Accepted Manuscript Title: Allosteric modulation of nicotinic acetylcholine receptors Author: Anna Chatzidaki Neil S. Millar PII: DOI: Reference:

S0006-2952(15)00417-7 http://dx.doi.org/doi:10.1016/j.bcp.2015.07.028 BCP 12321

To appear in:

BCP

Received date: Accepted date:

15-6-2015 24-7-2015

Please cite this article as: Chatzidaki Anna, Millar Neil S.Allosteric modulation of nicotinic acetylcholine receptors.Biochemical Pharmacology http://dx.doi.org/10.1016/j.bcp.2015.07.028 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Allosteric modulation of nicotinic acetylcholine receptors Anna Chatzidaki and Neil S. Millar Department of Neuroscience, Physiology & Pharmacology, University College London, London, WC1E 6BT, UK Address correspondence to: Neil S Millar, Department of Neuroscience, Physiology & Pharmacology, University College London, London, WC1E 6BT, UK. Tel. +44 (0)20 7679 7241; Email: [email protected] Graphical abstract ABSTRACT Nicotinic acetylcholine receptors (nAChRs) are receptors for the neurotransmitter acetylcholine and are members of the ‘Cys-loop’ family of pentameric ligand-gated ion channels (LGICs). Acetylcholine binds in the receptor extracellular domain at the interface between two subunits and research has identified a large number of nAChR-selective ligands, including agonists and competitive antagonists, that bind at the same site as acetylcholine (commonly referred to as the orthosteric binding site). In addition, more recent research has identified ligands that are able to modulate nAChR function by binding to sites that are distinct from the binding site for acetylcholine, including sites located in the transmembrane domain. These include positive allosteric modulators (PAMs), negative allosteric modulators (NAMs), silent allosteric modulators (SAMs) and compounds that are able to activate nAChRs via an allosteric binding site (allosteric agonists). Our aim in this article is to review important aspects of the pharmacological diversity of nAChR allosteric modulators and to describe recent evidence aimed at identifying binding sites for allosteric modulators on nAChRs. Keywords: Allosteric modulation; Ion channel; Neurotransmitter-gated ion channel; Nicotinic acetylcholine receptor; Receptor

1

1. Introduction Nicotinic acetylcholine receptors (nAChRs) are excitatory receptors for the neurotransmitter acetylcholine and play an important role in the central and peripheral nervous system [1-4]. They are members of the ‘Cys-loop’ family of pentameric ligand-gated ion channels (LGICs); a family that includes a subset of receptors for 5-hydroxytryptamine (5-HT), γ-aminobutyric acid (GABA) and glycine [5-7]. Nicotinic receptors are widely expressed in the brain in both pre- and post-synaptic locations [8, 9]. They are also expressed in both sympathetic and parasympathetic ganglia, where they are responsible for fast synaptic transmission. In addition, nAChRs are expressed in skeletal muscle, epithelial and immune cells [10-13]. Nicotinic receptors have been implicated in a number of neuromuscular, neurological and psychiatric disorders. For example, in recent years, neuronal nAChRs have been identified as important targets for therapeutic drug discovery, in connection with disorders such as Alzheimer’s disease and schizophrenia [14, 15]. Sixteen human nAChR subunits (α1-α7, α9, α10, β1-β4, γ, δ and ε) have been identified. The α1, β1, γ, δ and ε subunits are expressed in muscle, whereas the α2-α7, α9, α10 and β1-β4 subunits are expressed more widely and are commonly referred to as neuronal subunits [16, 17]. Receptors expressed at the neuromuscular junction are heteromeric complexes containing two copies of the α1 subunit co-assembled with three non-α subunits, with receptors in embryonic and adult muscle having the subunit composition of (α1)2β1γδ and (α1)2β1δε, respectively [18, 19]. There is, however, considerably greater subunit diversity amongst neuronal nAChRs [8, 17]. Some neuronal nAChR subunits, such as α7, can form functional homomeric [(α7)5] nAChRs [20], whereas the majority of neuronal nAChR subunits form heteromeric complexes, consisting of at least two α-type subunits co-assembled with at least two β-type subunits, for example (α4)2(β2)3 and (α4)3(β2)2 [17]. Each nAChR subunit contains an amino-terminal extracellular domain and four transmembrane helices (TM1TM4). The pore of the channel is lined by the second transmembrane domain (TM2) from five co-assembled subunits [21] (Figure 1).

2

2. Allosteric modulation of nAChRs The endogenous neurotransmitter of nAChRs, acetylcholine, binds in the receptor extracellular domain, at the interface between two subunits [22, 23]. Research conducted over many years has identified a large number of nAChR-selective ligands, including agonists and competitive antagonists that bind at the same site as acetylcholine (commonly referred to as the orthosteric binding site). In addition, more recent research has identified ligands that are able to modulate nAChR function by binding to sites that are distinct from the binding site for acetylcholine, including sites located in the transmembrane domain. As has been discussed previously, there can be some confusion as to the use of the term ‘allosteric’ in receptor pharmacology [24]. We will use the term ‘allosteric site’ to describe any nAChR ligand-binding site that is distinct from the conventional acetylcholine binding site and we will use the term ‘allosteric modulator’ to describe any ligand that alters the functional properties of nAChRs by interacting with a site that is distinct from the orthosteric site. In doing so, we do not necessarily intend to imply a particular mechanism of action for such ligands. Instead, our aim in this review is to describe evidence for receptor modulation that occurs as a consequence of ligands binding to sites that are distinct from that at which acetylcholine acts as an agonist. Allosteric modulators can either potentiate the effects of agonist-activation (positive allosteric modulators; PAMs), or inhibit agonist-activation (negative allosteric modulators; NAMs). Of course, quite correctly, the term non-competitive antagonist is also used extensively to describe ligands that are capable of inhibiting agonist-evoked responses through a distinct binding site. Possible mechanisms for modulation of nAChRs by allosteric ligands have been discussed previously [25-28]. For example, it has been proposed that PAMs that potentiate lower agonist concentration but have minimal effect on higher agonist concentrations may do so by facilitating agonist binding. PAMs may also increase agonist efficacy by reducing the energy barrier between closed and open states or by increasing the energy barrier between open and desensitised states (Figure 2). In addition, an increase in the apparent peak response and a large steady-state current might be a consequence of PAMs destabilising desensitised states. It has also been suggested that large molecules such as ivermectin may act as PAMs by binding at an inter-subunit transmembrane site and preventing the transition from the open to the closed state [28].

3

Allosteric modulators often have no intrinsic activity and only modulate the effects of an agonist. However allosteric ligands that can induce nAChR activation in the absence of an orthosteric agonist have been reported and referred to as either allosteric agonists or agoallosteric modulators [29, 30]. In addition, to PAMs and NAMs, compounds referred to as silent allosteric modulators (SAMs) have also been reported for nAChRs, as they have for other receptors [31-33]. Such compounds do not potentiate or inhibit responses to orthosteric agonists but they can block the effect of other allosteric modulators (PAMs, NAMs or allosteric agonists) by binding competitively at an allosteric binding site. One of the goals of therapeutic drug discovery is the identification of receptor ligands with high subtype selectivity. Although high subtype and species selectivity can be achieved with ligands binding to the orthosteric binding site (see, for example [34]), it is possible that receptor subtype selectivity might be achieved more easily by ligands binding to an allosteric site. This is because, whereas the acetylcholine-binding site is conserved within all nAChRs, there may be less need for conservation of structure or amino acid composition at other receptor sites. In addition, the low intrinsic activity of PAMs, in comparison to agonists, may be useful in reducing toxicity and off-target effects. This may be advantageous because it can enable the spatial and temporal pattern of signalling of endogenous acetylcholine-evoked responses to be retained [27, 35]. A number of previous reviews have discussed the structural diversity of nAChR allosteric modulators, including many compounds that have been described in patent applications [3640]. In addition, previous reviews have discussed the possible therapeutic uses of nAChRs allosteric modulators [27, 41-43] for example in the treatment of cognitive deficits [44]; depression [35]; pain [45, 46] and cancer [47]. In contrast, the present review does not attempt to provide a comprehensive overview of nAChR allosteric modulation but will focus on important aspects of the pharmacological diversity of allosteric modulators and on recent evidence aimed at identifying binding sites for allosteric modulators on nAChRs.

4

3. Positive allosteric modulators (PAMs) and allosteric agonists 3.1. Homomeric α7 nAChRs Much of the recent work concerning nAChR PAMs has focussed on the homomeric α7 receptor [36, 37, 40, 48], one of the main nAChR subtypes expressed in the mammalian brain. The α7 subtype is somewhat atypical in that it has a relatively low sensitivity to acetylcholine, high calcium permeability and exhibits very fast desensitisation [20]. In addition to potentiating agonist peak responses, some PAMs acting on α7 nAChRs have been reported to cause a slowing of receptor desensitisation, whereas other PAMs have little or no effect on the rate of receptor desensitisation. As a consequence of this differing effect on desensitisation, α7-selective PAMs have been classified into two categories: type I and type II [26, 49]. Type I PAMs increase peak current in the presence of an orthosteric agonist, without having an effect on receptor desensitisation. Type II PAMs, on the other hand, significantly reduce the fast desensitisation of α7 receptors. A further difference between α7-selective type I and type II PAMs is the ability of type II PAMs to allow receptor reactivation from the desensitised state [49, 50]. It is possible that the mechanism of action of type I PAMs consists of facilitating the transition to the open state by reducing the energy barrier between closed and open states. In addition, type II PAMs may cause destabilisation of the desensitised state. Although classifying PAMs acting on α7 nAChRs as either type I or type II can be useful in some circumstances, it is clear, as will be discussed later, that this is an over-simplification and that PAMs with intermediate properties have been identified. An early demonstration of allosteric modulation of nAChRs came from the observation that calcium potentiates α7 nAChR currents in a voltage-independent manner [51, 52]. Early studies with α7 nAChRs also demonstrated the ability of ivermectin, a large macrocyclic lactone (Figure 3), to potentiate agonist-evoked responses [53]. In addition, the acetylcholinesterase inhibitor galantamine has been reported to act as a relatively weak and non-selective PAM of α7 nAChRs [54]. Similarly, genistein (a tyrosine kinase inhibitor) and also 5-hydroxyindole (5-HI) are weak and relatively non-selective potentiators on α7 nAChRs [49, 55]. A number of proteins have been shown to enhance agonist responses on the α7 receptors, including a secreted mammalian Ly-6/uPAR-related protein (SLURP-1) and bovine serum albumin (BSA) [56, 57]. All of these compounds have only minimal effects on the

5

rapid desensitisation of α7 nAChRs and, in most cases, do not display high selectivity for this receptor subtype. Over the past decade there has been considerable interest from pharmaceutical companies and academic research groups in developing PAMs of α7 nAChRs that display higher potency and greater subtype selectivity. PNU-120596 was the first α7-selective PAM to be described in detail that causes a dramatic reduction in the rate of receptor desensitisation [50]. It is also one of the best-studied PAMs in pre-clinical models of pain, ischaemia, schizophrenia and cognitive deficits [50, 58-63]. Other α7-selective PAMs that cause a dramatic reduction in receptor desensitisation and, consequently, have been described as type II PAMs include TQS [49] and A-867744 [64, 65]. A feature of compounds such as PNU-120596, TQS and A-867744 is that they cause little or no receptor activation in the absence of an orthosteric agonist. However, a derivative of TQS (4BP-TQS) has been shown to activate recombinant and native α7 nAChRs in the absence of an orthosteric agonist [30, 66, 67]. In contrast to the activation of α7 nAChRs by acetylcholine, activation by 4BP-TQS occurs with minimal desensitisation [30]. In addition the agonist dose-response curve with 4BP-TQS is steeper and maximal responses are substantially larger than observed with orthosteric agonists, all of which argues that activation by 4BP-TQS occurs by a different mechanism of action [30]. As will be discussed later, there is evidence that 4BP-TQS causes receptor activation via a distinct allosteric site and, as a consequence, it has been described as an allosteric agonist [30]. Subsequently, a number of other TQS derivatives with either PAM or allosteric agonist effects have been reported [68, 69]. Interestingly, very minor changes to ligand structure, such as altering the size of a single halogen atom or the pattern of methyl substitution of an aromatic ring, can convert type II PAMs into potent allosteric agonists [33, 68]. In addition to allosteric modulators that reduce levels of agonist-induced desensitisation, a number of α7-selective type I PAMs have been described in recent years as a consequence of high-throughput compound screening. Amongst the first to be examined in detail were NS1738 [70] and CCMI (also described as ‘Compound 6’) [71]. Other compounds acting as type I PAMs of α7, but displaying less subtype selectivity, are LY-2087101, LY-2087133 and LY1078733 [72]. In addition to potentiating α7 nAChRs, these compounds also potentiate α2β4, α4β2, α4β4 subtypes [72].

6

Whilst the distinction between type I and type II PAMs can be useful, there is increasing evidence that it is an over-simplification. A number of studies have reported α7-selective PAMs with effects on desensitisation that are intermediate between classical type I and type II PAMs [73, 74]. A good example of this is a recently described series of structurally-related α7-selective PAMs (TBS-345, TBS-346, TBS-516, TBS-546 and TBS-556) with a range of effects on receptor desensitisation [75]. Another α7-selective PAM, RO5126946, has effects on desensitisation that are typical of a type II PAM [76]. However, it lacks the ability to facilitate reactivation of desensitised α7 nAChRs, a feature that is normally considered to be characteristic of type II PAMs [76]. 3.2. Heteromeric nAChRs Heteromeric α4β2-containing nAChRs are an abundant receptor subtype expressed in the human brain and have been a target for the development of orthosteric ligands. For example, the partial agonists varenicline and cytisine have been approved as aids for smoking cessation [77]. In addition, there has been considerable interest in the identification of PAMs that are selective for heteromeric nAChR subtypes such as α4β2. Steroids are among compounds that act as endogenous PAMs of nAChRs. 17β-estradiol increases acetylcholine-evoked currents in the human α4β2 receptor and, more modestly, in the α4β4 receptor [78-80]. Galantamine also potentiates α4β2 nAChRs, in addition to a number of other heteromeric nAChR subtypes, including α6β4 and α3β4 nAChRs [81]. Both estradiol and galantamine increase the potency (i.e. they reduce the EC50 value) of the orthosteric agonist, without having a significant effect on the size of the maximum response, which could be the result of enhanced agonist binding. In contrast, desformylflustrabromine (dFBr) and some of its derivatives are α4β2-selective PAMs that increase the size of the maximum response and have very modest effect on the potency of the agonist [82, 83]. LY2087101 is another relatively non-selective nAChR PAM that increases maximum response size as well as potency of acetylcholine on α4β2, [72]. Other compounds that have been reported to potentiate α4β2 and α4β4 nAChRs include the muscarinic antagonists atropine and scopolamine [84, 85]. Receptors containing α4 and β2 subunits can exist in two different stoichiometries with distinct functional properties. The (α4)3(β2)2 subtype has a lower sensitivity to acetylcholine

7

and it displays higher permeability to calcium and faster desensitisation kinetics in comparison to the (α4)2(β2)3 subtype [86-89]. In addition, the two subtypes display some ligand selectivity. NS9283 (alternatively A-969933) is an α4β2-selective PAM with in vivo efficacy in models of pain [90, 91] and is selective for the (α4)3(β2)2 subtype [92-94]. As will be discussed in more detail in section 6, NS206 is a PAM that is thought to act through a distinct binding site and can potentiate both the (α4)3(β2)2 and (α4)2(β2)3 subtypes [93]. In contrast, HEPES is a selective PAM for the (α4)2(β2)3 subtype [95]. In recent years, PAMs of α3-containing nAChRs have also been reported [96-98]. For example, the anthelmintic compounds levamisole and morantel potentiate acetylcholine response on α3β2 and α3β4. Zinc potentiates a number of nAChR subtypes, including both α3- and α4-containing subtypes [99, 100] but, interestingly, shows selectivity for α4β2 nAChRs with the (α4)3(β2)2 stoichiometry [101, 102]. Recent studies aimed at identifying binding sites for modulators such as zinc and morantel are discussed below.

4. Negative allosteric modulators (NAMs) Receptor antagonism can occur by a variety of mechanisms. A common distinction is made between antagonists that bind either competitively or non-competitively with respect to an orthosteric agonist. However, the terms ‘non-competitive antagonist’ and ‘negative allosteric modulator’ (NAM) are frequently used interchangeably. A distinction is also sometimes made between, on the one hand, antagonists that bind in the channel pore and thereby occlude the flow of ions (open-channel blockers) and, on the other hand, antagonists that modulate receptor properties such as gating or desensitisation equilibrium. A large number of compounds can act as open-channel blockers, including agonists at high concentrations. In addition, channel-blocking ligands frequently show relatively low receptor selectivity. Such compounds will not be discussed in detail but have been reviewed elsewhere [103, 104]. Many compounds have been reported to act as non-competitive antagonists of nAChRs. These include psychoactive drugs such as mecamylamine and phencyclidine [105-109], philanthotoxin [110], general anaesthetics [111], antidepressants [112-114], monoterpines such as camphor, menthol and propofol [115-119], aminoglycoside antibiotics [120] antipsychotics [121], the fluorophore crystal violet [122]; zinc [99, 123] and some detergents [124]. Several endogenous ligands have also been reported to have negative allosteric effects

8

on nAChRs. These include progesterone and neurosteroids [103, 125] and also proteins such as Lynx-1 [126, 127] and the cannabinoid anandamide [128].

5. Silent allosteric modulators (SAMs) The term ‘silent allosteric modulator’ (SAM) has been used to describe compounds on a variety of receptors that interact with a site that is distinct from the conventional allosteric site but do not exert a modulatory effect on responses evoked by an orthosteric ligand (i.e. neither a positive or negative allosteric effect) [31, 32, 129]. More recently, compounds with such properties have been reported for nAChRs [33]. Such compounds do not potentiate or inhibit responses to orthosteric agonists but they can influence allosteric modulation by blocking the effect of other allosteric modulators (PAMs, NAMs or allosteric agonists) by binding in a mutually exclusive manner at a common or overlapping allosteric site [33]. In addition, recent studies have demonstrated that relatively small changes in chemical structure, such as methyl substitution of a single aromatic ring, can determine whether a compound acts as a PAM, NAM or SAM on nAChRs [33].

6. Binding sites for allosteric modulators Several experimental techniques have been employed with the aim of identifying the binding sites of allosteric modulators on nAChRs (Table 1). The availability of a readily purified preparation of nAChR, for example from the electric organ of the marine ray Torpedo, has facilitated affinity labelling experiments that have provided direct evidence for the binding of ligands to sites other than the orthosteric site. This approach has identified sites for a variety of non-competitive antagonists interacting with the transmembrane domain [130-135]. Affinity labelling studies have also identified binding sites for cholesterol within the nAChR transmembrane domain [136]. In addition, more indirect approaches, such as computerdocking studies, have been used. For example, docking studies with cholesterol have predicted binding sites in cavities within the transmembrane region [137]. Other experimental techniques aimed at identifying binding sites have included the construction of recombinant subunit chimeras, site-directed mutagenesis, and the substituted cysteine accessibility method (SCAM). However, it is important to use a degree of caution when interpreting such studies. For example, if the mutation of an amino acid results in a change in the pharmacological properties of a ligand, it is not necessarily appropriate to conclude that this is a consequence of a change to the binding site [24]. Nevertheless, the combined use of a variety of

9

approaches, including electrophysiological, pharmacological, biochemical and computational techniques can provide strong evidence for the location of ligand binding sites [138]. Various sites and amino acids have been shown to be important for allosteric modulation. Studies using α7-5-HT3 subunit chimaeras, site-directed mutagenesis and docking simulations, have led to the proposal that α7-selective allosteric modulators such as NS-1738 (a type-I PAM) and PNU-120596 (a type II PAM) act via a binding site within an intrasubunit cavity that is located between the four transmembrane helices of a single subunit [139, 140]. More recent studies have examined a series of nineteen α7-selective allosteric modulators that differ only in methyl substitution of a single aromatic ring [33]. Despite relatively small changes in chemical structure, the compounds examined displayed five distinct pharmacological effects on α7 nAChRs. These included effects typical of type I PAMs, type II PAMs, NAMs, SAMs and allosteric agonists [33] and it has been proposed that all of these pharmacological effects can arise from ligands binding to a broadly similar or overlapping site located within the previously identified intra-subunit transmembrane cavity [33]. Other studies are consistent with α7-selective allosteric modulators such NS-1738, PNU120596 and A-867744 interacting with a transmembrane site [65, 141, 142]. As would be expected, allosteric modulators that have been proposed to bind in a transmembrane location do not displace the binding of orthosteric radioligands ligands such as [3H]-MLA or [3H]-αbungarotoxin [33, 65, 70, 75]. However, some unexpected results have been reported for the α7-selective PAM A-867744. Although A-867744 does not displace binding of [3H]-MLA from α7 nAChRs, in contrast to other PAMs, it has been reported to displace the binding of another agonist ([3H]-A-585539) that is thought to interact with the orthosteric site [65]. A transmembrane binding site in nAChRs has also been proposed for monoterpine compounds such as menthol and propofol [116, 117] and is consistent with evidence supporting a transmembrane binding site for monoterpines on other pentameric LGICs [143147]. Affinity labelling studies with a photoreactive analogue of propofol identified three sites at which it bound within the transmembrane domain of muscle-type nAChRs, but concluded that the functionally relevant site for the inhibitory action of propofol was an intra-subunit site [117], similar to that described earlier for α7-selective allosteric modulators such as NS-1738, PNU-120596 and 4BP-TQS [30, 139]. Similarly, there is evidence for a transmembrane binding site for nAChR NAMs such as the endogenous cannabinoid anandamide [148-150] and also antihistamine compounds [151].

10

As discussed earlier, the α7 subunit forms a homomeric nAChR. However, in addition, it can co-assemble with β2 to form a heteromeric α7β2 complex [152]. One difference that has been reported in the allosteric modulation of these two nAChR subtypes is that α7β2 receptors, but not α7 receptors, are inhibited by the volatile anaesthetic isoflurane [153]. On the basis of mutagenesis and computer docking studies, it has been proposed that isoflurane binds to the β2 transmembrane domain [153]. This is consistent with evidence for a transmembrane binding site for anaesthetics on other LGICs [154]. There is also evidence that the macrocyclic lactone ivermectin acts as a nAChR PAM by interacting with a transmembrane site [155]. Recently, a high resolution X-ray structure was obtained of ivermectin bound to a prokaryotic glutamate-gated chloride channel (GluCl) with close structural similarity to nAChRs [156]. These findings are consistent with ivermectin binding to an inter-subunit transmembrane site, rather than the intra-subunit transmembrane site that has been proposed for smaller allosteric modulators [30, 117, 139]. Spinosad, like ivermectin, is a macrocyclic lactone pesticide that acts as an allosteric modulator of nAChRs [157]. Recent studies of spinosad-resistant insects have identified a resistance-associated point mutation in the nAChR transmembrane domain that is consistent with spinosad binding to nAChRs in a similar inter-subunit transmembrane site to the known binding site of ivermectin in GluCl [158]. It addition to the transmembrane domain, there are many other potential sites at which allosteric modulators might potentially interact with nAChRs, as is the case with other ligandgated ion channels [25, 159]. For instance, a site on the C-terminus of the α4 subunit is believed to be necessary for estradiol PAM activity [78-80]. Whereas there are five potential agonist binding sites in a homomeric nAChR, there are expected to be just two or three functioning orthosteric agonist sites in heteromeric nAChRs. For example, the acetylcholine binding site in neuronal heteromeric nAChRs is at the interface between an α and β subunit, in which the α-type subunit forms the principal face and the β-type subunit the complementary face (designated α(+)/β(-)). However, it is possible that compounds can bind at equivalent positions at other subunit interfaces, for example at a β(+)/α(-) interface. Indeed, this has been proposed as a mechanism by which compounds such as morantel and desformylflustrabromine can potentiate agonist-evoked responses [96, 160-163]. Similarly, compounds interacting with the extracellular α4/α4 interface of (α4)3(β2)2 nAChR (or α/α interfaces in other nAChR subtypes) can act as PAMs [93, 164, 165], just as the inability of

11

an agonist to interact with this site may explain partial agonist activity [166-169]. In contrast, it has been proposed that HEPES acts as a potentiator of α4β2 nAChRs by interacting with the β(+)/β(-) interface and, as a consequence, is selective for receptors with the (α4)2(β2)3 stoichiometry [95]. Additionally, a binding site for NAMs at the α(+)/β(-) interface of α3β4 and α4β2 nAChRs has been predicted on the basis of computer docking studies and molecular dynamic simulations [170-172]. As described earlier, the acetylcholinesterase inhibitor galantamine is a relatively weak PAM of α7 nAChRs [173] and has been proposed to bind at the extracellular domain at a site that is distinct from the acetylcholine binding site [174, 175]. Galantamine also potentiates α4β2 nAChRs and a number of other nAChR subtypes including α6β4 and α3β4 nAChRs [81], by binding at a non-α subunit interface [176]. In addition, there is evidence that the binding site mediating the potentiating effects of calcium is located on the extracellular site of the α7 nAChR [125, 177]. Another divalent cation, zinc, acts as a PAM of α4β2 nAChRs but does so selectively on receptors with the stoichiometry (α4)3(β2)2 [102]. This has been explained by evidence that it interacts selectively with the α4(+)/α4(-) subunit interface, whilst inhibitory effects on (α4)2(β2)3 nAChRs are thought to be mediated by zinc binding to the α4(-)/β2(+) subunit interface [102].

7. Summary Rather than attempting to provide a comprehensive overview of all aspects of allosteric modulation of nAChRs, our aim in this brief review has been to summarise aspects of the pharmacological diversity of nAChR allosteric modulators and to describe evidence for the location of binding sites for such compounds. Although the focus of this review has been the family of nAChRs, the evidence for the existence of multiple modulatory sites, both in transmembrane and extra-transmembrane locations is a feature that appears to be common amongst neurotransmitter-gated ion channels.

Acknowledgements Research in the authors’ laboratory has been funded in recent years by the Biotechnology and Biological Sciences Research Council (BBSRC), Eli Lilly, the Medical Research Council (MRC), the Royal Society, Syngenta and the Wellcome Trust. Anna Chatzidaki was

12

supported by a Medical Research Council (MRC) Industrial CASE PhD studentship [Grant: G1001602] that was co-funded by the MRC and Eli Lilly.

References [1] Galzi JL, Revah F, Bessis A, Changeux JP. Functional architecture of the nicotinic acetylcholine receptor: from electric organ to brain. Annu Rev Pharmacol Toxicol. 1991;31:37-72. [2] Sargent PB. The diversity of neuronal nicotinic acetylcholine receptors. Annu Rev Neurosci. 1993;16:403-43. [3] McGehee DS, Role LW. Physiological diversity of nicotinic acetylcholine receptors expressed by vertebrate neurons. Annu Rev Physiol. 1995;57:521-46. [4] Dani JA, Bertrand D. Nicotinic acetylcholine receptors and nicotinic cholinergic mechanisms of the central nervous system. Ann Rev Pharmacol Toxicol. 2007;47:699-729. [5] Sine SM, Engel AG. Recent advances in Cys-loop receptor structure and function. Nature. 2006;440:448-55. [6] Lester HA, Dibas MI, Dahan DS, Leite JF, Dougherty DA. Cys-loop receptors: new twists and turns. Trends Neurosci. 2004;27:329-36. [7] Corringer PJ, Poitevin F, Prevost MS, Sauguet L, Delarue M, Changeux JP. Structure and pharmacology of pentameric receptor channels: from bacteria to brain. Structure. 2012;20:941-56. [8] Gotti C, Clementi F, Fornari A, Gaimarri A, Guiducci S, Manfredi I, et al. Structural and functional diversity of native brain neuronal nicotinic receptors. Biochem Pharmacol. 2009;78:703-11. [9] Wu J, Lukas RJ. Naturally-expressed nicotinic acetylcholine receptor subtypes. Biochem Pharmacol. 2011;82:800-7. [10] Matsunaga K, Klein TW, Friedman H, Yamamoto Y. Involvement of nicotinic acetylcholine receptors in suppression of antimicrobial activity and cytokine responses of alveolar macrophages to Legionella pneumophila infection by nicotine. J Immunol. 2001;167:6518-24. [11] Heeschen C, Jang JJ, Weis M, Pathak A, Kaji S, Hu RS, et al. Nicotine stimulates angiogenesis and promotes tumor growth and atherosclerosis. Nat Med. 2001;7:833-9. [12] Wang H, Yu M, Ochani M, Amella CA, Tanovic M, Susaria S, et al. Nicotinic acetylcholine receptor α7 subunit is an essential regulator of inflammation. Nature. 2003;421:384-8. [13] Gotti C, Clementi F. Neuronal nicotinic receptors: from structure to pathology. Prog Neurobiol. 2004;74:363-96. [14] Wallace TL, Bertrand D. Alpha7 neuronal nicotinic receptors as a drug target in schizophrenia. Expert Opin Ther Targets. 2013;17:139-55. [15] Lombardo S, Maskos U. Role of the nicotinic acetylcholine receptor in Alzheimer's disease pathology and treatment. Neuropharmacol. 2014. [16] Le Novère N, Corringer PJ, Changeux JP. The diversity of subunit composition in nAChRs: evolutionary origins, physiologic and pharmacologic consequences. J Neurobiol. 2002;53:447-56. [17] Millar NS, Gotti C. Diversity of vertebrate nicotinic acetylcholine receptors. Neuropharmacol. 2009;56:237-46.

13

[18] Mishina M, Takai T, Imoto K, Noda M, Takahashi T, Numa S, et al. Molecular distinction between fetal and adult forms of muscle acetylcholine receptor. Nature. 1986;313:364-9. [19] Missias AC, Chu GC, Klocke BJ, Sanes JR, Merlie JP. Maturation of the acetylcholine receptor in skeletal muscle: regulation of the AChR γ-to-ε switch. Dev Biol. 1996;179:22338. [20] Couturier S, Bertrand D, Matter JM, Hernandez MC, Bertrand S, Millar N, et al. A neuronal nicotinic acetylcholine receptor subunit (α7) is developmentally regulated and forms a homo-oligomeric channel blocked by α-BTX. Neuron. 1990;5:847-56. [21] Unwin N. Refined structure of the nicotinic acetylcholine receptor at 4 Å resolution. J Mol Biol. 2005;346:967-89. [22] Arias HR. Localization of agonist and competitive antagonist binding sites on nicotinic acetylcholine receptors. Neurochem Int. 2000;36:595-645. [23] Sine SM. The nicotinic receptor ligand binding domain. J Neurobiol. 2002;53:431-46. [24] Colquhoun D. Binding, gating, affinity and efficacy: the interpretation of structureactivity relationships for agonists and of the effects of mutating receptors. Br J Pharmacol. 1998;125:924-47. [25] Hogg RC, Buisson B, Bertrand D. Allosteric modulation of ligand-gated ion channels. Biochem Pharmacol. 2005;70:1267-76. [26] Bertrand D, Gopalakrishnan M. Allosteric modulation of nicotinic acetylcholine receptors. Biochem Pharmacol. 2007;74:1155-63. [27] Williams DK, Wang J, Papke RL. Positive allosteric modulators as an approach to nicotinic acetylcholine receptor-targeted therapeutics: advantages and limitations. Biochem Pharmacol. 2011;82:915-30. [28] Cecchini M, Changeux JP. The nicotinic acetylcholine receptor and its prokaryotic homologues: Structure, conformational transitions & allosteric modulation. Neuropharmacology. 2015;96:137-49. [29] Grønlien JH, Ween H, Thorin-Hagene K, Cassar S, Li J, Briggs CA, et al. Importance of M2-M3 loop in governing properties of genistein at the α7 nicotinic acetylcholine receptor inferred from α7/5-HT3A chimera. Eur J Pharmacol. 2010;647:37-47. [30] Gill JK, Savolainen M, Young GT, Zwart R, Sher E, Millar NS. Agonist activation of α7 nicotinic acetylcholine receptors via an allosteric transmembrane site. Proc Natl Acad Sci USA. 2011;108:5867-72. [31] Schann S, Mayer S, Franchet C, Frauli M, Steinberg E, Thomas M, et al. Chemical switch of a metabotropic glutamate receptor 2 silent allosteric modulator into dual metabotropic glutamate receptor 2/3 negative/positive allosteric modulators. J Med Chem. 2010;53:8775-9. [32] Burford NT, Clark JM, Wehrman TS, Gerritz SW, Banks M, O'Connell J, et al. Discovery of positive allosteric modulators and silent allosteric modulators of the μ-opioid receptor. Proc Natl Acad Sci USA. 2013;110:10830-5. [33] Gill-Thind JK, Dhankher P, D'Oyley JM, Sheppard TD, Millar NS. Structurally similar allosteric modulators of α7 nicotinic acetylcholine receptors exhibit five distinct pharmacological effects. J Biol Chem. 2015;290:3552-62. [34] Young GT, Broad LM, Zwart R, Astles PC, Bodkin M, Sher E, et al. Species selectivity of a nicotinic acetylcholine receptor agonist is conferred by two adjacent extracellular β4 amino acids that are implicated in the coupling of binding to channel gating. Mol Pharmacol. 2007;71:389-97. [35] Mantione E, Micheloni S, Alcaino C, New K, Mazzaferro S, Bermudez I. Allosteric modulators of α4β2 nicotinic acetylcholine receptors: a new direction for antidepressant drug discovery. Future Med Chem. 2012;4:2217-30.

14

[36] Faghih R, Gfesser GA, Gopalakrishnan M. Advances in the discovery of novel positive allosteric modulators of the α7 nicotinic acetylcholine receptor. Recent Pat CNS Drug Discov. 2007;2:99-106. [37] Faghih R, Gopalakrishnan M, Briggs CA. Allosteric modulators of the α7 nicotinic acetylcholine receptor. J Med Chem. 2008;51:701-12. [38] Lightfoot AP, Kew JN, Skidmore J. α7 nicotinic acetylcholine receptor agonists and positive allosteric modulators. Prog Med Chem. 2008;46:131-71. [39] Gündisch D, Eibl C. Nicotinic acetylcholine receptor ligands, a patent review (20062011). Expert Opin Ther Pat. 2011;21:1867-96. [40] Mazurov AA, Speake JD, Yohannes D. Discovery and development of α7 nicotinic acetylcholine receptor modulators. J Med Chem. 2011;54:7943-61. [41] Romanelli MN, Gualtieri F. Cholinergic nicotinic receptors: competitive ligands, allosteric modulators, and their potential applications. Med Res Rev. 2003;23:393-426. [42] Pandya AA, Yakel JL. Effects of neuronal nicotinic acetylcholine receptor allosteric modulators in animal behavior studies. Biochem Pharmacol. 2013;86:1054-62. [43] Uteshev VV. The therapeutic promise of positive allosteric modulation of nicotinic receptors. Eur J Pharmacol. 2014;727:181-5. [44] Haydar SN, Dunlop J. Neuronal nicotinic acetylcholine receptors - targets for the development of drugs to treat cognitive impairment associated with schizophrenia and Alzheimer's disease. Curr Top Med Chem. 2010;10:144-52. [45] Umana IC, Daniele CA, McGehee DS. Neuronal nicotinic receptors as analgesic targets: it's a winding road. Biochem Pharmacol. 2013;86:1208-14. [46] Nirogi R, Goura V, Abraham R, Jayarajan P. α4β2* neuronal nicotinic receptor ligands (agonist, partial agonist and positive allosteric modulators) as therapeutic prospects for pain. Eur J Pharmacol. 2013;712:22-9. [47] Ambrosi P, Becchetti A. Targeting neuronal nicotinic receptors in cancer: new ligands and potential side-effects. Recent Pat Anticancer Drug Discov. 2013;8:38-52. [48] Malysz J, Grønlien JH, Timmermann DB, Håkerud M, Thorin-Hagene K, Ween H, et al. Evaluation of α7 nicotinic acetylcholine receptor agonists and positive allosteric modulators using the parallel oocyte electrophysiology test station. Assay Drug Dev Technol. 2009;7:374-90. [49] Grønlien JH, Håkerud M, Ween H, Thorin-Hagene K, Briggs CA, Gopalakrishnan M, et al. Distinct profiles of α7 nAChR positive allosteric modulation revealed by structurally diverse chemotypes. Mol Pharmacol. 2007;72:715-24. [50] Hurst RS, Hajós M, Raggenbass M, Wall TM, Higdon NR, Lawson JA, et al. A novel positive allosteric modulator of the α7 neuronal nicotinic acetylcholine receptor: in vitro and in vivo characterization. J Neurosci. 2005;25:4396-405. [51] Vernino S, Amador M, Luetje CW, Patrick J, Dani JA. Calcium modulation and high calcium permeability of neuronal nicotinic acetylcholine receptors. Neuron. 1992;8:127-34. [52] Mulle C, Lena C, Changeux JP. Potentiation of nicotinic receptor response by external calcium in rat central neurons. Neuron. 1992;8:937-45. [53] Krause RM, Buisson B, Bertrand S, Corringer P-J, Galzi J-L, Changeux J-P, et al. Ivermectin: a positive allosteric effector of the α7 neuronal nicotinic acetylcholine receptor. Mol Pharmacol. 1998;53:283-94. [54] Lopes C, Pereira FR, Wu H-Q, Purushottamachar P, Njar V, Schwarcz R, et al. Competitive antagonism between the nicotinic allosteric potentiating ligand galantamine and kynurenic acid at α7* nicotinic receptors. J Pharm Exp Ther. 2007;322:48-58. [55] Zwart R, De Filippi G, Broad LM, McPhie GI, Pearson KH, Baldwinson T, et al. 5Hydroxyindole potentiates human α7 nicotinic receptor-mediated responses and enhances

15

acetylcholine-induced glutamate release in cerebellar slices. Neuropharmacol. 2002;43:37484. [56] Chimienti F, Hogg RC, Plantard L, Lehmann C, Brakch N, Fischer J, et al. Identification of SLURP-1 as an epidermal neuromodulator explains the clinical phenotype of Mal de Meleda. Hum Mol Genet. 2003;12:3017-24. [57] Conroy WG, Liu Q-S, Nai Q, Margiotta JF, Berg DK. Potentiation of α7-containing nicotinic acetylcholine receptors by select albumins. Mol Pharmacol. 2003;63:419-28. [58] Thomsen MS, El-Sayed M, Mikkelsen JD. Differential immediate and sustained memory enhancing effects of α7 nicotinic receptor agonists and allosteric modulators in rats. PLOS ONE. 2011;6:e27014. [59] McLean SL, Idris NF, Grayson B, Gendle DF, Mackie C, Lesage AS, et al. PNU120596, a positive allosteric modulator of α7 nicotinic acetylcholine receptors, reverses a subchronic phencyclidine-induced cognitive deficit in the attentional set-shifting task in female rats. J Psychopharmacol. 2012;26:1265-70. [60] Callahan PM, Hutchings EJ, Kille NJ, Chapman JM, Terry AV. Positive allosteric modulator of α7 nicotinic-acetylcholine receptors, PNU-120596 augments the effects of donepezil on learning and memory in aged rodents and non-human primates. Neuropharmacol. 2013;67:201-12. [61] Freitas K, Negus SS, Carroll FI, Damaj MI. In vivo pharmacological interactions between a type II positive allosteric modulator of α7 nicotinic ACh receptors and nicotinic agonists in a murine tonic pain model. Br J Pharmacol. 2013;169:567-79. [62] Sun F, Jin K, Uteshev VV. A type-II positive allosteric modulator of α7 nAChRs reduces brain injury and improves neurological function after focal cerebral ischemia in rats. PLOS ONE. 2013;8:e73581. [63] Kalappa BI, Sun F, Johnson SR, Jin K, Uteshev VV. A positive allosteric modulator of α7 nAChRs augments neuroprotective effects of endogenous nicotinic agonists in cerebral ischaemia. Br J Pharmacol. 2013;169:1862-78. [64] Faghih R, Gopalakrishnan SM, Gronlien JH, Malysz J, Briggs CA, Wetterstrand C, et al. Discovery of 4-(5-(4-chlorophenyl)-2-methyl-3-propionyl-1H-pyrrol-1yl)benzenesulfonamide (A-867744) as a novel positive allosteric modulator of the alpha7 nicotinic acetylcholine receptor. J Med Chem. 2009;52:3377-84. [65] Malysz J, Grønlien JH, Anderson DJ, Håkerud M, Thorin-Hagene K, Ween H, et al. In vitro pharmacological characterization of a novel allosteric modulator of the α7 neuronal acetylcholine receptor, 4-(5-(4-chlorophenyl)-2-methyl-3-propionyl-1H-pyrrol-1yl)benzenesulfonamide (A-867744), exhibiting unique pharmacological profile. J Pharm Exp Ther. 2009;330:257-67. [66] Gill JK, Chatzidaki A, Ursu D, Sher E, Millar NS. Contrasting properties of α7-selective orthosteric and allosteric agonists examined on native nicotinic acetylcholine receptors. PLOS ONE. 2013;8:e55047. [67] Chatzidaki A, Fouillet A, Li J, Dage J, Millar NS, Sher E, et al. Pharmacological characterisation of nicotinic acetylcholine receptors expressed in human iPSC-derived neurons. PLOS ONE. 2015;10:e0125116. [68] Gill JK, Dhankher P, Sheppard TD, Sher E, Millar NS. A series of α7 nicotinic acetylcholine receptor allosteric modulators with close chemical similarity but diverse pharmacological properties. Mol Pharmacol. 2012;81:710-8. [69] Ondrejcak T, Wang Q, Kew JN, Virley DJ, Upton N, Anwyl R, et al. Activation of α7 nicotinic acetylcholine receptors persistently enhances hippocampal synaptic transmission and prevents Aß-mediated inhibition of LTP in the rat hippocampus. Eur J Pharmacol. 2012;677:63-70.

16

[70] Timmermann DB, Grønlien JH, Kohlhaas KL, Nielsen EØ, Dam E, Jørgensen TD, et al. An allosteric modulator of the α7 nicotinic acetylcholine receptor possessing cognitionenhancing properties in vivo. J Pharm Exp Ther. 2007;323:294-307. [71] Ng HJ, Whittemore ER, Tran MB, Hogenkamp DJ, Broide RS, Johnstone TB, et al. Nootropic α7 nicotinic receptor allosteric modulator derived from GABAA receptor modulators. Proc Natl Acad Sci USA. 2007;104:8059-64. [72] Broad LM, Zwart R, Pearson KH, Lee M, Wallace L, McPhie GI, et al. Identification and pharmacological profile of a new class of selective nicotinic acetylcholine receptor potentiators. J Pharm Exp Ther. 2006;318:1108-17. [73] Dunlop J, Lock T, Jow B, Sitzia F, Grauer S, Jow F, et al. Old and new pharmacology: positive allosteric modulation of the α7 nicotinic acetylcholine receptor by the 5hydroxytryptamine2B/C receptor antagonist SB-206553 (3,5-dihydro-5-methyl-N-3pyridinylbenzo[1,2-b:4,5-b']di pyrrole-1(2H)-carboxamide). J Pharm Exp Ther. 2009;328:766-76. [74] Dinklo T, Shaban H, Thuring JW, Lavreysen H, Stevens KE, Zheng L, et al. Characterization of 2-[[4-fluoro-3-(trifluoromethyl)phenyl]amino]-4-(4-pyridinyl)-5thiazolethanol (JNJ-1930942), a novel positive allosteric modulator of the α7 nicotinic acetylcholine receptor. J Pharm Exp Ther. 2011;336:560-74. [75] Chatzidaki A, D'Oyley JM, Gill-Thind JK, Sheppard TD, Millar NS. The influence of allosteric modulators and transmembrane mutations on desensitisation and activation of α7 nicotinic acetylcholine receptors. Neuropharmacol. 2015;97:75-85. [76] Sahdeo S, Wallace T, Hirakawa R, Knoflach F, Bertrand D, Maag H, et al. Characterization of RO5126946, a Novel α7 nicotinic acetylcholine receptor-positive allosteric modulator. J Pharm Exp Ther. 2014;350:455-68. [77] Rollema H, Coe JW, Chambers LK, Hust RS, Stahl SM, Williams KE. Rationale, pharmacology and clinical efficacy of partial agonists of α4β2 nACh receptors for smoking cessation. Trends Pharmacol Sci. 2007;28:316-25. [78] Paradiso K, Zhang J, Steinbach JH. The C terminus of the human nicotinic alpha4beta2 receptor forms a binding site required for potentiation by an estrogenic steroid. J Neurosci. 2001;21:6561-8. [79] Curtis L, Buisson B, Bertrand S, Bertrand D. Potentiation of human α4β2 neuronal nicotinic acetylcholine receptor by estradiol. Mol Pharmacol. 2002;61:127-35. [80] Jin X, Steinbach JH. A portable site: a binding element for 17β-estradiol can be placed on any subunit of a nicotinic α4β2 receptor. J Neurosci. 2011;31:5045-54. [81] Samochocki M, Höffle A, Fehrenbacher A, Jostock R, Ludwig J, Vhristner C, et al. Galantamine is an allosterically potentiating ligand of neuronal nicotinic but not of muscarinic acetylcholine receptors. J Pharm Exp Ther. 2003;305:1024-36. [82] Kim JS, Padnya A, Weltzin M, Edmonds BW, Schulte MK, Glennon RA. Synthesis of desformylflustrabromine and its evaluation as an α4β2 and α7 nACh receptor modulator. Bioorg Med Chem Lett. 2007;17:4855-60. [83] Sala F, Mulet J, Reddy KP, Bernal JA, Wikman P, Valor LM, et al. Potentiation of human α4β2 neuronal nicotinic receptors by a Flustra foliacea metabolite. Neurosci Lett. 2005;373:144-9. [84] Zwart R, Vijverberg HP. Potentiation and inhibition of neuronal nicotinic receptors by atropine: competitive and noncompetitive effects. Mol Pharmacol. 1997;52:886-95. [85] Smulders CJ, Zwart R, Bermudez I, van Kleef RG, Groot-Kormelink PJ, Vijverberg HP. Cholinergic drugs potentiate human nicotinic alpha4beta2 acetylcholine receptors by a competitive mechanism. Eur J Pharmacol. 2005;509:97-108.

17

[86] Moroni M, Zwart R, Sher E, Cassels BK, Bermudez I. α4β2 nicotinic receptors with high and low acetylcholine sensitivity: pharmacology, stoichiometry, and sensitivity to long-term exposure to nicotine. Mol Pharmacol. 2006;70:755-68. [87] Zwart R, Vijverberg HPM. Four pharmacologically distinct subtypes of α4β2 nicotinic acetylcholine receptor expressed in Xenopus laevis oocytes. Mol Pharmacol. 1998;54:112431. [88] Buisson B, Bertrand D. Chronic exposure to nicotine upregulates the human α4β2 nicotinic acetylcholine receptor function. J Neurosci. 2001;21:1819-29. [89] Nelson ME, Kuryatov A, Choi CH, Zhou Y, Lindstrom J. Alternate stoichiometries of α4β2 nicotinic acetylcholine receptors. Mol Pharmacol. 2003;63:332-41. [90] Lee CH, Zhu C, Malysz J, Campbell T, Shaughnessy T, Honore P, et al. α4β2 neuronal nicotinic receptor positive allosteric modulation: an approach for improving the therapeutic index of α4β2 nAChR agonists in pain. Biochem Pharmacol. 2011;82:959-66. [91] Zhu CZ, Chin CL, Rustay NR, Zhong C, Mikusa J, Chandran P, et al. Potentiation of analgesic efficacy but not side effects: co-administration of an α4β2 neuronal nicotinic acetylcholine receptor agonist and its positive allosteric modulator in experimental models of pain in rats. Biochem Pharmacol. 2011;82:967-76. [92] Timmermann DB, Sandager-Nielsen K, Dyhring T, Smith M, Jacobsen AM, Nielsen E, et al. Augmentation of cognitive function by NS9283, a stoichiometry-dependent positive allosteric modulator of α2- and α4-containing nicotinic acetylcholine receptors. Br J Pharmacol. 2012;167:164-82. [93] Olsen JA, Kastrup JS, Peters D, Gajhede M, Balle T, Ahring PK. Two distinct allosteric binding sites at α4β2 nicotinic acetylcholine receptors revealed by NS206 and NS9283 give unique insights to binding activity-associated linkage at Cys-loop receptors. J Biol Chem. 2013;288:35997-6006. [94] Grupe M, Jensen AA, Ahring PK, Christensen JK, Grunnet M. Unravelling the mechanism of action of NS9283, a positive allosteric modulator of (α4)3(β2)2 nicotinic ACh receptors. Br J Pharmacol. 2013;168:2000-10. [95] Weltzin MM, Huang Y, Schulte MK. Allosteric modulation of α4β2 nicotinic acetylcholine receptors by HEPES. Eur J Pharmacol. 2014;732:159-68. [96] Wu TY, Smith CM, Sine SM, Levandoski MM. Morantel allosterically enhances channel gating of neuronal nicotinic acetylcholine α3β2 receptors. Mol Pharmacol. 2008;74:466-75. [97] Bürgi JJ, Awale M, Boss SD, Schaer T, Marger F, Viveros-Paredes JM, et al. Discovery of potent positive allosteric modulators of the α3β2 nicotinic acetylcholine receptor by a chemical space walk in ChEMBL. ACS Chem Neurosci. 2014;5:346-59. [98] Levandoski MM, Piket B, Chang J. The anthelmintic levamisole is an allosteric modulator of human neuronal nicotinic acetylcholine receptors. Eur J Pharmacol. 2003;471:920. [99] Hsiao B, Dweck D, Luetje CW. Subunit-dependent modulation of neuronal nicotinic receptors by zinc. J Neurosci. 2001;21:1848-56. [100] Vázquez-Gómez E, García-Colunga J. Neuronal nicotinic acetylcholine receptors are modulated by zinc. Neuropharmacol. 2009;56:1035-40. [101] Hsiao B, Mihalak KB, Repicky SE, Everhart D, Mederos AH, Malhotra A, et al. Determinants of zinc potentiation on the α4 subunit of neuronal nicotinic receptors. Mol Pharmacol. 2006;69:27-36. [102] Moroni M, Vijayan R, Carbone A-L, Zwart R, Biggin PC, Bermudez I. 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. 2008;28:6884-94.

18

[103] Arias HR. Binding sites for exogenous and endogenous non-competitive inhibitors of the nicotinic acetylcholine receptor. Biochim Biophys Acta. 1998;1376:173-220. [104] Arias HR. Noncompetitive inhibition of nicotinic acetylcholine receptors by endogenous molecules. J Neurosci Res. 1998;52:369-79. [105] Papke RL, Sanberg PR, Shytle RD. Analysis of mecamylamine stereoisomers on human nicotinic receptor subtypes. J Pharm Exp Ther. 2001;297:646-56. [106] Varanda WA, Aracava Y, Sherby SM, VanMeter WG, Eldefrawi ME, Albuquerque EX. The acetylcholine receptor of the neuromuscular junction recognizes mecamylamine as a noncompetitive antagonist. Mol Pharmacol. 1985;28:128-37. [107] Banerjee S, Punzi JS, Kreilick K, Abood LG. [3H]mecamylamine binding to rat brain membranes. Studies with mecamylamine and nicotine analogues. Biochem Pharmacol. 1990;40:2105-10. [108] Martin TJ, Suchocki J, May EL, Martin BR. Pharmacological evaluation of the antagonism of nicotine's central effects by mecamylamine and pempidine. J Pharm Exp Ther. 1990;254:45-51. [109] Connolly J, Boulter J, Heinemann SF. α4-2 β2 and other nicotinic acetylcholine receptor subtypes as targets of psychoactive and addictive drugs. Br J Pharmacol. 1992;105:657-66. [110] Rozental R, Scoble GT, Albuquerque EX, Idriss M, Sherby S, Sattelle DB, et al. Allosteric inhibition of nicotinic acetylcholine receptors of vertebrates and insects by philanthotoxin. J Pharm Exp Ther. 1989;249:123-30. [111] Zhang L, Oz M, Stewart RR, Peoples RW, Weight FF. Volatile general anaesthetic actions on recombinant nAChα7 5-HT3 and chimeric nAChα7-5-HT3 receptors expressed in Xenopus oocytes. Brit J Pharmacol. 1997;120:353-5. [112] Gumilar F, Arias HR, Spitzmaul G, Bouzat C. Molecular mechanisms of inhibition of nicotinic acetylcholine receptors by tricyclic antidepressants. Neuropharmacol. 2003;45:96476. [113] Fryer JD, Lukas RJ. Antidepressants noncompetitively inhibit nicotinic acetylcholine receptor function. J Neurochem. 1999;72:1117-24. [114] Fryer JD, Lukas RJ. Noncompetitive functional inhibition at diverse, human nicotinic acetylcholine receptor subtypes by bupropion, phencyclidine, and ibogaine. J Pharm Exp Ther. 1999;288:88-92. [115] Hans M, Wilhelm M, Swandulla D. Menthol suppresses nicotinic acetylcholine receptor functioning in sensory neurons via allosteric modulation. Chem Senses. 2012;37:463-9. [116] Ashoor A, Nordman JC, Veltri D, Yang K-HS, Kury LA, Shuba Y, et al. Menthol binding an inhibition of α7-nicotinic acetylcholine receptors. PLOS ONE. 2013;8:e67674. [117] Jayakar SS, Dailey WP, Eckenhoff RG, Cohen JB. Identification of propofol binding sites in a nicotinic acetylcholine receptor with a photoreactive propofol analog. J Biol Chem. 2013;288:6178-89. [118] Park TJ, Seo HK, Kang BJ, Kim KT. Noncompetitive inhibition by camphor of nicotinic acetylcholine receptors. Biochem Pharmacol. 2001;61:787-93. [119] Ton HT, Smart AE, Aguilar BL, Olson TT, Kellar KJ, Ahern GP. Menthol Enhances the Desensitization of Human α3β4 Nicotinic Acetylcholine Receptors. Mol Pharmacol. 2015;doi:10.1124/mol.115.098285. [120] Hoda JC, Krause R, Bertrand S, Bertrand D. Unexpected sensitivity of the human alpha7 neuronal nicotinic acetylcholine receptor to aminoglycosides. Neuroreport. 2006;17:65-70. [121] Singhal SK, Zhang L, Morales M, Oz M. Antipsychotic clozapine inhibits the function of α7-nicotinic acetylcholine receptors. Neuropharmacol. 2007;52:387-94.

19

[122] Arias HR, Bhumireddy P, Spitzmaul G, Trudell JR, Bouzat C. Molecular mechanisms and binding site location for the noncompetitive antagonist crystal violet on nicotinic acetylcholine receptors. Biochem. 2006;45:2014-26. [123] Palma E, Maggi L, Miledi R, Eusebi F. Effects of Zn2+ on wild and mutant neuronal α7 nicotinic receptors. Proc Natl Acad Sci U S A. 1998;95:10246-50. [124] Oz M, Spivak CE, Lupica CR. The solubilizing detergents, Tween 80 and Triton X-100 non-competitively inhibit alpha 7-nicotinic acetylcholine receptor function in Xenopus oocytes. J Neurosci Methods. 2004;137:167-73. [125] Pereira EFR, Hilmas C, Santos MD, Alkondon M, Maelicke A, Albuquerque EX. Unconventional ligands and modulators of nicotinic receptors. J Neurobiol. 2002;53:479-500. [126] Ibañez-Tallon I, Miwa JM, Wang HL, Adams NC, Crabtree GW, Sine SM, et al. Novel modulation of neuronal nicotinic acetylcholine receptors by association with the endogenous prototoxin lynx1. Neuron. 2002;33:893-903. [127] Miwa JM, Stevens TR, King SL, Caldarone BJ, Ibanez-Tallon I, Xiao C, et al. The prototoxin lynx1 acts on nicotinic acetylcholine receptors to balance neuronal activity and survival in vivo. Neuron. 2006;51:587-600. [128] Oz M. Receptor-independent effects of endocannabinoids on ion channels. Curr Pharm Des. 2006;12:227-39. [129] Noblin DJ, Bertekap RL, Burford NT, Hendricson A, Zhang L, Knox R, et al. Development of a high-throughput calcium flux assay for identification of all ligand types including positive, negative, and silent allosteric modulators for G protein-coupled receptors. Assay Drug Dev Tech. 2012;10:457-67. [130] Pedersen SE, Sharp SD, Liu WS, Cohen JB. Structure of the noncompetitive antagonistbinding site of the Torpedo nicotinic acetylcholine receptor. [3H]meproadifen mustard reacts selectively with alpha-subunit Glu-262. J Biol Chem. 1992;267:10489-99. [131] Middleton RE, Strnad NP, Cohen JB. Photoaffinity labeling the torpedo nicotinic acetylcholine receptor with [(3)H]tetracaine, a nondesensitizing noncompetitive antagonist. Mol Pharmacol. 1999;56:290-9. [132] Chiara DC, Dangott LJ, Eckenhoff RG, Cohen JB. Identification of nicotinic acetylcholine receptor amino acids photolabeled by the volatile anesthetic halothane. Biochemistry. 2003;42:13457-67. [133] Ziebell MR, Nirthanan S, Husain SS, Miller KW, Cohen JB. Identification of binding sites in the nicotinic actylcholine receptor for [3H]azietomidate, a photoactivatable general anesthetic. J Biol Chem. 2004;279:17640-9. [134] Nirthanan S, Garcia G, Chiara DC, Husain SS, Cohen JB. Identification of binding sites in the nicotinic acetylcholine receptor for TDBzl-etomidate, a photoreactive positive allosteric effector. J Biol Chem. 2008;283:22051-62. [135] Hamouda AK, Jayakar SS, Chiara DC, Cohen JB. Photoaffinity labeling of nicotinic receptors: diversity of drug binding sites! J Mol Neurosci. 2014;53:480-6. [136] Hamouda AK, Chiara DC, Sauls D, Cohen JB, Blanton MP. Cholesterol interacts with transmembrane alpha-helices M1, M3, and M4 of the Torpedo nicotinic acetylcholine receptor: photolabeling studies using [3H]Azicholesterol. Biochemistry. 2006;45:976-86. [137] Brannigan G, Hénin J, Law R, Eckenhoff R, Klein ML. Embedded cholesterol in the nicotinic acetylcholine receptor. Proc Natl Acad Sci U S A. 2008;105:14418-23. [138] Millar NS. A review of experimental techniques used for the heterologous expression of nicotinic acetylcholine receptors. Biochem Pharmacol. 2009;78:766-76. [139] Young GT, Zwart R, Walker AS, Sher E, Millar NS. Potentiation of α7 nicotinic acetylcholine receptors via an allosteric transmembrane site. Proc Natl Acad Sci USA. 2008;105:14686-91.

20

[140] Collins T, Young GT, Millar NS. Competitive binding at a nicotinic receptor transmembrane site of two α7-selective positive allosteric modulators with differing effects on agonist-evoked desensitization. Neuropharmacol. 2011;61:1306-13. [141] Bertrand D, Bertrand S, Casser S, Gubbins E, Li J, Gopalakrishnan M. Positive allosteric modulation of the α7 nicotinic acetylcholine receptor: ligand interactions with distinct binding sites and evidence for a prominent role of the M2-M3 segment. Mol Pharmacol. 2008;74:1407-16. [142] Sattelle DB, Buckingham SD, Akamatsu M, Matsuda K, Pienarr I, Jones AK, et al. Comparative pharmacology and computational modelling yields insights into allosteric modulation of human α7 nicotinic acetylcholine receptors. Biochem Pharmacol. 2009;78:83643. [143] Nury H, Van Renterghem C, Weng Y, Tran A, Baaden M, Dufresne V, et al. X-ray structures of general anaesthetics bound to a pentameric ligand-gated ion channel. Nature. 2011;469:428-31. [144] Ashoor A, Nordman JC, Veltri D, Yang K-HS, Shuba Y, Kury LA, et al. Menthol inhibits 5-HT3 receptor-metiated currents. J Pharm Exp Ther. 2013;247:398-409. [145] Yip GMS, Chen Z-W, Edge CJ, Smith CH, Dickinson R, Hohenester E, et al. A propofol binding site on mammalian GABAA receptors identified by photolabeling. Nature Chem Biol. 2013;9:715-20. [146] Lynagh T, Laube B. Opposing effects of the anesthetic propofol at pentameric ligandgated ion channels mediated by a common site. J Neurosci. 2014;34:2155-9. [147] Lansdell SJ, Sathyaprakash C, Doward A, Millar NS. Activation of human 5hydroxytryptamine type 3 receptors via an allosteric transmembrane site. Mol Pharmacol. 2015;87:87-95. [148] Oz M, Zhang L, Ravindran A, Morales M, Lupica CR. Differential effects of endogenous and synthetic cannabinoids on α7-nicotinic acetylcholine receptor-mediated responses in Xenopus oocytes. J Pharm Exp Ther. 2004;310:1152-11160. [149] Oz M, Jackson SN, Woods AS, Morales M, Zhang L. Additive effects of endogenous cannabinoid anandamide and ethanol on α7-nicotinic acetylcholine receptor-mediated responses in Xenopus oocytes. J Pharm Exp Ther. 2005;313:1272-80. [150] Jackson SN, Singhal SK, Woods AS, Morales M, Shippenberg T, Zhang L, et al. Volatile anesthetics and endogenous cannabinoid anandamide have additive and independent inhibitory effects on α7-nicotinic acetylcholine receptor-mediated responses in Xenopus oocytes. Eur J Pharmacol. 2008;582:42-51. [151] Sadek B, Khanian SS, Ashoor A, Prytkova T, Ghattas MA, Atatreh N, et al. Effects of antihistamines on the function of human α7-nicotinic acetylcholine receptors. Eur J Pharmacol. 2015;746:308-16. [152] Khiroug SS, Harkness PC, Lamb PW, Sudweeks SN, Khiroug L, Millar NS, et al. Rat nicotinic ACh receptor α7 and β2 subunits co-assemble to form functional heteromeric nicotinic receptor channels. J Physiol. 2002;540:425-34. [153] Mowrey DD, Liu Q, Bondarenko V, Chen Q, Seyoum E, Xu Y, et al. Insights into distinct modulation of α7 and α7β2 nicotinic acetylcholine receptors by the volatile anesthetic isoflurane. J Biol Chem. 2013;288:35793-800. [154] Olsen RW, Li GD, Wallner M, Trudell JR, Bertaccini EJ, Lindahl E, et al. Structural models of ligand-gated ion channels: sites of action for anesthetics and ethanol. Alcohol Clin Exp Res. 2014;38:595-603. [155] Collins T, Millar NS. Nicotinic acetylcholine receptor transmembrane mutations convert ivermectin from a positive to a negative allosteric modulator. Mol Pharmacol. 2010;78:198-204.

21

[156] Hibbs RE, Gouaux E. Principles of activation and permeation in an anion-selective Cysloop receptor. Nature. 2011;474:54-60. [157] Kirst HA. The spinosyn family of insecticides: realizing the potential of natural products research. J Antibiot (Tokyo). 2010;63:101-11. [158] Puinean AM, Lansdell SJ, Collins T, Bielza P, Millar NS. A nicotinic acetylcholine receptor transmembrane point mutation (G275E) associated with resistance to spinosad in Frankliniella occidentalis. J Neurochem. 2013;124:590-601. [159] Forman SA, Miller KW. Anesthetic sites and allosteric mechanisms of action on Cysloop ligand-gated ion channels. Can J Anaesth. 2011;58:191-205. [160] Seo S, Henry JT, Lewis AH, Wang N, Levandoski MM. The positive allosteric modulator morantel binds at noncanonical subunit interfaces of neuronal nicotinic acetylcholine receptors. J Neurosci. 2009;29:8734-42. [161] Cesa LC, Higgins CA, Sando SR, Kuo DW, Levandoski MM. Specificity determinants of allosteric modulation in the neuronal nicotinic acetylcholine receptor: a fine line between inhibition and potentiation. Mol Pharmacol. 2012;81:239-49. [162] Short CA, Cao AT, Wingfield MA, Doers ME, Jobe EM, Wang N, et al. Subunit interfaces contribute differently to activation and allosteric modulation of neuronal nicotinic acetylcholine receptors. Neuropharmacol. 2015;91:157-68. [163] Weltzin MM, Schulte MK. Desformylflustrabromine modulates α4β2 nAChR high- and low- sensitivity isoforms at allosteric clefts containing the β2 subunit. J Pharm Exp Ther. 2015;doi:10.1124/jpet.115.223933. [164] Olsen JA, Ahring PK, Kastrup JS, Gajhede M, Balle T. Structural and functional studies of the modulator NS9283 reveal agonist-like mechanism of action at α4β2 nicotinic acetylcholine receptors. J Biol Chem. 2014;289:24911-21. [165] Wang J, Kuryatov A, Sriram A, Jin Z, Kamenecka TM, Kenny PJ, et al. An Accessory Agonist Binding Site Promotes Activation of α4β2* Nicotinic Acetylcholine Receptors. J Biol Chem. 2015;290:13907-18. [166] Mazzaferro S, Benallegue N, Carbone A, Gasparri F, Vijayan R, Biggin PC, et al. Additional acetylcholine (ACh) binding site at alpha4/alpha4 interface of (alpha4beta2)2alpha4 nicotinic receptor influences agonist sensitivity. J Biol Chem. 2011;286:31043-54. [167] Mazzaferro S, Gasparri F, New K, Alcaino C, Faundez M, Iturriaga Vasquez P, et al. Non-equivalent ligand selectivity of agonist sites in (α4β2)2α4 nicotinic acetylcholine receptors: a key determinant of agonist efficacy. J Biol Chem. 2014;289:21795-806. [168] Harpsøe K, Ahring PK, Christensen JK, Jensen ML, Peters D, Balle T. Unraveling the high- and low-sensitivity agonist responses of nicotinic acetylcholine receptors. J Neurosci. 2011;31:10759-66. [169] Eaton JB, Lucero LM, Stratton H, Chang Y, Cooper JF, Lindstrom JM, et al. The unique α4+/-α4 agonist binding site in (α4)3(β2)2 subtype nicotinic acetylcholine receptors permits differential agonist desensitization pharmacology versus the (α4)2(β2)3 subtype. J Pharm Exp Ther. 2014;348:46-58. [170] González-Cestari TF, Henderson BJ, Pavlovicz RE, McKay SB, El-Hajj RA, Pulipaka AB, et al. Effect of novel negative allosteric modulators of neuronal nicotinic receptors on cells expressing native and recombinant nicotinic receptors: implications for drug discovery. J Pharm Exp Ther. 2009;328:504-15. [171] Henderson BJ, Pavlovicz RE, Allen JD, González-Cestari TF, Orac CM, Bonnell AB, et al. Negative allosteric modulators that target human α4β2 neuronal nicotinic receptors. J Pharm Exp Ther. 2010;334:761-74.

22

[172] Pavlovicz RE, Henderson BJ, Bonnell AB, Boyd RT, McKay DB, Li C. Identification of a negative allosteric site on human α4β2 and α3β4 neuronal nicotinic acetylcholine receptors. PLOS ONE. 2011;6:e24949. [173] Maelicke A, Albuquerque EX. Allosteric modulation of nicotinic acetylcholine receptors as a treatment strategy for Alzheimer's disease. Eur J Pharmacol. 2000;393:165-70. [174] Ludwig J, Höffle-Maas A, Samochocki M, Luttmann E, Albuquerque EX, Fels G, et al. Localization by site-directed mutagenesis of a galantamine binding site on α7 nicotinic acetylcholine receptor extracellular domain. J Recept Signal Transduct Res. 2010;30:469-83. [175] Luttmann E, Ludwig J, Höffle-Maas A, Samochocki M, Maelicke A, Fels G. Structural model for the binding sites of allosterically potentiating ligands on nicotinic acetylcholine receptors. ChemMedChem. 2009;4:1874-82. [176] Hansen SB, Taylor P. Galanthamine and non-competitive inhibitor binding to AChbinding protein: evidence for a binding site on non-α-subunit interfaces of heteromeric neuronal nicotinic receptors. J Mol Biol. 2007;369:895-901. [177] Galzi JL, Bertrand S, Corringer PJ, Changeux JP, Bertrand D. Identification of calcium binding sites that regulate potentiation of a neuronal nicotinic acetylcholine receptor. EMBO J. 1996;15:5824-32. [178] Calimet N, Simoes M, Changeux JP, Karplus M, Taly A, Cecchini M. A gating mechanism of pentameric ligand-gated ion channels. Proc Natl Acad Sci USA. 2013;110:E3987-96.

Figure captions Fig. 1. Cartoon representation of nAChR structure and subunit topology. Five co-assembled nAChR subunits are shown in a lipid bilayer and annotated to show locations that have been proposed as binding sites for allosteric modulators (A). The transmembrane topology of a single nAChR subunit is shown, in which the polypeptide chain is denoted by a blue line and the four transmembrane helices by cylinders (A and B). Also illustrated (C) is the nAChR transmembrane domain, formed by twenty transmembrane helices (four from each of the five co-assembled subunits). The central ion channel pore is lined by the second transmembrane helix (TM2) from each of the five subunits. Fig. 2. Representation of the equilibrium between closed, open and desensitised states of a nAChR. The open state is stabilised by the binding of an agonist, such as the orthosteric agonist acetylcholine (depicted as blue dots). In addition, there is evidence that receptors can enter an agonist-bound, non-conducting, desensitised state. Although only three states are illustrated in this figure, it should be noted that receptors are expected to transition through multiple intermediate states. There is evidence, for example of both short-lived and longerlasting desensitised states. As discussed in the text, the transition between these various states can be influenced by allosteric modulators interacting with a variety of receptor binding sites.

23

Fig. 3. Chemical structures of selected nAChR allosteric modulators.

Fig. 1

Fig. 2

Fig. 3

24

Table 1 Experimental techniques that have helped to establish the location of binding sites for nAChR allosteric modulators. (Note: References cited are a representative series, rather than a comprehensive list) Experimental Technique

References

Affinity labelling

[117, 130-136, 145, 159]

Computer-docking and molecular dynamics studies

[30, 117, 122, 137, 139, 140, 142, 172, 175, 178]

Subunit chimaeras

[29, 65, 72, 75, 79, 81, 93, 111, 139, 141, 147]

Site-directed mutagenesis (SDM)

[30, 33, 53, 68, 78, 80, 93, 95, 139, 140, 146, 147, 155, 158, 160-163, 172, 174, 175, 177]

Radioligand binding

[29, 33, 65, 70-72, 74-76, 117, 122, 147]

X-ray crystallography and NMR

[143, 153, 156, 176]

25

Allosteric modulation of nicotinic acetylcholine receptors.

Nicotinic acetylcholine receptors (nAChRs) are receptors for the neurotransmitter acetylcholine and are members of the 'Cys-loop' family of pentameric...
1KB Sizes 1 Downloads 27 Views