binding at the main domain domain triggers the opening of the ion channel, an intrinsic process called gating [1]. Determining how the agonist-induced structural changes, that start in the agonist binding pocket, are propagated through a distance of ~50 Å to the gate is central for the understanding of the receptor function. One of the AChR domains that have recently attracted attention among researchers is the extracellular–transmembrane interface. Tis is a very unique transitional zone where β-sheets from the extracellular domain merge with α-helices from the transmembrane domain, fnally allowing functional communication between both domains [1,25,26]. Overall Structure of AChRs AChRs are pentameric proteins where the subunits are arranged around an axis perpendicular to the membrane, and each subunit shares a basic scafold composed of: (1) a large N-terminal extracellular domain of ~200 amino acids; (2) three transmembrane domains (M1- M3) separated by short loops; (3) a cytoplasmic loop of variable size and amino acid sequence; and (4) a fourth transmembrane domain (M4) with a relatively short and variable extracellular COOH-terminal sequence (Figure 1). AChRs are therefore built on a modular basis, with the extracellular domain containing the agonist binding sites, the transmembrane domain containing the pore, selectivity flter, and channel gate, and the cytoplasmic domain performing additional modulatory activities [19,20]. Recent structural studies have provided details of the three dimensional structure of AChRs and consequently for other members of this receptor superfamily. In particular, the structural model of the Torpedo AChR at 4 Å resolution [19] has revealed important information and has been invaluable in the interpretation of functional and pharmacological data. Although no structural information is available for any Cys-loop receptor at the atomic resolution level, the extracellular domain of the AChR α1 subunit has been resolved at 1.94 Å [28]. Additional high resolution structural information has become available from studies of proteins which show close structural similarity to AChRs, including soluble AChBPs from a variety of animals from the Mollusca [21,29] and Annelida [30] phyla [1,17], as well as prokaryotic proton-gated ion channels from the bacteria Erwinia chrysanthemi (ELIC) and Gloebacter violaceus (GLIC), respectively [1,18]. Te characterization of GLIC showed that it forms a cation-selective channel that is activated by protons, where currents do not decay during activation, suggesting no or very slow desensitization [31]. Recent results using ELIC showed that diferent primary amines, including GABA, can also activate this channel [22]. Te ELIC X-ray structure shows 16% sequence identity to αAChR subunits. In general, the extracellular domain is very similar to its eukaryotic counterpart and to AChBPs, but lacks the N-terminal α-helix. However, the putative binding site and several of the aromatic residues found in AChRs are conserved. Te central part of the Cys-loop is also conserved but lacks the fanking disulfde-bridge. Figure 1: Torpedo AChR model showing the structural differences between the open (magenta) and closed (blue) states at 6.2 Å resolution (modifed from [20] ). Only subtle differences can be seen between both conformational states. AChRs are formed by three main domains: The extracellular domain shows the β-strands forming the β-sandwich structure.The transmembrane domain is formed by 20 α-helices, four α-helices per subunit. The cytoplasmic domain is the smallest domain.
Journal of Thermodynamics & Catalysis s atalysis &C urnal of Th Jo dynami mo c er ISSN: 2157-7544 Research Article Review Article Arias, J Thermodynam Cat 2012, 3:4 DOI: 10.4172/2157-7544.1000116 Open OpenAccess Access Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins Hugo R. Arias* Department of Medical Education, College of Medicine, California Northstate University, CA, USA Abstract Nicotinic acetylcholine receptors (AChRs) are the best characterized ion channels representing the Cys-loop ligand-gated ion channel superfamily. Studies using Torpedo AChRs in the closed and open states and acetylcholine binding proteins (AChBPs) from different origins have elucidated the most important structural and functional features of the agonist/competitive antagonist binding sites. The first step in recognizing the neurotransmitter ACh and other agonists is fundamental in the process of agonist-induced activation, including the opening of the intrinsic cation channel. The AChBP studies demonstrated that Loop C is an important structural feature that is modified by ligand binding. These studies defined important pharmacologic features of AChR ligands, including the differences between full and partial agonists, agonists and competitive antagonists, peptidic and non-peptidic ligands, and between high affinity and high selectivity. The studies showing the structural mechanisms by which specific ligands can activate, inhibit, and potentiate different AChR subtypes could be of therapeutic importance. Keywords: Cys-loop ligand-gated ion channels; Nicotinic acetylcholine receptors; Acetylcholine binding proteins; Loop C; Crystallography Abbreviations: AChR: Nicotinic acetylcholine receptor; AChBP: Acetylcholine bound protein; 5-HT: 5-Hydroxytryptamine (serotonin); ACh: Acetylcholine; GABA: γ-aminobutyric acid; DMXBA: 3-(2,4 dimethoxybenzylidene)-anabaseine; α-BTx: α-bungarotoxin; κ-BTx: κ-bungarotoxin; α-CbTx: α-cobrotoxin; MLA: Methyllycaconitine; d-TC: d-tubocurarine; DHβE: Dihydro-β-erythroidine; SPX: 13-desmethyl spirolide C; GYM: Gymnodimine A; α-CTx: α-conotoxin; MTSET+: 2-(trimethylammonium) ethylmethanethiosulfonate; MMTS: methylmethanethiosulfonate; Ls: Lymnaeastagnalis; Ac: Aplysia californica; Bt: Bulinus truncates; Bg: Biomphalaria glabrata; Ct: Capitella teleta; GLIC: Gloebacter violaceus; ELIC: Erwinia chrysanthemi Introduction A vast amount of evidence indicates that nicotinic acetylcholine receptors (AChRs) are important for the homeostasis and function of our body. AChRs are cation channels members of the Cys-loop ligand gated ion channel superfamily, including type 3 serotonin (5-hydroxytryptamine; 5-HT) cation channels as well as anion channels such as type A and C γ-aminobutyric acid (GABA) and glycine receptors [1-3]. A large number of subunits have been cloned for all members of the Cys-loop super family from vertebrates and invertebrates (Ligand-gated ion channel database, http://www.ebi. ac.uk/compneur-srv/LGICdb/cys-loop.php) [4]. In vertebrates, AChR subunits are classified in two types, α (i.e., α1-α10) and non-α (i.e., β1β4, γ, δ, and ε), where α subunits contain a disulphide bridge in the binding site, whereas non-α subunits do not. AChRs can be formed by the same subunit comprising homomeric receptors [e.g., α7, α8 (expressed only in chicks), and α9] or by different subunits comprising heteromeric receptors (e.g., α4β2, α3β4) [5,6]. However, not all possible subunit arrangements are functional. For example, the neuronal AChR subunits α2-α4, β2, and β4 can co-assemble in pair wise combinations forming functional AChRs (e.g., α4β2, α3β4). Although the α9 subunit is actually expressed in the cochlea, the functional receptor in hear cells is the α9α10 AChR subtype. Although the homomeric α7 AChR is functional, α7 AChRs containing other subunits (e.g., α7β2) have J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal been characterized endogenously and probably these are the native AChRs. Each α5, α6, and β3 subunit does not form functional AChRs by binary combinations with another subunit, instead they prefer ternary combinations [e.g., (α4)2α5(β2)2]. AChRs can even be formed by assembling four different subunits [e.g., α4α6α5(β2)2], but there is no evidence of a receptor subtype formed by five different subunits. In addition, two different stoichiometries of the α4β2 AChR (and probably for the α3β4 AChR as well), (α4)3(β2)2 and (α4)2(β2)3, have been found in heterologous [7-10] and endogenous [11] cells. The (α4)2(β2)3 stoichiometry is more sensitive to the action of agonists including agonist induced up-regulation, desensitizes less rapidly, and has lower Ca2+ permeability compared to that for the (α4)3(β2)2 stoichiometry. The use of subunit concatemers might be useful in determining the subunit composition and stoichiometry of native AChRs [5]. Vertebrate AChRs are expressed in neuronal and non-neuronal tissues. In the peripheral and central nervous systems, presynaptic AChRs modulate the release of several neurotransmitters, including ACh, 5-HT, GABA, dopamine, norepinephrine, and glutamate, whereas postsynaptic AChRs mediate rapid transmission by converting a chemical signal into membrane depolarization (i.e., electrical signal), or trigger cytoplasmic cascades. These combined actions modulate several important functions in our body, including cognition, memory, pain perception, auditory response, and muscle contraction. In nonneuronal tissues, AChRs are involved in angiogenesis and immune responses [12,13]. Since AChRs are involved in such important physiological functions, their improper activities (e.g., decreased number, mutations, or hypo/hyperactivity) can produce several diseases, *Corresponding author: Hugo R. Arias, Department of Medical Education, College of Medicine, California Northstate University, USA, Tel: 916-686-7300; Fax: 916-6867310; E-mail: hugo.arias@cnucom.org Received October 12, 2012; Accepted October 18, 2012; Published October 22, 2012 Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Copyright: © 2012 Arias HR. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. Volume 3 • Issue 4 • 1000116 Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Page 2 of 7 including Alzheimer’s disease, Parkinson’s disease, schizophrenia, nocturnal frontal lobeepilepsy, attention deficit hyperactivity disorder, Tourette’s syndrome, drug and nicotine addictions, depression and anxiety, myasthenia gravis, myasthenic syndromes, tumor growth, and decreased immune response [2,3,12-15]. Several AChR subunits have been also cloned in insects and invertebrates [16]. In invertebrates, AChRs have become key elements for the development of neuroactive pesticides. For example, several neonicotinoids (e.g., imidacloprid, clothianidin, and thiacloprid) have insecticide activities but extremely low mammalian toxicities due to their high affinity and specificity for insect AChRs [16]. Breakthroughs in the Study of AChRs Three main breakthroughs in the last 10 years have helped in elucidating the functionally relevant structural features of AChRs and their cousins [1,17,18]: (1) the Torpedo AChR structures in the closed [19] and open [20] states, showing the main features at the extracellular, transmembrane, and intracellular domains (Figure 1), (2) the crystal structures of several acetylcholine binding proteins (AChBPs) showing details of the binding sites for agonists and competitive antagonists [21], and (3) the recently elucidated prokaryotic cation channels showing subtle differences between the open and closed conformations, as well as between activated (i.e., with several primary amines [22]), blocked (i.e., with several cations and open-channel blockers [23]), and inhibited (i.e., with several allosteric modulators [24]) states. Table 1 summarizes the different three dimensional structures obtained so far for Torpedo AChRs and different AChBPs bound to a variety of ligands. Structurally, AChRs are pentameric proteins with an extracellular domain that carries the binding sites for agonists (e.g., the neurotransmitter ACh and nicotine) and competitive antagonists [e.g., methyllycaconitine (MLA)]. In addition to these orthosteric sites, several additional sites have been characterized in this domain for allosteric modulators [25,26]. The transmembrane domain, specifically the M2 transmembrane segments from each subunit, forms the ion channel that is essential for cation flux (i.e., Na+ and Ca2+influx and K+ efflux), finally producing membrane depolarization, and in the case of Ca2+, triggering different intracellular pathways. M3) separated by short loops; (3) a cytoplasmic loop of variable size and amino acid sequence; and (4) a fourth transmembrane domain (M4) with a relatively short and variable extracellular COOH-terminal sequence (Figure 1). AChRs are therefore built on a modular basis, with the extracellular domain containing the agonist binding sites, the transmembrane domain containing the pore, selectivity filter, and channel gate, and the cytoplasmic domain performing additional modulatory activities [19,20]. Recent structural studies have provided details of the three dimensional structure of AChRs and consequently for other members of this receptor superfamily. In particular, the structural model of the Torpedo AChR at 4 Å resolution [19] has revealed important information and has been invaluable in the interpretation of functional and pharmacological data. Although no structural information is available for any Cys-loop receptor at the atomic resolution level, the extracellular domain of the AChR α1 subunit has been resolved at 1.94 Å [28]. Additional high resolution structural information has become available from studies of proteins which show close structural similarity to AChRs, including soluble AChBPs from a variety of animals from the Mollusca [21,29] and Annelida [30] phyla [1,17], as well as prokaryotic proton-gated ion channels from the bacteria Erwinia chrysanthemi (ELIC) and Gloebacter violaceus (GLIC), respectively [1,18]. The characterization of GLIC showed that it forms a cation-selective channel that is activated by protons, where currents do not decay during activation, suggesting no or very slow desensitization [31]. Recent results using ELIC showed that different primary amines, including GABA, can also activate this channel [22]. The ELIC X-ray structure shows 16% sequence identity to αAChR subunits. In general, the extracellular domain is very similar to its eukaryotic counterpart and to AChBPs, but lacks the N-terminal α-helix. However, the putative binding site and several of the aromatic residues found in AChRs are conserved. The central part of the Cys-loop is also conserved but lacks the flanking disulfide-bridge. The ion channel is the maindomain for the interaction with a very broad group of compounds called noncompetitive antagonists [14,15,27]. In addition, several binding sites for negative allosteric modulators have been characterized between the four M1-M4 transmembrane segments [25,26]. Agonist binding at the main domain domain triggers the opening of the ion channel, an intrinsic process called gating [1]. Determining how the agonist-induced structural changes, that start in the agonist binding pocket, are propagated through a distance of ~50 Å to the gate is central for the understanding of the receptor function. One of the AChR domains that have recently attracted attention among researchers is the extracellular–transmembrane interface. This is a very unique transitional zone where β-sheets from the extracellular domain merge with α-helices from the transmembrane domain, finally allowing functional communication between both domains [1,25,26]. Overall Structure of AChRs AChRs are pentameric proteins where the subunits are arranged around an axis perpendicular to the membrane, and each subunit shares a basic scaffold composed of: (1) a large N-terminal extracellular domain of ~200 amino acids; (2) three transmembrane domains (M1J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal Figure 1: Torpedo AChR model showing the structural differences between the open (magenta) and closed (blue) states at 6.2 Å resolution (modified from [20] ). Only subtle differences can be seen between both conformational states. AChRs are formed by three main domains: The extracellular domain shows the β-strands forming the β-sandwich structure.The transmembrane domain is formed by 20 α-helices, four α-helices per subunit. The cytoplasmic domain is the smallest domain. Volume 3 • Issue 4 • 1000116 Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Page 3 of 7 The Extracellular Domain The solution of the high-resolution structure of the AChBP from Lymnaea stagnalis (Ls-AChBP) was a giant step forward for our knowledge of the structure of the extracellular domain of AChRs [1,17,21]. Since then, several other AChBPs from mollusks and annelids have been characterized (Table 1). In general, AChBPs lack the transmembrane region but contain many of the structural features that give AChRs their unique signature and have therefore become functional and structural surrogates of the extracellular domain of the Cys-loop receptor superfamily. The AChR extracellular domain contains 210 amino acids and shares ∼15-24% sequence identity to aligned sequences of the amino-terminal, extracellular halves of Cys-loop receptor subunits. Each AChBP monomer consists of an N-terminal α-helix, two short 310 helices, and a core of 10 β-strands that form a β-sandwich structure. The inner β-sheet is formed by the β1, β2, β3, β5, β6 and β8 strands, and the outer β-sheet by the β4, β7, β9 and β10 strands. The N- and C- terminals are located at the top and bottom of the pentamer, respectively. In Cys-loop receptors, the end of β10 connects to the start of M1. Located at the bottom of the subunit; the linker between β6 and β7 strands is the signature Cys-loop found in all members of the superfamily, including bacterial ion channels. Agonist binding sites are located at the subunit interfaces [13,17,18]. Each binding site isformed by two faces. One face, called the principal or “positive” face at the α-subunit, is formed by β-strands connected by three loops harboring key aromatic residues [i.e., Loop A (β4β5 loop), Loop B (β7β8 loop), and Loop C (β9β10 loop)]. The complementary or “negative” face at the non-α-subunit contributes with three β-strands clustered in segments by Loops D-F. Thus, key residues (corresponding to Torpedo α1-subunit) from the principal face come from Loop A (Trp86 and Tyr93), Loop B (Trp149 and Gly153) and Loop C (Tyr190, Cys192, Cys193 and Tyr198). The complementary face is formed by residues from Loop D (Trp55 and Asp57), Loop E (Leu109, Arg111, Thr117 and Leu119), and Loop F (Asp174 and Glu176) (residues from the Torpedo δ- or γ-subunit) [1,21,32]. β-sheet, causing tilting of the Cys-loop (Loop C) away from the fivefold axis [20]. The results using AChBPs with different ligands indicate that Loop C from the principal face is in an extended (“open”) conformation in the resting AChR (no agonist), whereas in the presence of full agonists, Loop C is contracted (“closed”) and caps the entrance to the binding cavity, trapping the agonist [36,37]. Cysteine substitution and subsequent oxidation studies on Loop C of muscle AChRs indicate that Loop C capping is involved in the transition of the closed receptor to an activated pre-open intermediate state [38]. The interaction of several agonists (full and partial), competitive antagonists, and allosteric modulators with their binding sites was studied in minute detail by co-crystallization of various ligands with AChBPs from different species (Table 1). In addition to structural differences between ligands, diverse AChBP selectivity for several ligands was observed [17,39]. The structural basis to distinguish high affinity vs high selectivity for different ligands [e.g., d-tubocurarine (d-TC) and strychnine] was also described [40]. Another important structural difference is the interaction between full and partial agonists. In general, AChBP-agonist complexes show a fully contracted (closed) state, whereas AChBP-antagonist complexes show a more extended (open) state. However, additional experiments comparing the full agonist nicotine and the partial agonists cytisine and varenicline could not discriminate any variation in the Loop C closure [41]. In particular, varenicline interacts with highly conserved aromatic amino acids at the principal face of the binding site, and with less conserved hydrophobic residues at the complementary face (e.g., at Loop E) [30]. Interestingly, dihydro-β-erythroidine (DHβE), a potent competitive antagonist of β2containing AChRs, imposes closure of the Loop C as agonists do, but also induces a structural change perpendicular to the observed Loop C movements [42]. To illustrate some important differences between agonists and competitive antagonists, AChBP structures complexed with the partial agonists varenicline and lobeline (Figure 2) and the competitive antagonist DHβE (Figure 3) are shown. The ancestral Cys-loop receptor was likely homomeric and contained five identical binding sites, similarly to present day homomeric receptors, such as α7 and 5-HT3A receptors [32,33]. Evolution led to the appearance of new subunits which lost the ability to form agonist binding sites, giving rise to heteromeric receptors with fewer than five binding sites. The prototypic heteromeric receptors, muscle AChR and GABAARs, contain only two agonist binding sites, which have to be both occupied to allow appropriate gating. Although homomeric receptors contain five identical binding sites, it was shown that occupancy of only three of the five sites is required for optimal activation [34]. Structural Changes of Loop C when Interacting with Agonists and Small Competitive Antagonists Several lines of evidence indicate that ligands at the agonist binding site are stabilized by π-cation, dipole-cation, hydrogen bonding, and van der Waals interactions [35,36]. Agonist interaction produces activation (opening) of the AChR ion channel. The transition from the resting to the activated state is relatively fast, although the time regime is different among receptor subtypes. Molecular details about the activation process were determined by studies on AChBP- and AChR-bound ligand structures. For example, structural differences between the closed and open states permitted to determine that ACh elicits clockwise rotation of the inner β-sheet with respect to the outer J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal Figure 2: Structure of AChBP from Capitella teleta (Ct-AChBP) complexed with the partial agonists lobeline (A-C) and varenicline (D) respectively (modified from [30]). AChBP-lobeline structures showing the N-terminus pointing away (A) or toward (B) the viewer, and the subunit interface (principal face in yellow and complementary face in blue) towards the viewer (C). Each of the five subunits is displayed in a different color. Oxygen, red; nitrogen, blue; sulfur, green. The glycosyl chain N-linked to N122 is shown in magenta. The paucimannose chain is composed of N-acetylglucosamine (NAG), α-Dmannose (α-MAN), and β-D-mannose (β-MAN) (D) Ct-AChBP-varenicline complex. Lobeline and varenicline complexed with Ct-AChBP are shown as spheres, whereas the insets show the electron density of each molecule. Volume 3 • Issue 4 • 1000116 Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Page 4 of 7 Structure Ligand bound Conformational state PDB Resolution (Å) Reference AChR AChR None ACh ACh α-BTx α-BTx HEPES HEPES (-)-Nicotine CCh α-CbTx DHβE n-Acetyl-d-glucosamine (+)-Epibatidine CAPS None (+)-Epibatidine α-Lobeline MLA α-CTx ImI HEPES α-CTx PnIA(A10L/D14K) SPX GYM α-CTx ImI α-CTx TxIA(A10L) Cocaine Galantamine Imidacloprid Thiacloprid Clothianidin Imidacloprid Sulfate Anabaseine DMXBA 4OH-DMXBA Tropisetron None Compound 31 Compound 35 Metocurine d-TC Fragment 1 Compound 3 Compound 4 Compound 6 MTSET+ MMTS and ACh d-TC Strychnine Mutant I Mutant II + MLA Mutant III + MLA Compound 18 (complex I) Compound 6 (complex II) Compound 6 (complex III) Varenicline Cytisine Triazole 18 None None None Varenicline α-Lobeline Compound 5 Closed (Resting) Closed (Desensitized) Open Open Intermediate Intermediate Closed Closed Open Closed Apo Closed Intermediate Apo Closed Closed; g-to-t Intermediate Open Intermediate Open Intermediate Intermediate Open Open Unchanged Unchanged Closed Closed Closed Closed Coordinated with Lys residues Closed Intermediate Intermediate Intermediate Apo; g Intermediate Intermediate Open; g g-to-t g-to-t endo Open Closed Open Closed Apo Open Open Open Closed Closed Pentagonal dodecahedron 2BG9 4AQ5 4AQ9 2QC1 1HCB 1I9B 1UX2 1UW6 1UV6 1YI5 4ALX 3SQ9 3SQ6 2BJ0 2BYN 2BYQ 2BYS 2BYR 2BYP 2BR7 2BR8 2WZY 2X00 2C9T 2UZ6 2PGZ 2PH9 3C79 3C84 2ZJV 2ZJU 3GUA 2WNL 2WNJ 2WN9 2WNC 2Y7Y 2W8F 2W8G 3PEO 3PMZ 2Y54 2Y56 2Y57 2Y58 2XZ6 2XZ5 2XYT 2XYS 3T4M 3SH1 3SIO 2XNT 2XNU 2XNV 4AFT 4AFO 4DBM 4AOD 4AOE 4AFG 4AFH 4B5D 4.0 6.2 6.2 1.94 1.8 2.70 2.2 2.2 2.5 4.20 2.51 3.10 2.80 2.0 2.02 3.40 2.05 2.45 2.07 3.0 2.4 2.51 2.40 2.2 2.4 1.76 2.88 2.48 1.94 2.70 2.58 3.10 2.70 1.80 1.75 2.20 1.89 2.7 2.8 2.10 2.44 3.65 3.59 3.30 3.25 3.14 2.80 2.05 1.91 3.00 2.90 2.32 3.21 2.55 2.44 3.20 2.88 2.30 ~6 ~6 Low resolution 2.0 1.9 2.3 [19] [20] α1 α1-13-residue peptide Ls-AChBP Ls-AChBP Ls-AChBP Ls-AChBP hα7/Ls-AChBP chimera Bt-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP Ac-AChBP (Y53C mutant) Ac-AChBP hα7/Ac-AChBP chimera Ac-AChBP Ac-AChBP Ac-AChBP Bg-AChBP1 Bg-AChBP2 Ct-AChBP Ct-AChBP Ct-AChBP Apo Intermediate; g-to-t Closed; g-to-t Intermediate [28] [52] [21] [53] [54] [42] [62] [55] [46] [56] [49] [57] [50] [43] [58] [16] [59] [45] [44] [60] [39] [51] [40] [61] [63] [41] [64] [29] [65] [30] [66] AChRs are obtained from Torpedo marmorata electric fish. AChBPs are obtained from Lymnaea stagnalis (Ls), Aplysia californica (Ac), Bulinus truncates (Bt), Biomphalaria glabrata (Bg), and Capitella teleta (Ct), respectively. The close (i.e., contracted) and opened (i.e., extended) states of Loop C in the AChBP-ligand complexes indicate the active/desensitized (i.e., agonists) and resting (i.e., antagonists) states, respectively. Several ligands, including partial agonists, adopt an intermediate configuration between that for full agonists and competitive antagonists, whereas allosteric modulators (e.g., cocaine and galantamine) do not produce any apparent change on Loop C. The g-to-t (also called Tyr-flip) conformational state corresponds to the opening of the lobeline pocket. The g state is the closed lobeline pocket, whereas in the endo configuration the ligand is unable to induce the opening of the lobeline pocket. Table 1: Conformation states of AChRs and AChBPs in unbound and bound conditions. J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal Volume 3 • Issue 4 • 1000116 Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Page 5 of 7 to induce the opening of the lobeline pocket. These results suggest that the different occupation and orientation between full and partial agonists may produce distinct conformations of Loop C. Interestingly, α-lobeline binds to Ct- and Ac-AChBPs at different configurations [30,46]. In Ct-AChBP, the α-lobeline piperidine ring adopts a chair conformation and its hydroxyl group is in the S-configuration, whereas in the Ac-AChBP the ring adopts a half-chair conformation and its hydroxyl group is in the R-configuration. The AChBP structures in the absence of ligands or buffer molecules (i.e., Apo) show that the Loop C is more flexible than that in the presence of ligands, and that there are ordered water molecules filling the pocket. This structure is reminiscent to the Torpedo AChR model in the resting state [19]. Instead, buffer molecules (e.g., HEPES, CAPS) in the binding pocket move Loop C in a more open configuration, suggesting an intermediate state [46]. Structural Changes of Different Loops when Interacting with Large Competitive Antagonists Figure 3: Structure of AChBP from Lymnaea stagnalis (Ls-AChBP) complexed with the competitive antagonist dihydro-β-erythroidine (DHβE) (modified from [42] ). (a) Structure of DHβE. (b) Complex viewed along the five-fold symmetry axis. The five subunits are shown in different colors and DHβE in red spheres. (c) Interfacial binding pocket formed by the highly conserved aromatic residues Tyr89, Trp143, Tyr185, and Tyr192 from the principal side of the interface (yellow) and Trp53 from the complementary side (limon). DHβE is shown in red, and hydrogen bonds between DHβE and its surroundings are shown as stippled lines. Some other ligands produce an intermediate state, whereas the allosteric modulators cocaine and galantamine do not produce any apparent conformational change [43]. For example, the dibenzosuberylatropine analogs, compounds 31 and 35, which pharmacologically act as mixed competitive/noncompetitive antagonists on AChRs [44], induce an intermediate open configuration of Loop C, suggesting an intermediate resting/activated state. Interestingly, partial agonists of the α7 AChR such as DMXBA [3-(2,4-dimethoxybenzylidene)anabaseine], its hydroxyl metabolite 4OH-DMXBA, and tropisetron, display multiple orientations within the five binding sites and adopt an intermediate configuration between that for full agonists and competitive antagonists [45]. Nevertheless, the partial agonist α-lobeline induces a strong Loop C closure [30,46]. More specifically, the interaction of α-lobeline with the AChBP site opens a subpocket to accommodate the α-hydroxyphenetyl moiety, inducing the g-to-t (also called Tyr-flip) conformational state [39], where the g state is the closed lobeline pocket, whereas in the endo configuration the ligand is unable J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal Competitive antagonists overlap the agonist sites and inhibit their pharmacological action. In addition to this competitive (steric) mechanism, there is information that some antagonists [e.g., αand κ-bungarotoxin (α- and κ-BTx), α-cobrotoxin (α-CbTx), and methyllycaconitine (MLA)] maintain the AChR in the resting state [36,47,48], probably inhibiting the gating process. At the molecular level, MLA, the potent competitive antagonist of the α7 AChR, induces an intermediate open configuration of Loop C [46]. Although the Loop C conformation distinguishes agonists from competitive antagonists, additional results indicate that Loop F is another structural component responsible for ligand selectivity; especially for antagonists [45]. Although the toxins 13-desmethyl spirolide C (SPX) obtained from Alexandrium ostenfeldii and gymnodimine A (GYM) obtained from Karenia selliformis are among the most potent non-peptidic antagonists (i.e., binding affinities in the subnanomolar concentration range), they are nonselective antagonists, probably because they interact in less proportion with Loop F in the complementary face [49]. Crystallographic information supports the view that in general peptidic antagonists (i.e., snail and snake neurotoxins) maintain the AChBP in the resting state, whereas other nonpeptidic antagonists induce an intermediate state (Table 1). Different snail α-conotoxins (α-CTxs) also differ in their intrinsic orientations [50], probably resembling their receptor selectivities. The Ac-AChBP-Y53C mutant binds to Cys-modifying agents by different manners [51]. In this regard, MTSET+ [2-(trimethylammonium) ethylmethanethiosulfonate] produces a less constricted Loop C, similar to peptidic antagonists, whereas the conformation of Loop C in the presence of MMTS (methylmethanethiosulfonate) and ACh resembles that for agonists (Table 1). Since the location of the conserved amino acid Y53 is at Loop D, these results also emphasize the concept of concerted interactions between Loops D and C. Another important difference between agonists and antagonists is that cation–π interactions are formed between the conserved Trp residue in Loop B (i.e., W143 in Ls-AChBP) and agonists such as (-)-nicotine [53], whereas no cation–π interactions are formed with antagonists such as α-CTxs [46,50,56,57] and neonicotinoids [16,58]. Taking into account all studied ligand interactions, a good correlation (with some few exceptions) between Loop C closure and the type of ligand (i.e., full and partial agonists, competitive antagonists, and modulators) was found [1]). In addition to details on ligand interactions, AChBP studies have also helped in finding an ion selectivity filter in the extracellular domain Volume 3 • Issue 4 • 1000116 Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Page 6 of 7 of Cys-loop ligand-gated ion channels [59]. This selectivity filter is negatively charged in cation ion channels and positively charged in anion ion channels. Concluding Remarks AChRs mediate rapid transmission throughout the nervous system and also present functional roles in non-neuronal tissues. Structural and functional studies permitted to highlight the importance of the extracellular domain for the binding of agonists, competitive antagonists, and allosteric modulators. It is remarkable the advancement in our knowledge on ligand recognition and binding through the study of AChBPs from mollusks and annelids bound to a variety of ligands at atomic resolution. The current knowledge of the structural components of the AChR binding sites is paramount in differentiating agonists (full vs. partial) from competitive antagonists (small and large) and modulators. The crystallographic results in combination with mutagenesis, biochemical, electrophysiological, and animal behavior studies will help in the design of more selective agonists and competitive antagonists that can be used for the treatment of AChR-related diseases. For example, compound 5 (3-[(2(S)azetidinyl)methoxy]-5-[(1S,2R)-2-(2-hydroxyethyl)cyclopropyl] pyridine) is a highly selective partial agonist of the α4β2 AChR with antidepressant properties that has been recently co-crystallized with an AChBP [60-66]. Acknowledgement I would like to thank Katarzyna Targowska-Duda (Medical University of Lublin, Poland) for superimposing the open and closed Torpedo AChR models. References 1. Arias HR (2011) Functional components of nicotine acetylcholine receptors: Pharmacology of Nicotinic Acetylcholine Receptors from the Basic and Therapeutic Perspectives. In: H.R. Arias (ed). Chapter 1. Research Signpost, Kerala, India. 1-18. 2. Albuquerque EX, Pereira EF, Alkondon M, Rogers SW (2009) Mammalian nicotinic acetylcholine receptors: from structure to function. Physiol Rev 89: 73-120. 3. Kalamida D, Poulas K, Avramopoulou V, Fostieri E, Lagoumintzis G, et al. (2007) Muscle and neuronal nicotinic acetylcholine receptors. Structure, function and pathogenicity. FEBS J 274: 3799-3845. 4. Le Novère N, Corringer PJ, Changeux JP (2002) The diversity of subunit composition in nAChRs: Evolutionary origins, physiologic and pharmacologic consequences. J Neurobiol 53: 447-456. 5. Millar NS, Harkness PC (2008) Assembly and trafficking of nicotinic acetylcholine receptors (Review). Mol Membr Biol 25: 279-292. 6. Araud T, Wonnacott S, Bertrand D (2010) Associated proteins: The universal tool box controlling ligand gated ion channel function. Biochem Pharmacol 80: 160-169. 7. Moroni M, Zwart R, Sher E, Cassels BK, Bermudez I (2006) α4β2 nicotinic receptors with high and low acetylcholine sensitivity: pharmacology, stoichiometry, and sensitivity to long-term exposure to nicotine. Mol Pharmacol 70: 755-768. 8. Moroni M, Vijayan R, Carbone A, Zwart R, Biggin PC, et al. (2008) Non-agonistbinding subunit interfaces confer distinct functional signatures to the alternate stoichiometries of the α4β2 nicotinic receptor: an α4-α4 interface is required for Zn2+ potentiation. J Neurosci 28: 6884-6894. 9. Tapia L, Kuryatov A, Lindstrom J (2007) Ca2+ permeability of the (α4)3(β2)2 stoichiometry greatly exceeds that of (α4)2(β2)3 human acetylcholine receptors. Mol Pharmacol 71: 769-776. 10. Anderson DJ, Malysz J, Grønlien JH, El Kouhen R, Håkerud M, et al. (2009) 86 Rb+ efflux mediated by α4β2*-nicotinic acetylcholine receptors with high and low-sensitivity to stimulation by acetylcholine display similar agonist-induced desensitization. Biochem Pharmacol 78: 844-851. J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal 11. Grady SR, Salminen O, McIntosh JM, Marks MJ, Collins AC (2010) Mouse striatal dopamine nerve terminals express α4α5β2 and two stoichiometric forms of α4β2*-nicotinic acetylcholine receptors. J Mol Neurosci 40: 91-95. 12. Mousa SA, Arias HR (2010) Angiogenesis modulation by nicotine and nicotinic ligands. J Pediatr Biochem 1: 91-104. 13. Mousa SA, Mousa SS, Arias HR (2011) Physiological functions of non-neuronal nicotinic receptors: Angiogenesis modulation: Pharmacology of Nicotinic Acetylcholine Receptors from the Basic and Therapeutic Perspectives (H.R. Arias, Ed.), Research Signpost, Kerala, India, Chapter 5: 83-99. 14. Arias HR (2009) Is the inhibition of nicotinic acetylcholine receptors by bupropion involved in its clinical actions? Int J Biochem Cell Biol 41: 2098-2108. 15. Arias HR (2010) Molecular interaction of bupropion with nicotine acetylcholine receptors. J Pediatr Biochem 1: 185-197. 16. Ihara M, Okajima T, Yamashita A, Oda T, Hirata K, et al. (2008) Crystal structures of Lymnaea stagnalis AChBP in complex with neonicotinoid insecticides imidacloprid and clothianidin. Invert Neurosci 8: 71-81. 17. Akdemir A, Leurs R, de Esch IJP (2011) The acetylcholine binding protein as a template for the ligand binding domains of the homologous nicotinic receptors: Pharmacology of Nicotinic Acetylcholine Receptors from the Basic and Therapeutic Perspectives. Research Signpost, Kerala, India. 18. Corringer PJ, Baaden M, Bocquet N, Delarue M, Dufresne V, et al. (2010) Atomic structure and dynamics of pentameric ligand-gated ion channels: new insight from bacterial homologues. J Physiol 588: 565-572. 19. Unwin N (2005) Refined structure of the nicotinic acetylcholine receptor at 4 Å resolution. J Mol Biol 346: 967-989. 20. Unwin N, Fujiyoshi Y (2012) Gating movement of acetylcholine receptor caught by plunge-freezing. J Mol Biol 422: 617-634. 21. Brejc K, van Dijk WJ, Klaassen RV, Schuurmans M, van Der Oost J, et al. (2001) Crystal structure of an ACh-binding protein reveals the ligand-binding domain of nicotinic receptors. Nature 411: 269-276. 22. Zimmermann I, Dutzler R (2011) Ligand activation of the prokaryotic pentameric ligand-gated ion channel ELIC. PLOS Biol 9: 1-12. 23. Hilf RJ, Bertozzi C, Zimmermann I, Reiter A, Trauner D, et al. (2010) Structural basis of open channel block in a prokaryotic pentameric ligand-gated ion channel. Nat Struct Mol Biol 17: 1330-1336. 24. Nury H, Van Renterghem C, Weng Y, Tran A, Baaden M, et al. (2011) X-ray structures of general anaesthetics bound to a pentameric ligand-gated ion channel. Nature 469: 428-431. 25. Arias HR (2010) Positive and negative modulation of nicotinic receptors: Advances in ProteinChemistry and Structural Biology. R Donev (Ed.), Elsevier Inc, USA 80: 153-204. 26. Arias HR (2011) Allosteric modulation of nicotine acetylcholine receptors: Pharmacology of Nicotinic Acetylcholine Receptors from the Basic and Therapeutic Perspectives. Research Signpost, Kerala, India. 27. Arias HR, Bhumireddy P, Bouzat C (2006) Molecular mechanisms and binding site locations for noncompetitive antagonists of nicotinic acetylcholine receptors. Int J Biochem Cell Biol 38: 1254-1276. 28. Dellisanti CD, Yao Y, Stroud JC, Wang ZZ, Chen L (2007) Crystal structure of the extracellular domain ofthe nAChR α1 bound to α-bungarotoxin at 1.94 Å resolution. Nat Neurosci 10: 953-962. 29. Saur M, Moeller V, Kapetanopoulos K, Braukmann S, Gebauer W, et al. (2012) Acetylcholine-binding protein in the hemolymph of the planorbid snail Biomphalaria glabrata is a pentagonal dodecahedron (60 subunits). PLoS ONE 7: 1-14. 30. Billen B, Spurny R, Brams M, van Elk R, Valera-Kummer S, et al. (2012) Molecular actions of smoking cessation drugs at α4β2 nicotinic receptors defined in crystal structures of a homologous binding protein. Proc Natl Acad Sci U S A 109: 9173-9178. 31. Changeux JP, Taly A (2008) Nicotinic receptors, allosteric proteins and medicine. Trends Mol Med 14: 93-102. 32. Tasneem A, Iyer LM, Jakobsson E, Aravind L (2005) Identification of the prokaryotic ligand-gated ion channels and their implications for the mechanisms and origins of animal Cys-loop ion channels. Genome Biol 6: R4.1-R4.12. Volume 3 • Issue 4 • 1000116 Citation: Arias HR (2012) Molecular Interactions between Ligands and Nicotinic Acetylcholine Receptors Revealed by Studies with Acetylcholine Binding Proteins. J Thermodynam Cat 3:116. doi:10.4172/2157-7544.1000116 Page 7 of 7 33. Rayes D, De Rosa MJ, Sine SM, Bouzat C (2009) Number and locations of agonist binding sites required to activate homomeric Cys-loop receptors. J Neurosci 29: 6022-6032. of a Cbtx-AChBP complex reveals essential interactions between snake α-neurotoxins and nicotinic receptors. EMBO J 24: 1512-1522. 34. Xiu X, Puskar NL, Shanata JA, Lester HA, Dougherty DA (2009) Nicotine binding to brain receptors requires a strong cation-π interaction. Nature 458: 534-537. 54. Celie PH, Klaassen RV, van Rossum-Fikkert SE, van Elk R, van Nierop, P, et al. (2005) Crystal structure of acetylcholine-binding protein from Bulinustruncatus reveals the conserved structural scaffold and sites of variation in nicotinic acetylcholine receptors. J Biol Chem 280: 26457-26466. 35. Arias HR, Gu H, Feuerbach D, Wei DQ (2010) Different interaction between the agonist JN403 and the competitive antagonist methyllycaconitine with the human α7 nicotinic receptor. Biochemistry 49: 4169-4180. 55. Celie PHN, Kasheverov IE, Mordvintsev DY, Hogg RC, van Nierop P, et al. (2005) Crystal structure of nicotinic acetylcholine receptor homolog AChBP in complex with an α-conotoxinPnIA variant. Nature Struct Mol Biol 12: 582-588. 36. Law RJ, Henchman RH, McCammon JA (2005) A gating mechanism proposed from a simulation of a human a7 nicotinic acetylcholine receptor. Proc Natl Acad Sci USA 102: 6813-6818. 56. Ulens C, Hogg RC, Celie PH, Bertrand D, Tsetlin V, et al. (2006) Structural determinants of selective α-conotoxin binding to a nicotinic acetylcholine receptor homolog AChBP. Proc Natl Acad Sci USA 103: 3615-3620. 37. Mukhtasimova N, Lee WY, Wang HL, Sine SM (2009) Detection and trapping of intermediate states priming nicotinic receptor channel opening. Nature 459: 451-454. 57. Talley TT, Harel M, Hibbs RE, Radic Z, Tomizawa M, et al. (2008) Atomic interactions of neonicotinoid agonists withAChBP: Molecular recognition of the distinctive electronegative pharmacophore. Proc Natl Acad Sci USA 105: 76067611. 38. Edink E, Rucktooa P, Retra K, Akdemir A, Nahar T, et al. (2011) Fragment growing induces conformational changes in acetylcholine-binding protein: a structural and thermodynamic analysis. J Am Chem Soc 133: 5363-5371. 39. Brams M, Pandya A, Kuzmin D, van Elk R, Krijnen L, et al. (2011) A structural and mutagenic blueprint for molecular recognition of strychnine and d-tubocurarine by different Cys-loop receptors. PLoS Biol 9: 1-12. 40. Rucktooa P, Haseler CA, van Elk R, Smit AB, Gallagher T, et al. (2012) Structural characterization of the binding mode of smoking cessation drugs to nicotinic acetylcholine receptors through the study of ligand complexes with acetylcholine binding protein. J Biol Chem 287: 23283-23293. 41. Shahsavar A, Kastrup JS, Nielsen EØ, Kristensen JL, Gajhede M, et al. (2012) Crystal structure of Lymnaea stagnalis AChBP complexed with the potent nAChR antagonist DHβE suggests a unique mode of antagonism. PLoS ONE 7: 1-6. 58. Hansen SB, Wang HL, Taylor P, Sine SM (2008) An ion selectivity filter in the extracellular domain of Cys-loop receptors reveals determinants for ion conductance. J Biol Chem 283: 36066-36070. 59. Talley TT, Harel M, Yamauchi GJ, Radic Z, Hansen S, et al. (2010) The curare alkaloids: analyzing the poses of complexes with the acetylcholine binding protein in relation to structure and binding energetics. RCSB Protein Data Bank. 60. Nemecz A, Taylor PW (2011) Creating an α7 nicotinic acetylcholine recognition domain from the acetylcholine binding protein: crystallographic and ligand selectivity analyses. J Biol Chem 286: 42555-42565. 61. Li SX, Huang S, Bren N, Noridomi K, Dellisanti C, et al. (2011) Ligand-binding domain of an α7-nicotinic receptor chimera and its complex with agonist. Nature Neurosci 14: 1253-1259. 42. Hansen SB, Taylor P (2007) Galanthamine and non-competitive inhibitor binding to ACh-binding protein: Evidence for a binding site on non-a-subunit interfaces of heteromeric neuronal nicotinic receptors. J Mol Biol 369: 895-901. 62. Akdemir A, Rucktooa P, Jongejan A, Elk R, Bertrand S, et al. (2011) Acetylcholine binding protein (AChBP) as template for hierarchical in silico screening procedures to identify structurally novel ligands for the nicotinic receptors. Bioorg Med Chem 19: 6107-6119. 43. Ulens C, Akdemir A, Jongejan A, van Elk R, Bertrand S, et al. (2009) Use of acetylcholine binding protein in the search for novel α7 nicotinic receptor ligands. In silico docking, pharmacological screening, and X-ray analysis. J Med Chem 52: 2372-2383. 63. Grimster NP, Stump B, Fotsing JR, Weide T, Talley TT, et al. (2012) Generation of candidate ligands for nicotinic acetylcholine receptors via in situ click chemistry with a soluble acetylcholine binding protein template. J Am Chem Soc 134: 6732-6740. 44. Hibbs RE, Sulzenbacher G, Shi J, Talley TT, Conrod S, et al. (2009) Structural determinants for interaction of partial agonists with acetylcholine binding protein and neuronal α7 nicotinic acetylcholine receptor. EMBO J 28: 3040-3051. 45. Hansen SB, Sulzenbacher G, Huxford T, Marchot P, Taylor P, et al. (2005) Structures of Aplysia AChBP complexes with agonists and antagonists reveal distinctive binding interfaces and conformations. EMBO J 24: 3635-3646. 46. Moore MA, McCarthy MP (1995) Snake venom toxins, unlike smaller antagonists, appear to stabilize a resting state conformation of the nicotinic acetylcholine receptor. Biochim Biophys Acta 1235: 336-342. 47. Nasiripourdori A, RanjbarB, Naderi-Manesh H (2009) Binding of long-chain α-neurotoxin would stabilize the resting state of nAChR: A comparative study with α-conotoxin. Theor Biol Med Model 6: 1-15. 64. McCormack T, PetrovichRM, MercierKA, De RoseEF, Cuneo MJ, et al. (2010) Identification and functional characterization of a novel acetylcholine-binding protein from the marine annelid Capitella teleta. Biochemistry 49: 2279-2287. 65. Zhang HK, Eaton JB, Yu LF, Nys M, Mazzolari A, et al. (2012) Insights into the structural determinants required for high-affinity binding of chiralcyclopropanecontaining ligands to α4β2-nicotinic acetylcholine receptors: An integrated approach to behaviorally active nicotinic ligands. J Med Chem 55: 8028-8037. 66. Zhang H, Tückmantel W, Eaton JB, Yuen PW, Yu LF, et al. (2012) Chemistry and behavioral studies identify chiral cyclopropanes as selective α4β2-nicotinic acetylcholine receptor partial agonists exhibiting an antidepressant profile. J Med Chem 55: 717-724. 48. Bourne Y, Radić Z, Aráoz R, Talley TT, Benoit E, et al. (2010) Structural determinants in phycotoxins and AChBP conferring high affinity binding and nicotinic AChR antagonism. Proc Natl AcadSci USA 107: 6076-6081. 49. Dutertre S, Ulens C, Büttner R, Fish A, van Elk R, et al. (2007) AChBP-targeted α-conotoxin correlates distinct binding orientations with nAChR subtype selectivity. EMBO J 26: 3858-3867. 50. Brams M, Gay EA, Sáez JC, Guskov A, van Elk R, et al. (2011) Crystal structures of a cysteine-modified mutant in loop D of acetylcholine-binding protein. J Biol Chem 286: 4420-4428. 51. Harel M, Kasher R, Nicolas A, Guss JM, Balass M, et al. (2001) The binding site of acetylcholine receptor as visualized in the X-ray structure of a complex between α-bungarotoxin and a mimotope peptide. Neuron 32: 265-275. 52. Celie PH, van Rossum-Fikkert SE, van Dijk WJ, Brejc K, Smit AB, et al. (2004) Nicotine and carbamylcholine binding to nicotinic acetylcholine receptors as studied in AChBP crystal structures. Neuron 41: 907-914. 53. Bourne Y, Talley TT, Hansen SB, Taylor P, Marchot P (2005) Crystal structure J Thermodynam Cat ISSN: 2157-7544 JTC, an open access journal Volume 3 • Issue 4 • 1000116