Skip to main content
eLife logoLink to eLife
. 2023 Jul 3;12:e86496. doi: 10.7554/eLife.86496

Agonist efficiency links binding and gating in a nicotinic receptor

Dinesh C Indurthi 1, Anthony Auerbach 1,
Editors: László Csanády2, Kenton J Swartz3
PMCID: PMC10317499  PMID: 37399234

Abstract

Receptors signal by switching between resting (C) and active (O) shapes (‘gating’) under the influence of agonists. The receptor’s maximum response depends on the difference in agonist binding energy, O minus C. In nicotinic receptors, efficiency (η) represents the fraction of agonist binding energy applied to a local rearrangement (an induced fit) that initiates gating. In this receptor, free energy changes in gating and binding can be interchanged by the conversion factor η. Efficiencies estimated from concentration-response curves (23 agonists, 53 mutations) sort into five discrete classes (%): 0.56 (17), 0.51(32), 0.45(13), 0.41(26), and 0.31(12), implying that there are 5 C versus O binding site structural pairs. Within each class efficacy and affinity are corelated linearly, but multiple classes hide this relationship. η unites agonist binding with receptor gating and calibrates one link in a chain of coupled domain rearrangements that comprises the allosteric transition of the protein.

Research organism: Other

Introduction

The primary job of a receptor is to convert chemical energy from ligand binding into mechanical work of protein conformational change. Each agonist has two affinities that measure how strongly it binds to resting and active target sites, and an efficacy that measures how well it activates once bound (Cecchini and Changeux, 2022; Ehlert, 2015; Foreman et al., 2011). A third agonist property, efficiency (η; eta), is the correlation between efficacy and affinity and is the receptor’s output/input energy ratio (Nayak et al., 2019). Here, we describe and interpret the distribution of η values estimated from concentration(dose)-response curves (CRCs) of adult-type skeletal muscle nicotinic acetylcholine receptors (AChRs) activated by different agonists and having different binding site mutations.

AChRs are five subunit, ligand-gated ion channels that have two neurotransmitter sites in the extracellular domain (ECD), and a narrow equatorial gate in the lumen of the transmembrane domain (TMD) that governs the passive flow of cations across the membrane (Gharpure et al., 2020; Unwin, 1995; Zarkadas et al., 2022). AChRs are allosteric proteins that alternate spontaneously between global off (C, closed-channel) and on (O, open-channel) conformations in which each neurotransmitter site binds agonists weakly (with low affinity; LA) or strongly (with high affinity; HA) (Cecchini and Changeux, 2022; Jackson, 2006; Phillips, 2020). The AChR neurotransmitter sites and gate are separated by ~60 Å, but both regions change structure and function within the gating isomerization, CLA⇄OHA.

In adult-type AChRs and without any bound agonists, the probability of the O conformation (PO) is small (~10–6) so baseline current is negligible (Jackson, 1986; Nayak et al., 2012; Purohit and Auerbach, 2009). However, when both agonist sites are occupied by the neurotransmitter ACh, the increase in favorable binding free energy realized in the spontaneous isomerization from C to O increases PO substantially (to ~0.96), to generate a cellular current (Figure 1).

Figure 1. Bind and gate.

Figure 1.

(A) Potential energy surfaces (landscapes) for receptor activation without (left) and with (right) bound agonists (A). C, closed channel and low affinity (LA); O, open channel and high affinity (HA). Red, without agonists intrinsic C→O gating is uphill; blue lines, agonists increase PO because they bind more favorably to O; green, agonists also stabilize the gating transition state (‡) to increase the opening rate constant because the LA→HA induced fit (‘hold’, Figure 2) occurs at the start of the gating conformational change. Calibration: ACh, adult-type AChRs, –100 mV, 23 °C. (B) Corresponding reaction scheme (2 equivalent and independent neurotransmitter sites). Horizontal, ligand-protein complex formation; KdC and KdO, LA and HA equilibrium dissociation constants to C and O. Vertical, global conformational change; Ln, gating equilibrium constant with n bound agonists. Red dashed line, pathway that determines the CRC when L0 is negligible (Equation 4a, Equation 4b, Equation 4c, Equation 4d, Methods).

Our approach to investigating the AChR allosteric transition is to measure free energy changes associated with each component event. Energy and structure are related, and a longer term goal is to associate these energy changes with those in structure. Briefly, we use electrophysiology to compile a CRC and from the minimum, maximum and midpoint calculate the ligand’s LA and HA equilibrium dissociation constants for binding to C and to O (KdC and KdO). The logs of these are proportional to the LA and HA binding free energies, ΔGLA and ΔGHA. As shown below, agonist efficiency depends only on these two values.

Previously, it was observed that in AChRs the ratio ΔGLA/ΔGHA is approximately the same for a group of agonists despite wide variation in individual affinities and efficacies (Jadey and Auerbach, 2012). Within this group, whose members included full agonists (like ACh), partial agonists (like nicotine) and extremely weak agonists (like choline), the binding free energy ratio was 0.51±0.01 (mean ± sd). That is, for all members in the group binding to O is ~two-fold stronger than to C, regardless of agonist affinity or efficacy. Additional experiments revealed two more groups having a shared binding energy ratio, either 0.58 (for example, epibatidine) or 0.46 (for example, tetramethylammonium) (Indurthi and Auerbach, 2021; Nayak et al., 2019). In these groups, agonist binding free energy increases by ~1.7 and ~2.2-fold in the channel-opening conformational change.

The observation of a common ΔGLA/ΔGHA ratio for even one group of ligands is remarkable because it implies that the only 2 events in receptor activation that involve the agonist directly - LA binding to C and the switch to HA that happens within gating - are not independent. Rather, the constant ratio indicates that the energy (structure) changes that underpin LA binding to C and HA binding to O are related (Auerbach, 2016; Jadey and Auerbach, 2012; Purohit et al., 2014).

Below, we define η empirically as the receptor’s output/input energy ratio, and propose that in AChRs it is the fraction of agonist binding energy applied to the mechanical work of a local rearrangement of the neurotransmitter site (an induced fit) that initiates the gating conformational cascade (Nayak et al., 2019). In principle, η can be calculated from any CRC to estimate the degree of coupling between the maxima of receptor output (efficacy) and agonist input (affinity).

We measured η for various agonists of AChRs, both wild-type (wt) and following mutation of a binding site residue. Here we report 16 new values (shown in Tables 1 and 2) that, combined with 60 previous measurement describe a spectrum of 5 η classes. The presence of multiple η classes obscures the underlying correlations between affinity and efficacy and, further, suggests that there are multiple C versus O binding site structures. Importantly, the existence of efficiency classes highlights that binding and gating are energy-linked stages of a unified allosteric transition.

Table 1. Agonist efficiencies.

Table 1—source data 1. Agonist efficiency.
Agonist EC50 (μM) (sem) POmax (sem) KdC (μM) KdO (nM) c η n
BzTMAa 1070 (200) 0.60 (0.05) 930 (112.0) 650 (90) 1424 (18.0) 0.51 (0.01) 3
Decb 190 (20) 0.79 (0.03) 90 (7.0) 140 (10) 643 (21) 0.41 (0.01) 4
SCha 20 (3) 0.84 (0.02) 50 (4.0) 20 (1) 3177 (118) 0.45 (0.01) 3
BzTEAb+c 2 (0.1) 0.85 (0.02) 0.80 (0.3) 3 (2) 316 (14) 0.29 (0.00) 3
TriMAab 16000 (1200) 0.78 (0.03) 7720 (316.0) 12580 (670) 615 (18) 0.57 (0.01) 4

Mean EC50 and Pomax were measured from each CRC (intra-cluster interval duration histograms in Figure 3 and Figure 3—figure supplement 1), with standard error of mean (sem). KdC and KdO were calculated (Equation 4a) after correcting the background mutations that only changed L0. c, coupling constant (KdC/KdO); η, efficiency (Equation 2); error estimates (calculated by error propagation) for calculated KdC, KdO, c, and η values given by (sem); n, number of CRCs. Membrane potential, +70 mV (to minimize channel block by the agonist). Agonist structures are in Figure 5; superscripts indicate mutation backgrounds: aεS450W, bεL269F, cεE181W that increase Lo (increase responses to weak agonists).

Table 2. Mutation efficiencies.

Table 2—source data 1. Mutation efficiency.
mutation agonist EC50 (μM) POmax KdC (μM) KdO (nM) c ηmut n ηwt
D200A ACha 110 (10) 0.80 (0.02) 52 (3) 150 (5) 338 (12) 0.37 (0.00) 3 0.50
CCha 320 (70) 0.45 (0.02) 138 (18) 900 (94) 153 (3) 0.36 (0.00) 4 0.52
TMAa 13350 (2970) 0.37 (0.03) 5683 (740) 4382 (890) 130 (5) 0.48 (0.02) 5 0.54
Ebta+b 110 (20) 0.50 (0.03) 47 (5) 680 (47) 69 (2) 0.30 (0.00) 3 0.41
Ebxa+b 060 (20) 0.49 (0.03) 26 (5) 390 (58) 67 (3) 0.29 (0.01) 3 0.46
K145A ACha 110 (10) 0.95 (0.06) 83 (5) 50 (3) 1672 (82) 0.44 (0.00) 3 -
CCha 380 (30) 0.53 (0.01) 165 (8) 400 (15) 411 (5) 0.41 (0.00) 4 -
TMAa 2200 (590) 0.19 (0.01) 922 (125) 5010 (712) 184 (3) 0.43 (0.01) 3 -
Ebta 40 (4) 0.81 (0.04) 20 (2) 20 (2) 792 (52) 0.38 (0.00) 3 -
Ebxa 40 (3) 0.83 (0.01) 20 (1) 20 (1) 847 (17) 0.38 (0.00) 3 -
G153S Ebta 2 (0.5) 0.65 (0.01) 2 (0.3) 7 (1) 313 (4) 0.31 (0.01) 4 -

Measured EC50 & Pomax and calculated Kd, c and h, mean (sem). For mutation location see Figure 6, inset. KdC and KdO were calculated from CRC parameters (Figure 6) by using Eq. 4. c, coupling constant (KdC/KdO); n, number of CRCs. L0 was corrected for background mutations (Methods): aεS450W, bεL269F, cεE181W. See Figure 6—figure supplement 1 and Figure 6—figure supplement 2 for intra-cluster interval duration histograms.

Results

Background and definitions

The standard conception of receptor activation incorporates two seemingly disparate events – bind, the formation of a ligand-protein complex, and gate, the global isomerization of the protein (Figure 1). However, as described below (Figure 2), in AChRs these are composite reactions and are connected by a pair of local, induced-fit rearrangements of the agonist site (Jadey and Auerbach, 2012).

For clarity, we define the universal agonist attributes affinity and efficacy. Affinity is the strength at which the ligand binds to its target site. In receptors ligands have two affinities, weak binding to C and strong binding to O (Figure 1). The corresponding binding free energies, ΔGLA and ΔGHA, are calculated (in kcal/mol) as +RT times the natural logarithms of the apparent KdC and KdO, where R is the gas constant and T is the absolute temperature (RT = 0.59 at 23 °C). For ACh at adult-type binding sites KdC = 174 μM (ΔGLA = −5.1 kcal/mol) and KdO = 29 nM (ΔGHA = −10.2 kcal/mol) (Jadey and Auerbach, 2012).

Efficacy can be defined in several ways. Oftern, it is simply the high-concentration asymptote (maximum response) of an unnormalized CRC that in our experiments is POmax. Considering just bind and gate (Figure 1B), this limit depends only on the fully-liganded gating equilibrium constant L2 so this constant, too, defines agonist efficacy (Equation 4a). Another definition derives from considering the full cycle of receptor activation. In adult-type AChRs the 2 neurotransmitter binding sites are approximately equivalent and independent and there is no significant input of external energy (Nayak and Auerbach, 2017), so

L2L0=(KdCKdO)2. (1)

The subscripts of the gating equilibrium constants (L) refer to the number of bound agonists, and the equilibrium dissociation constant (Kd) ratio is called the coupling constant (c). L0 is agonist-independent so differences in L2 (efficacy) among agonists depend only on differences in c. Below, we use the logarithm of c (λ) as the index of relative agonist efficacy,

λ=ΔGHAΔGLA.

The relative efficacy of an agonist depends only on the difference between binding free energies, O minus C (blue lines in Figure 1A).

Unlike voltage and mechanical stimuli, a small, thermalized ligand can deliver only a small force to a large receptor (Howard, 2001). The tiny momentum imparted to the protein by the ligand is obscured by those from collisions with water molecules. In the absence of external energy, agonists promote conformational change only by providing more favorable (stabilizing) binding energy to active compared to resting states that interconvert spontaneously. This mechanism likely pertains to all large receptors activated by small agonists.

Efficiency (η) is defined empirically as the maximum output/input energy ratio (the efficacy/high-affinity energy ratio),

η=(ΔGHAΔGLA)ΔGHA=1ΔGLAΔGHA (2)

Accordingly, the agonist-dependent free energy changes in gating (ΔGHA-ΔGLA) and in binding (ΔGHA) are interchangeable, with η as the conversion factor. Again, efficacy (λ) is a free energy difference and the maximum amount the agonist can deliver to the receptor’s gating machinery, and efficiency (η) is a free energy ratio and this amount normalized by the ligand’s maximum binding energy. With regard to equilibrium dissociation constants, λ relates to log(KdC/KdO) and η relates to log(KdC)/log(KdO) (Figure 7—figure supplement 2).

Insofar as KdC and KdO are universal agonist attributes, so too is η. We propse that an η value can be calculated from the ratio of logarithms of the 2 dissociation constants for every agonist of every receptor. However, this does not imply that η has a physical meaning, or that all agonists of a given receptor have the same (or a unique) η value. Equation 2 only shows how to calculate η from the 2 equilibrium dissociation constants. We estimated these from CRCs, but other experimental approaches would suffice.

In AChRs η does have a physiochemical meaning because KdC and KdO derive mainly from a pair of induced fit rearrangements at the ligand site. Although bind and gate are usually denoted as single-step events (Figure 1B), in AChRs both are composite reactions that harbor intermediate states that are too short-lived to be detected individually and directly (Figure 2, Figure 2—figure supplement 1). Induced fits are common, and undetected intermediate states have been invoked with great success previosuly to explain experimental results (for example ES in enzymology, and AC in pharmacology). Here, we invoke intermediate states to deconstruct η (Figure 2).

Figure 2. Catch and hold.

(A) Bind and gate. Top, as one-step reactions. Bottom, as composite reactions. The main undetected intermediate states are AC, the encounter complex (inside bind) and ACHA, a high affinity and closed channel state (inside gate). Black, the 2 agonist-dependent induced fit rearrangements are called catch (AC⇄ACLA) and hold (ACLA⇄ACHA). Gray, diffusion (A+C⇄AC) and receptor conformation changes in other domains (ACHA⇄AO) are approximately agonist-independent (see Figure 2—figure supplement 1). (B) Catch-and-hold free energy landscape. An η class indicates that catch and hold are linked in a linear free energy relationship (LFER; dashed lines) (Howard, 2001). Green, weak agonist; black, strong agonist. Ligand binding energy ‘tilts’ the entire landscape to the total extent ΔGHA (black side bars). For relative agonist actions, the energy change in catch (brown) determines KdC and the energy change in hold (blue) determines the coupling constant. η is the fraction of the total applied to hold (blue/black) and η is the fraction apllied to catch (brown/black). Agonist affinity and efficacy differ substantially, but η is constant and depends only on the left-right position of ACLA in the reaction.

Figure 2.

Figure 2—figure supplement 1. Inside bind and gate.

Figure 2—figure supplement 1.

See Figure 1B for 2 equivalent sites. LA, low affinity; HA, high affinity. (Top). Standard activation scheme; bind and gate transitions bracket AC (a LA closed state, Delcastillo, et al., 1957). In gate, both channel conductance and binding site affinity increase (ACLA⇄AOHA). Middle. Expanded activation scheme showing undetected intermediates inside bind and gate. In bind, the agonist arrives at the target by diffusion (diffuse) and forms an ultra-LA AC encounter complex, followed by the first stage of the induced fit (catch1) that forms ACLA. In gate (boxed), the second stage of the rearrangement (hold) forms ACHA, followed by additional rearrangements in distant domains (etc) that eventually lead to AOHA. Experimental KdC and KdO values are dominated by energy changes in the two stages of the induced fit. Bottom. Expansion of gating. The distribution of Φ values2 suggests that the global isomerization involves passage through 4 short-lived (~100 ns) CHA states (denoted with ‘) associated with sequential rearrangements of the ECD (twist), TMD (tilt) and gate region (dilation), followed by pore water/membrane movements to allow ion transit (pop); Φ value given below each transition (Purohit et al., 2013; Gupta et al., 2017). The only agonist dependent rearrangements are catch (AC⇄ACLA; ΔGLA) and hold (ACLA⇄AC’; DGHA-ΔGLA), stages of the induced fit linked in a LFER (Figure 2). Together, sojourns in the ensemble of CHA states appear in single-channel currents as a brief gap (F for ‘flip’) (Auerbach, 1993; Lape et al., 2008; Mukhtasimova et al., 2005; Shi et al., 2023). The longitudinally-decreasing, coarse gradient in Φ suggests that the channel-opening gating transition is a conformational cascade (‘wave’) that propagates from the agonist to the gate, block-wise (Grosman et al., 2000), perhaps as an extended LFER. To emphasize that in AChRs binding ispart of gating, we show that a CRC and synaptic decay time constant can be calculated from the agonist association rate constant, kon to C (catch) if η and L0 are known a priori. (A) CRC. For many agonists, koff,C ~15,000 s–1 (Jadey and Auerbach, 2012). Calculate (i) KdC~1.5 × 104 s–1/kon,C, (ii) KdO (Equation 2), (iii) L2 (Equation 1) and (iv) POmax and EC50 (Equation 4a, Equation 4b, Equation 4c, Equation 4d). For example, η=0.5 and L0=7.4 × 10–7 (at Vm=-100 mV). ACh: measure kon,C=108 M–1s–1, calculate KdC = 150 μM, KdO = 22 nM, L2=33, POmax = 0.97 and EC50=31 μM. Choline: measure konC = 5 × 106 M–1s–1, calculate KdC = 3 mM, KdO = 9 μM, L2=0.08, POmax = 0.08 and EC50=6.8 mM. The procedure can be reversed (konC can be estimated from a CRC). (B) Synaptic decay time constant (τ). In adult-type AChRs the diliganded channel-closing rate constant (b2) for many agonists is ~2500 s–1 (–100 mV and 23 °C) (Grosman et al., 2000). τ~0.4 (1+f2/2*koff), where f2 is the diliganded opening rate constant, f2=L2*2500 s–1. Using the above kon,C for ACh yields τ=1.5ms. That the agonist’s association rate constant can approximate POmax, EC50 and τ demonstrates the entanglement between binding and gating. Unlike diffusion, natural selection can adjust the catch ‘induced fit’ (kon,C) to fine tune physiological responses.
1Experimental evidence for catch in AChRs: kon to C is (i) slower than the limit set by diffusion, (ii) correlated with agonist potency rather than diffusion constant, (iii) slower than kon to O that is approximately diffusional (Nayak and Auerbach, 2017)3 and (iv) for choline highly temperature dependent (Gupta et al., 2017). Evidence that the hold stage of the induced fit occurs at the start of the global isomerization is that (i) Φ~0.95 for agonists and binding site residues (Purohit et al., 2013) (see below), and (ii) agonists increase the channel-opening rate constant (the affinity increase occurs before the transition state; Figure 1A).
2 Φ is log f2/log L2 for a series of perturbations and reports the free energy change of the perturbed location at the gating transition state (relative to A2O). In AChRs, a longitudinal, blocky, decreasing gradient in Φ (neurotransmitter site to gate) suggests the allosteric transition is a cascade of discrete domain rearrangements that connects A2C and A2O (Auerbach, 2005). Although Φ values sequence (1–0, early to late) and locate gating rearrangements, they do not provide temporal information or quantify energy coupling between domains.
3 The barrier that prevents agonists from forming ACLA by diffusioserves a purpose. In AChRs, kon,C correlates with agonist potency (Jadey and Auerbach, 2012; Jadey et al., 2011; Nayak and Auerbach, 2017), so the weak-agonist choline (present at the synapse at a high concentration) that would otherwise interfere with signaling is excluded from the pocket, preventing competitive antagonism. Extracellular cations compete with agonists to slow kon,C (Cs+>K+>Na+>Li+) (Akk and Auerbach, 1996), and the mutation εE184Q eliminates this competition (Akk et al., 1999). We hypothesize that the ions and agonist compete at the encounter complex site (for instance, K++C⇄KC) rather that the aromatic pocket. Agonist occupancy of the ultra-low-affinity AC binding site triggers the catch-and hold rearrangement.

In bind, the agonist diffuses to the target and forms an encounter complex (Held et al., 2011; Homans, 2007; Schiebel et al., 2018), after which a local rearrangement (an induced fit) called ‘catch’ establishes the LA complex (A+C⇄AC⇄ACLA). Gate starts with the second stage of the induced fit called ‘hold’ that increases agonist affinity and, after several additional conformational changes in other protein domains, terminates with rearrangements in the pore that allow water and ions to pass (ACLA⇄ACHA⇄…⇄AOHA).

According to Equation 2, efficiency depends only on the LA/HA binding energy ratio. Regarding ΔGLA (proportional to logKdC), the agonists we examined (Figure 5) all have approximately the same diffusion constant so we surmise that differences are caused by differences in catch. Regarding ΔGHA (proportional to logKdO), the effect of perturbations away from the neurotransmitter site are agonist independent so we surmise that differences are caused by differences in hold. Therefore, with regard to η and the CRC (red pathway, Figure 1B), we attribute differences between agonists exclusively to differences in the free energy changes associated with induced fit that is the pair of catch-hold rearrangements, AC⇄ACLA⇄ACHA (Figure 2B).

The energy change in catch is ΔGLA and in hold is ΔGHA-ΔGLA. Their sum, ΔGHA, is the total free energy change delivered by the ligand. From Equation 2, η is the fraction of this total associated with hold, and 1-η is the fraction associated with catch. Multiplying η by 100% gives the percent of agonist binding energy used for the hold rearrangement of the binding site that jumpstarts the full conformational cascade that connects the neurotransmitter sites with the gate.

In a pure binding reaction, for example A+C⇄AC, Kd is the concentration where the energy gained from formation of the complex is equal to the entropy lost from removing a ligand from solution that is indexed to a reference concentration (Phillips, 2020). However, if KdC and KdO are dominated by free energy changes associated with the induced fit, the entropy components become negligible (Figure 7—figure supplement 1). As far as comparative agonist action in AChRs is concerned, it appears that only the energy changes in catch and hold are germane.

EC50 and Pomax were measured from each CRC (intra-cluster interval duration histograms in Figure 3 and Figure 3—figure supplement 1). KdC and KdO were calculated (Equation 4a, Equation 4b, Equation 4c, Equation 4d) after correction for background mutations that only changed L0. c, coupling constant (KdC/KdO); η, efficiency (Equation 2); n, number of CRCs. Membrane potential,+70 mV (to minimize channel block by the agonist). Agonist structures are in Figure 5; superscripts indicate mutation backgrounds: aεS450W, bεL269F, cεE181W that increase Lo (to increase responses to weak agonists).

Figure 3. Measuring η.

(A) Top, single-channel current traces, low time-resolution (Vm = +70 mV; O is up). Clusters of openings are bind-and-gate (Figure 1B); silent periods between clusters are desensitized. Bottom, example clusters and intra-cluster interval duration histograms. PO was calculated at each [agonist] from shut- and open-interval time constants. (B) CRCs. PO values were fitted to estimate POmax and EC50 from which KdC and KdO were calculated (Equation 4a, Equation 4b, Equation 4c, Equation 4d, Materials and methods) (Table 1). The logs of these constants are proportional to ΔGLA and ΔGHA, the ratio of which gives η (Equation 2). The profile for SCh is relatively left-shifted because this agonist is ~10% more efficient than Dec. Symbols are mean ± sem (Table 1; see Figure 3—figure supplement 1 for other agonists). The background mutation εS450W compensates for the effects of depolarization on the gating rate constants (see Materials and methods).

Figure 3.

Figure 3—figure supplement 1. CRCs.

Figure 3—figure supplement 1.

Left, example histograms and clusters (see also text Figures 3 and 5). Vm = +70 mV; O is up. Efficiencies and background mutations are in Table 1. Symbols are mean + sem.

Agonists

We measured η for 5 agonists using adult-type AChRs with wt neurotransmitter binding sites (Table 1). For each, single-channel currents were recorded at different agonist concentrations and PO values calculated from shut and open interval durations were compiled into a CRC. L0 was known a priori (Nayak et al., 2012) so KdC and KdO could be calculated from EC50 and POmax by using Equation 4a, Equation 4b, Equation 4c, Equation 4d (Materials and methods). As shown elsewhere, CRCs compiled from whole-cell currents serve equally well for η estimation (Indurthi and Auerbach, 2021).

Figure 3 shows CRCs for succinylcholine (SCh) and decamethonium (Dec). Although POmax is similar for both agonists, EC50 (potency) of SCh is substantially lower than that of Dec. For each agonist, ΔGLA and ΔGHA were calculated to yield an η value (Equation 2) with the result were ηSCh=0.45 and ηDec=0.41 (Table 1). SCh is 10% more efficient than Dec, which is the root cause of the abovementioned mismatch between POmax and potency. Figure 3 shows that given L0, agonist efficiency can be calculated from a single CRC. Because η is the ratio of two logarithms, error arising from errors in POmax and EC50 are small (Indurthi and Auerbach, 2021). CRCs for the other agonists are in Figure 3—figure supplement 1.

η values were calculated from previous measurements of KdC for 18 other agonists (Figure 4A). An x-means cluster analysis of all 23 agonist η values indicates that there are 5 groups (mean ± sd): 0.32±0.035, 0.41±0.005, 0.45±0.014, 0.51±.008 and 0.55±0.015.

Figure 4. Agonist h values.

Figure 4.

(A) Each symbol represents the average η value of one agonist, calculated from CRCs (Table 1, Table 1—source data 1). The x-means algorithm was used to separate the agonists into 5 η classes (mean, vertical bars are ± sd). (B) Efficiency plots. Each color group was fitted by a straight line (Equation 3) with L0=5.2 × 10–7 (sd of each point smaller than the symbol). η values calculated from the slopes are the same as in panel A. x- and y-axes are proportional to the agonist’s free energy changes in catch (ΔGLA) and hold (ΔGHA-ΔGLA) induced fits (Figure 2B).

Figure 5 shows the agonists grouped by η class. The neurotransmitter ACh, its breakdown product choline (Cho), and the partial agonists carbamylcholine (CCh) and nicotine all belong to the η~0.51 class. The second-most common class, η~0.41, includes ligands that have a bridge nitrogen (for instance, Ebt) as well as others that do not (for instance, Dec and TMP). Small ligands (for instance, TriMA, MW 60) appear to be the most efficient and large rigid ligands (for instance, varenicline, MW 211) appear to be the least efficient.

Figure 5. Agonists grouped by η class (Figure 4).

Figure 5.

See Materials and methods for abbreviations.

It is valuable to combine Equations 1 and 2 (Nayak et al., 2019),

logL2=logL0+mlog(1KdC)η=m(m+2). (3)

Equation 3 describes an ‘efficiency’ plot, log affinity (1/KdC) versus log relative efficacy (L2). An average η for a group of agonists is estimated from the slope of the straight line fit (m). Equation 4a, Equation 4b, Equation 4c, Equation 4d converts readily measured CRC parameters (POmax and EC50) into equilibrium constants (KdC and L2), and Equation 3 converts these into fundamental constants that pertain to the agonist (η) and the receptor (L0). The value of the efficiency plot is that it increases the accuracy of the η estimates because if L0 is known a priori, the y-intercept can be added as a fixed point to all lines. For receptors in which L0 has not been measured, the efficiency plot offers a convenient way to do so, as was done previously for glutamate, GABA, glycine, and muscarinic recepotors (Nayak et al., 2019).

η values estimated from the slopes for the 23 agonists (Figure 4B) are the same as from the x-mean cluster analysis (Figure 4A) (mean ± sd): 0.31±0.018, 0.41±0.002, 0.46±0.004, 0.51±0.002 and 0.56±0.008. Affinity and efficacy are correlated significantly within the 3 classes having >2 members (19 agonists; Pearson’s correlation test p-values for η=0.41 and 0.51,<0.0001, and for η=0.56, 0.019). Also, the slopes for these 3 classes are significantly different (ANCOVA, F=129, p<0.001). The association of the other agonists with a discrete η class is less certain but supported by mutation studies (see below).

Figure 4B shows that in AChRs, agonists having the same affinity can have different efficacies, and vice versa. Absent classification of an agonist according to η (imagine all symbols the same color) there is no global correlation between these two agonist properties. However, a linear correlation between log affinity and log efficacy is clear within each class. In AChRs, the presence of multiple η classes precludes a global correlation between efficacy and affinity.

Mutations

η values were measured previously in adult-type AChRs having one of 42 binding site mutations (Table 2—source data 1). To these we add 11 more (Table 2), for 5 agonists and 3 binding-site mutations (αD200A, αK145A, αG153S; Figure 6, Figure 6—figure supplement 1, Figure 6—figure supplement 2).

Figure 6. CRCs for αK145A and αD200A.

Agonist efficiencies calculated from the fitted CRC parameters are in Table 2. Intra-cluster interval histograms are in Figure 6—figure supplement 1 and Figure 6—figure supplement 2. Inset, α−δ subunit interface of an AChR neurotransmitter binding site occupied by CCh (cyan) (7QL6.pdb; Zarkadas et al., 2022).

Figure 6.

Figure 6—figure supplement 1. αK145A.

Figure 6—figure supplement 1.

Example histograms and clusters (agonist structures in Figure 5). Vm = +70, O is up. CRCs are in Figure 6; efficiencies and backgrounds are in Table 2.
Figure 6—figure supplement 2. αD200A and αG153S.

Figure 6—figure supplement 2.

Example histograms and clusters (agonist structures in Figure 4). Vm = +70, O is up. CRCs in Figure 6. (Efficiencies and backgrounds in Table 2).

αD200 and αK145, along with αY190, have been suggested to work together to initiate the channel-opening conformational change (Mukhtasimova et al., 2005). The mutation αY190A reduces ηACh from 0.50 to 0.35, but αY190F is without effect (Bruhova and Auerbach, 2017). However, αY190F does cause substantial losses in LA and HA binding energies for ACh, to an extent that depends on the αK145 side chain (Bruhova and Auerbach, 2017). These results support the suggestion that these side chains work together, but exactly how and to what effect remains unclear. The agonists we tested with αD200A or αK145A were from 4 different η classes (wt class value): TMA (0.54), CCh (0.51), ACh (0.50), Ebx (0.46), and Ebt (0.41).

For mutation location see Figure 6, inset. KdC and KdO were calculated from CRC parameters (Figure 6) by using Equation 4c, coupling constant (KdC/KdO); n, number of CRCs. L0 was corrected for background mutations (Materials and methods): aεS450W, bεL269F, cεE181W. See Figure 6—figure supplement 1 and Figure 6—figure supplement 2 for intra-cluster interval duration histograms.

The results are in Table 2. The substitution αD200A reduces the efficiency of 4 of the 5 agonists to 0.33±0.04 (mean ± sd), or to about the same level as with αY190A with ACh. The exceptional ligand was TMA for which η remained relatively high at 0.48. In contrast, in αK145A η for all 5 agonists was 0.41±0.03. Interestingly, this mutation reduces η for ACh, CCh and TMA but not significantly for Ebx and Ebt. Note that the mutation αW149A increases η for ACh to 0.60 (Purohit et al., 2014).

Table 2—source data 1 shows KdC and KdO values measured previously for 4 different agonists in AChRs having a substitution at αG153 (Jadey et al., 2013), including a Ser that causes a congenital myasthenic syndrome (Engel et al., 1982). After converting the published equilibrium constants to ΔGLA and ΔGHA, the results indicate that on average the mutations reduce η for Cho, DMP, TMA and nicotine to 0.41±0.03 (mean ± sd), or by ~25%. To these we add the new result that αG153S reduces η of Ebt from 0.42 to 0.31, or also by ~25%.

Figure 7A shows x-means cluster analysis of efficiencies for 53 AChR binding site mutations. Mutant η values (mean ± sd) segregate into the same 5 η classes that were apparent with agonists: 0.33±0.03, 0.40±0.02, 0.44±0.01, 0.49±0.01 and 0.56±0.03. The η classes that were under-represented and poorly defined with agonists (Figure 4B) are more common and clearer with mutations.

Figure 7. Mutation η values.

Each symbol is η value calculated for an individual mutation (various agonists; Table 2 and Table 2—source data 1). (A) There are 5 efficiency classes (mean ± sd). As with agonists (Figure 4), the ~0.5 and~0.4 classes predominate. (B) Weak (LA) versus strong (HA) binding energies for the mutants (sd of each point is smaller than the symbol). The 5 slopes reflect different correlations between catch and hold free energy changes (Figure 2B). The efficiency distribution for the mutations (η=1-slope; see text for mean ± sd) is the same as for agonists. C. Spectrum of efficiencies, adult AChR neurotransmitter binding sites. The width of each line is proportional to prevalence (agonists and mutations).

Figure 7.

Figure 7—figure supplement 1. Estimating L0 (αD200).

Figure 7—figure supplement 1.

All currents were recorded in the complete absence of agonists. Top traces, clusters of unliganded single-channel openings are mainly C⇄O (Vm=-100 mV; O is down); silent periods between clusters are desensitized. Bottom, interval duration histograms and an example cluster. Left, the background mutations together increase L0 from 7.4x10–7 to 0.17 (by a factor of 2.46x105). Right, the background mutations plus αD200A increase L0 from 7.4x10–7 to 0.64 (by a factor of 3.76x105). Hence, the A substitution at αD200 increases L0 by 3.76-fold. L0 is voltage dependent and is 5.2x10–7 at –70 mV (Nayak et al., 2012). We calculate that at this membrane potential (Figure 6) L0 with αD200A is 3.76-fold grater, or 1.9x10–6. L0 was measured individually using the above method for all 53 mutations (Table 2 and Table 2—source data 1).
Figure 7—figure supplement 2. Evidence that in AChRs KdC reflects mainly the free energy change of the catch rearrangement (see Figure 2—figure supplement 1).

Figure 7—figure supplement 2.

In A+C⇄AC ⇄AC (Figure 2A), the first step includes a term for the entropy decrease from the loss of a free ligand, indexed to a reference concentration (Phillips, 2020). This term does not cancel in Equation 2, but in AChRs the second step dominates so η is the same with or without this term. (A) Agonist η-plots (0.51 and 0.42 class ligands; Figures 4B and 5) using different reference concentrations. The slopes of the straight-line fits and η (Equation 3) are independent of the reference concentration. (B) Mutation plots. The entropy term was removed by first normalizing KdC and KdO by those of a standard agonist (here, choline; 2 other agonists give the same result). The slopes of plots, ΔΔGLA/ΔΔGHA, are approximately same as without normalization (Figure 7B).

With mutations, an η-plot is not useful because substitutions can change L0 (see Figure 7—figure supplement 1) and, hence, the y-intercept of each line. Instead, a plot of ΔGLA versus ΔGHA shows the binding energy correlations directly. This plot for mutations (Figure 7B) shows 5 slopes (Pearson’s correlation test p-value <0.0006 for all classes). The distribution of η values (1-slope) is (mean ± sd): 0.32±0.007, 0.40±0.004, 0.44±0.003, 0.49±0.002 and 0.56±0.007, with all slopes being significantly different (ANCOVA, F-value, 63.35; p-value,<0.0001).

Combining the results for agonists and mutations, the overall distribution of η (relative prevalence) is: 0.56 (17%), 0.51 (31%), 0.45 (13%), 0.41 (26%), and 0.31 (12%). As was the case with agonists and mutations separately, the 0.51 (example, ACh) and 0.41 (example, Ebt) classes predominate. Figure 7C shows the distribution of agonist plus mutation η values as a spectrum in which line thickness represents relative prevalence.

Discussion

Efficiency

Efficiency is the missing link that connects binding to gating (Equation 2). Insofar as KdC and KdO apply generally (Figure 1), η is a universal agonist attribute that depends only on these two constants. In AChRs, KdC and KdO are set mainly by energy changes in a pair of local rearrangements of the binding site (the catch-and-hold induced fit), with η being the fraction of the total used to initiate the allosteric transition of the receptor. As such, η calibrates the fundamental connection that defines receptor action.

Converting ligand binding energy into energy for an otherwise unfavorable protein conformational change is an induced fit. We assume that without an agonist present, both catch and hold rearrangements (of an aromatic pocket) are energetically unfavorable and generate the high barrier to unliganded opening (Figure 1). In enzymes, a fraction of substrate binding energy is used to promote a local protein rearrangement that stabilizes the reaction transition state (Richard, 2022). In AChRs, neurotransmitter binding energy is divided equally between two stages of an intrinsically unfavorable rearrangement that forms the LA and HA complexes and starts the gating isomerization. In brief, η quantifies the split in ligand binding energy between the two steps in the catch-and-hold induced fit (Figure 2B).

η classes

Agonists having radically different resting affinities and efficacies can have the same η. For example, ACh (KdC = 175 μM, POmax = 0.96) and choline (KdC = 4 mM, POmax = 0.05) both divide their binding energy equally between catch and hold. In AChRs, agonists segregate into discrete ηclasses within which all members use approximately the same fraction of their post-diffusion binding energy for catch. So far we have identified 5 η classes with catch percentages ranging from 47% to 71% (Table 1). Below, we discuss the possibility that larger ligands apply a larger percentage of their binding energy to catch.

η classes are notable because its two component energy changes (Equation 3) are realized in bind and in gate, usually considered to be qualitatively different processes (Figure 1B). Nevertheless, the existence of an η class indicates that in AChRs these two binding energy (structure) changes are correlated and joined in a linear free energy relationship (LFER) (Figure 2B). Any change in bind (catch) free energy compels one in gate (hold), reciprocally. Rather than being independent processes, bind and gate are completely entangled even if they are depicted as different dimensions of a reaction scheme (Figure 1B).

Although an empirical η value can be calculated for every agonist of every receptor, the generality of η classes is less certain. The components, KdC and KdO for many agonists, have not been measured extensively in other receptors, but some results suggest that a catch-and-hold induced fit, with stages linked in a LFER, is not uncommon.

Regarding hold, it is likely that an increase in affinity caused by a local binding site rearrangement inside the gating isomerization is typical (ACLA⇄ACHA). In enzymes and receptors, agonists induce a ‘clamshell’ closure of the ligand site that is associated with increased binding strength (Armstrong and Gouaux, 2000; Masiulis et al., 2019; Pless and Lynch, 2009; Traynelis et al., 2010). Also, in many receptors agonists increase the opening rate constant as well as PO (Clements et al., 1998; Clements and Westbrook, 1991; Maconochie et al., 1994) indicating that the stabilization by the lgiadn of an otherwise unfavorable rearrangement (the affinity increase) occurs early in the isomerization (Figure 1A). Finally, in other receptors an ACHA intermediate state has been detected directly in single-channel currents (Lape et al., 2008; Shi et al., 2023) or identified in structures (Sauguet et al., 2014).

Regarding catch, this first step in the induced fit serves a purpose (Figure 2—figure supplement 1) and may also pertain to other receptors. As in AChRs, in several receptors kon to C is slower than diffusion and correlated with agonist potency (Dravid et al., 2008; Grewer, 1999; Lewis et al., 2003; Mortensen et al., 2010). Further, a catch-and-hold LFER is implied in other receptors (including a GPCR; Sykes et al., 2009) because efficiency plots are linear, with outliers that possibly indicate different η classes (Nayak et al., 2019). In biological reactions, LFERs are empirical descriptors and more experiments are needed to determine the generality of η classes. In addition to measurements of KdC and KdO in other receptors, and identifying the structures of intermediate liganded states AC and ACLA would corroborate catch.

In AChRs, multiple η classes together preclude the appearance of a global affinity-efficacy correlation. If η classes turn out to be common, then the consensus view that there is no correlation between these agonist properties needs to be reexamined (Colquhoun, 1998; Kenakin and Onaran, 2002). The clear correlation between log affinity and log efficacy within each class vanishes when agonists from different classes are lumped together (Figure 4B).

We detected 5 η classes in AChRs but additional experiments could reveal more. In particular, agonist classes with the highest and lowest η values have the greatest scatter and could be amalgams. Nonetheless, the ability to corral 76 different agonist-receptor combinations into just 5 classes represents a significant reduction in complexity. If nothing else, η is a useful way to classify agonists and, possibly, binding sites. Also, η could prove to be useful CRC analysis because it allows EC50 to be computed from POmax and vice versa by using (Equation 3, Equation 4a, Equation 4b, Equation 4c, Equation 4d; Indurthi and Auerbach, 2021).

Mutations of binding site residues segregate into the same 5 efficiency classes as do agonists. This supports the idea that an agonist at a binding site is much like an ordinary side chain, with exceptions. Most importantly, agonists are not linked covalently and so are both hypermobile in the pocket and free to come and go to serve as a signal. Also, by definition a bound agonist is more stable in O versus C (Figure 1A) but this is the opposite of most AChR wt side chains (Purohit et al., 2013). η measurements reinforce the standard view that ligands only perturb the intrinsic activity of allosteric proteins.

η and structure

In AChRs, the pair of structural changes at the binding sites associated with energy changes in hold (ΔGHA-ΔGLA) and in catch (ΔGLA) are not known with certainty. Although there are structures of apo-C (Zarkadas et al., 2022) and desensitized (perhaps the same as AOHA) (Auerbach, 2020; Rahman et al., 2020; Zarkadas et al., 2022), those of the three relevant liganded-closed intermediate states AC, ACLA and ACHA are not available. Here, we discuss inferences about the structures of these three transient states that can be made from η.

If every agonist had a unique η then little could be said about the underlying conformational changes. However, the existence of an η class and, therefore, an LFER between catch-and-hold energy changes suggests that the corresponding structural changes, too, are correlated, although perhaps not linearly. That is, catch and hold are two related stages of a single induced fit rearrangement of the binding site.

Regarding hold, at HA sites loop C covers (‘caps’) the aromatic pocket (Nys et al., 2013) that appears to be contracted. Also, simulations suggest the possibilities that in the hold step the ligand flips its orientation (Tripathy et al., 2019), and that water is extruded from the binding interface (Young et al., 2007). The components of the hold free energy change arise from these and other restructuring events, but the breakdown is not known. Regarding catch, little is known about this rearrangement directly because the structures of neither end state have been solved. (Note that apo-C and AC are different). However, because of the LFER, we hypothesize that all or some of the above putative rearrangements in hold (cap, contract, flip, de-wet) also happen in catch, but perhaps to lesser extents.

A surprising inference about the binding site structural change comes from measurements of agonist association rate constants (kon) to C versus O (Grosman and Auerbach, 2001; Nayak and Auerbach, 2017). Whereas kon to C is slower than diffusion and correlated strongly with agonist potency, kon to O is approximately at the diffusion limit and weakly agonist dependent. Hence, slow association to C cannot be attributed to the ligand itself but rather suggests that the hold stage of the induced fit removes a barrier(s) that otherwise prevents free entry of the ligand into the aromatic binding pocket (Figure 2—figure supplement 1). Neither the site of the encounter complex nor the nature of this barrier are known. Because loop C capping and pocket contraction would be expected to decrease rather than increase kon, we speculate that in AChRs other, still-unidentified structural changes in hold serve to remove the barrier. Removing loop C eliminates activation by agonists but not unliganded gating (Purohit and Auerbach, 2009). Apparently, clamshell closure is necessary for transducing ligand binding energy into the receptor isomerization, but not for the global isomerization itself. However, it is uncertain if other rearrangements in the binding site induced fit occur in unliganded gating after loop C removal (see below).

Energy change reflects a change in structure, so multiple η classes (logKdC/logKdO ratios) imply there are multiple pairs of C versus O (post-catch and post-hold) binding site conformations. The existence of 5 discrete ΔGLA/ΔGHA ratios indicates that there are at least this many ACLA/ACHA binding site structural pairs. Possible combinations that give rise to 5 η classes are (number of ACLA/number of AOHA) 1/5, 2/3, 3/2, 4/2, and 5/1. For example, with the 1/5 combination a single ACLA binding site structure selects its target AOHA structure from among 5 alternatives.

The discrete distribution of η classes suggests that the structural changes in the two stages of the induced fit also are discrete rather than continuous. The AChR allosteric transition combines elements of induced fit (use of binding energy to drive a local rearrangement) and conformational selection (constitutive activity and the adoption of pre-existing target structures) (Changeux and Edelstein, 2005).

The binding energy increase (in hold) occurs at the onset of the channel-opening gating isomerization (Grosman et al., 2000). The existence of multiple, distinct C versus O structural pairs at each binding site indicates that the end states in the simple activation scheme (Figure 1B) do not represent single stable structures but rather ensembles made up of receptors having at least five distinct binding site structural pairs. This raises the possibility that there could be five isomerization pathways that might culminate in five different structural perturbations of the gate region. Although there is no evidence in AChRs of any agonist-dependent variations in output properties (conductance or ion selectivity), in other receptors functional output is agonist dependent (Kenakin and Christopoulos, 2013). Efficiency measurements of these receptors would test the possibility that η classes are associated with the phenomenon of biased agonism (Ehlert, 2018).

Finally, regarding the agonist (Figure 5), there is a trend for those with cationic centers that occupy smaller volumes in vacuo to have greater η (Indurthi and Auerbach, 2021). That is, agonists with larger volumes tend to use a greater fraction of their binding energy for catch. In AChRs and other receptors, the volume of the binding pocket appears to be smaller in O versus C (Tripathy et al., 2019), so it is also possible that unfavorable VdW interactions caused by pocket contraction guide the selection of (for example) the target O structure, to decrease η. The relationship between catch and hold energy change and the corresponding structural elements (agonist, protein, water) is complex and further investigations are needed.

Extending η

The catch-and-hold LFER implied by an η class indicates that structure (energy) changes in these two binding site rearrangements are related. Microscopic reversibility demands that if catch promotes hold, then hold promotes catch. Below, we present evidence suggesting that this bidirectional linkage is the tip of an iceberg, and that the entire AChR activation cascade - from an agonist at the encounter complex to water at the gate (AC to AO) - is linked as an extended, multi-stage LFER (Figure 2—figure supplement 1).

There have been several proposals in AChRs regarding structural changes that link the neurotransmitter sites and the gate, but these are difficult to assess because they are without corroborating energy measurements. Likewise, η measurements reveal a binding-gating energy link at the agonist site, but lack corroborating structural information. However, separate activation domains in AChR activation have been identified. For instance, η calibrates the catch-hold connection at the agonist binding site, mutations decouple the ECD-TMD interface (Cymes and Grosman, 2021; Shi et al., 2023) and interactions between gate residues and pore water have been investigated (Beckstein and Sansom, 2006; Rasaiah et al., 2008; Yazdani et al., 2020). Analyses of the gating transition state suggest that there are 5 discrete domain rearrangements (Figure 2—figure supplement 1). In opening, hold is followed by movements of the ECD, TMD, gate region, and, finally, pore water and membrane lipids (Gupta et al., 2017).

Evidence for the extended LFER hypothesis is the experimental measurements of kon to C versus O. As mentioned above, when an unliganded AChR opens constitutively, kon increases to approach the diffusion limit, and also loses its agonist dependence. This indicates that with loop C intact, adopting the global O shape in the absence of agonists is sufficient to induce the local change in structure that removes the entry barrier, in hold. That a distant mutation (for example, of a gate residue) increases constitutive opening and also influences the ligand association rate constant is evidence that binding and gating are entangled.

This result, along with η and Φ measurements, leads us to propose that everything in the protein’s activation cascade, AC to AO, is energy-linked in a reversible, extended LFER. In this sequence, each individual domain energy change (rearrangement) influences those of its neighbors, bidirectionally. Usually, the first link in activation (after the encounter complex) is catch-and-hold, and the last is gate-and-water. However, in a LFER the consequence of a perturbation propagates both forward and backward, so a gate mutation would also perturb the agonist site and one at the ECD-TMD would induce rearrangements both there and at the gate. An extended LFER offers a way to transfer energy (change structure) over distance by using only local domain rearrangements. In AChRs, the allosteric transition appears to be a energy-coupled chain of domain rearrangements that crosses seamlessly the traditional divide between binding and gating. η is at the core receptor operation insofar as it links the induced fit with the conformational gating cascade, Φ reveals the cascade’s components (and the sequence of domain rearrangements), and kon to O reveals that a reverse cascade, gate→binding site, exists without agonists. Transit across the whole extended LFER is rapid (~5 μs; Lape et al., 2008), with each interemdiate states lasting on the order of 100 ns (Gupta et al., 2017).

There are many unanswered questions. We do not know the structural changes in the induced fit, or their energy consequences. Regarding agonists, we do not know the full spectrum of η classes, or their structural correlates. Regarding mutations, we do not know the reasons specific agonist/side chain combinations change η or whether mutations of additional residues (including at a distance from the agonist site) can also do so. Regarding binding sites, we know little about the location of the encounter complex, why agonists are more efficient at the fetal AChR neurotransmitter site (Nayak et al., 2019), or η values in other subtypes of nicotinic receptor or atl allosteric modulator binding sites. The extended LFER hypothesis needs to be tested, with both structure and energy changes identified and quantified. Regarding other receptors, there are few reports of the key experimental values (KdO, kon to O, Φ) so it is unknown whether η classes, a staged induced fit, a conformational cascade, or an extended LFER apply generally. We anticipate that additional structure and energy measurements, in combination with computation, will reveal the molecular forces that underpin activation and lead to a deeper understanding of how ligands promote conformational change in allosteric proteins.

Materials and methods

Expression

Human embryonic kidney (HEK) 293 cells were maintained in Dulbecco’s minimal essential medium (DMEM) supplemented with 10% FBS and 1% penicillin–streptomycin (pH 7.4). HEK cells obtained from ATCC (CRL-1573, lot no. 57925149) are authenticated using STR profiling and tested free of mycoplasma contamination. Mutations were incorporated into AChR subunits using the Quickchange II site directed mutagenesis kit (Agilent Technologies, CA) according to manufacturer’s instructions. Sequence was verified by nucleotide sequencing (IDT DNA, I). AChRs were transiently expressed in HEK 293 cells by transfecting (CaPO4 precipitation) (Purohit et al., 2014) mouse α1 (GFP encoded between M3-M4),β1,δ,ε subunits (3–5 μg total/ 35 mm culture dish) in a ratio of 2:1:1:1 for ~16 hrs. Most electrophysiological experiments were done 24–48 hr post-transfection.

Electrophysiology

Single-channel currents were recorded in cell-attached patches (23 °C). The bath solution was (in mM) 142 KCl, 5.4 NaCl, 1.8 CaCl2, 1.7 MgCl2, 10 HEPES/KOH (pH 7.4). High extracellular [K+] ensured that the membrane potential Vm was ~0 mV. Patch pipettes were fabricated from borosilicate glass, coated with sylgard (Dow Corning, Midland, MI) to a resistance of ~10 MΩ when filled with pipette solution (Dulbecco’s phosphate-buffered saline PBS) (in mM): 137 NaCl, 0.9 CaCl2, 2.7 KCl, 1.5 KH2PO4, 0.5 MgCl2, and 8.1 Na2HPO4 (pH 7.3/NaOH). Single channel currents were recorded using a PC505 amplifier (Warner instruments, Hamden, CT), low-pass filtered at 20 kHz and digitized at a sampling frequency of 50 kHz using a data acquisition board (SCB-68, National instruments, Austin, TX). For liganded activation experiments, agonists were added to the pipette solution at the desired concentrations. For unliganded activation experiments, we used pipettes and wires that were never exposed to agonists. To reduce the effect of channel block without affecting binding of agonist to the receptor, membrane potential (Vm) was held at +70 mV when agonists were used (Jadey et al., 2011).

Current analysis

Analyses of the single-channel currents were performed by using QUB software (Nicolai and Sachs, 2013). Single-channel currents occur in clusters when the opening rate constant is significantly large. For analysis, we selected clusters of shut/open intervals that appeared (by eye) to be homogeneous, with regard to Po. We limited the analysis to intracluster interval durations and thus excluded sojourns arising from desensitized states (shut intervals between clusters >20ms). The clusters were idealized into noise-free intervals after digitally filtering the data at 10–15 kHz (Qin, 2004). First, the idealized, intra-cluster intervals were fitted by a two-state model, C⇄O. Then, additional nonconducting and conducting states were added, one at a time, connected only to the first O state, until the log likelihood failed to improve by 10 units (Qin et al., 1997). Cluster PO at each agonist concentration was calculated from the time constants of the predominant components of the shut- (τs) and open-time distributions (τo): τo/(τSo). In this way, an equilibrium CRC was constructed as a plot of the absolute PO (not normalized) versus the agonist concentration (see Figure 2).

Equilibrium constants

Two equilibrium dissociation constants comprise efficiency, KdC and KdO (Figure 1; Equation 3). These, and the fully-liganded gating constant L2, were estimated in two ways, with both methods producing the same results.

In the primary approach, KdC and KdO were estimated from CRC parameters. The CRC was fitted by the Hill equation to estimate POmax (the high concentration asymptote) and EC50 (the agonist concentration that produces a half-maximum PO). Equilibrium constants were calculated from the reaction scheme pertaining to the main activation pathway that assumes L0 and KdO are negligible. Let x=[A]/KdC. For a one-site receptor the scheme is A+C⇄AC⇄AO and PO([A])= xL1/(1+x + xL1). For adult AChRs that have two equal and independent binding sites the scheme is A+C⇄AC⇄A2C⇄A2O (red, Figure 1B) and PO([A])= x2L2/(1+2x+x2+x2L2). Relating the two site scheme to CRC parameters, POmax is the infinite-concentration asymptote, POmin is the zero-concentration asymptote and EC50 is the concentration at which PO is half POmax,

L2=-POmaxPOmax-1 (4a)
L0=-POminPOmin-1 (4b)
KdC=EC50L2+11+L2+2 (4c)
KdO=KdCL2L0 (4d)

where KdO is solved by using Equation 1. Because L2=L0c (Equation 1), any change in L0 (see below) will change all three CRC parameters (POmax, POmin and EC50) even if the equilibrium dissociation constant ratio KdC/KdO remains unchanged. Knowledge of, and ability to manipulate, L0 was the key to measuring η.

Voltage, L0 and background mutations

To reduce channel block by the agonists, the membrane was depolarized to +70 mV. To compensate for changes in τS and τo caused by depolarization, we added the background mutation εS450W. This residue is far from the binding site (M4 transmembrane region of the ε subunit), has no effect on KdC, and has equal but opposite effect on unliganded gating as does this extent of membrane depolarization (Jadey et al., 2011). L0 is 7.4x10–7 at Vm=-100 mV and reduced e-fold by a 60 mV depolarization (Nayak et al., 2012). Hence, we calculate that L0 is 5.2x10–7 at Vm=-70 mV as well as in our experiments at Vm = +70 mV plus εS450W.

In some conditions, for instance low efficacy agonists (Dec, TriMA, and BzTEA) and aD200A, the wt opening rate constant was small and single-channel clusters were poorly defined. Accordingly, we added background mutations to facilitate PO measurements (Tables 1 and 2). These were εL269F (located in the M2 helix) and εE181W (located in strand β9) that increase the L0 by 179- and 5.5-fold (1084-fold for the pair) without effecting KdC. First, we obtained the apparent L2 from the CRC PO max (Equation 4a). Second, we divided this value by the fold increase in L0 caused. By the background to obtain a corrected L2. Finally, agonist KdC was estimated from EC50 and the corrected L2 (Equation 4c).

L0 for αD200A

In wt adult AChRs, L0 is 5.2x10–7 at Vm = +70 mV (Nayak et al., 2012) and ofcourse is the same for all agonists. L0 has been reported previously for the mutations αK145A (Bruhova and Auerbach, 2017) and αG153S (Jadey et al., 2013). To estimate this L0 for αD200A, the pipette solution was free of any agonist and currents were measured at a membrane potential of −100 mV. The AChRs had added background mutations far from αD200 and each other (εL269F+εE181W+δV269A) that together increased unliganded activity substantially to allow cluster formation. Individually, these mutations increase L0 by 179- (Jha et al., 2009), 5.5- (Purohit et al., 2013) and 250-fold (Cymes et al., 2002), respectively (Figure 7—figure supplement 1). Assuming no interaction (Gupta et al., 2017), the expected net increase in L0 for this background combination is the product, ~2.5 × 105. L0 was measured experimentally using this background plus αD200A, from the durations of intra-cluster intervals (see above). The unliganded opening (f0) and closing (b0) rate constants were estimated from the idealized interval durations by using a maximum-interval likelihood algorithm after imposing a dead time of 25 μs and L0 was calculated from the ratio. Using a similar approach, L0 was estimated previously for each mutation shown in Figure 7.

Statistics

For single-channel CRCs, the midpoint and maximum (EC50 and POmax) were estimated by fitting to monophasic Hill equation (PO = POmax/(1+(EC50/[A])nH)) using GraphPad Prism 6 (GraphPad). nH values contain information (Qin, 2010) but were not used because the number of binding sites was known a priori. A x-means cluster analysis algorithm (QUB online: qub.mandelics.com/online/xmeans.html) was used to define agonist (Figure 4A) and mutant (Figure 7A) classes, considering cluster must have at least two elements. Optimal clustering was determined based on the Sum Square Residual (SSR) and the corrected Akaike Information Criterion (AICc): for agonists SSR = 4.76 × 10–3 and AICc = −96.5 (n=5 classes); for mutations SSR = 2.41 × 10–2 and AICc=-396 (n=5 classes). Pearson’s correlation test was performed to determine correlation significance between the two variables logKdC versus logKdO and logL2 versus log[1/KdC]. Since the goal was to measure the correlation rather than the magnitude difference between classes, Pearson’s correlation was used instead of Cohen’s. The P-value (two-tail) and r2 value for that 3 agonist efficiency classes with >2 elements (Figure 5A) were 0.019 and 0.96 (η=0.56), <0.0001 and 0.99 (η=0.51) and <0.0001 and 0.99 (η=0.41). Significance for classes with <3 data points (η=0.31 and 0.46) could not be determined. The p-value (two-tail) and r2 value for mutants (Figure 7A) were 0.0006 and 0.83(η=0.56), <0.0001 and 0.98 (η=0.51), <0.0001 and 0.97 (η=0.45), <0.0001 and 0.93 (η=0.41), and <0.0001 and 0.98 (η=0.31). η is the ratio of logarithms and as such is precise.

The position of the intermediate state ACLA in the catch-and-hold reaction sequence (1-η; Figure 2) is analogous to the position of the transition state (Φ orβ) in a single-step reaction. With Φthe intermediate state in the LFER is an energy barrier, whereas with η it is an energy well. Φ gives the fraction of the total energy change at ‡, and 1-η gives the fraction of the total energy change at ACLA.

Agonists

Abbreviations: acetylcholine (ACh), trimethyl ammonium (TRiMA), tetramethyl ammonium, dimethylpyrrolidium (DMP), dimethylthiazolidinium (DMT), nornicotine, nicotinic (Nic), carbamylcholine (CCh), anabasine (Singh et al.), dimethylphenylpiperazinium (DMPP), benzyltrimethyl ammonium (BzTMA), choline (Cho), 3-hydroxypropyltrimethylammonium (3-OH), 4-hydroxybutyltrimethylammonium (4-OH), dimethylthiazolidinium (DMT), dimethylpyrrolidium (DMP), succinylcholine (SCh), decamethonium (Dec), epiboxidine (Ebx), epibatidine (Ebt), cytisine (Cyt), tetraethyl ammonium (TEA), tetramethyl phosphonium (TMP), varenicline (Var) and benzyltriethyl ammonium (BzTEA), Choline (Cho). Agonists were from Sigma (St. Louis, MO) except DMP, DMT, 3-OH and 4-OH that were synthesized as described previously (Bruhova et al., 2013).

Acknowledgements

We thank Jan Jordan and Mary Teeling for technical assistance, and John Richard and Rob Phillips for helpful discussions.

Funding Statement

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Contributor Information

Anthony Auerbach, Email: auerbach.anthony@gmail.com.

László Csanády, Semmelweis University, Hungary.

Kenton J Swartz, National Institute of Neurological Disorders and Stroke, National Institutes of Health, United States.

Funding Information

This paper was supported by the following grant:

  • NIH Blueprint for Neuroscience Research NS-064969 to Anthony Auerbach.

Additional information

Competing interests

No competing interests declared.

Author contributions

Data curation, Software, Formal analysis, Investigation, Methodology.

Conceptualization, Supervision, Funding acquisition, Writing - review and editing.

Additional files

MDAR checklist

Data availability

The source data is available in the Tables and table source data files and referred to in the legend of the associated figure.

References

  1. Akk G, Auerbach A. Inorganic, Monovalent Cations compete with agonists for the transmitter binding site of nicotinic acetylcholine receptors. Biophysical Journal. 1996;70:2652–2658. doi: 10.1016/S0006-3495(96)79834-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Akk G, Zhou M, Auerbach A. A mutational analysis of the acetylcholine receptor channel transmitter binding site. Biophysical Journal. 1999;76:207–218. doi: 10.1016/S0006-3495(99)77190-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Armstrong N, Gouaux E. Mechanisms for activation and antagonism of an AMPA-sensitive glutamate receptor: crystal structures of the Glur2 ligand binding core. Neuron. 2000;28:165–181. doi: 10.1016/s0896-6273(00)00094-5. [DOI] [PubMed] [Google Scholar]
  4. Auerbach A. A statistical analysis of acetylcholine receptor activation in Xenopus myocytes: stepwise versus concerted models of gating. The Journal of Physiology. 1993;461:339–378. doi: 10.1113/jphysiol.1993.sp019517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Auerbach A. Dose-response analysis when there is a correlation between affinity and efficacy. Molecular Pharmacology. 2016;89:297–302. doi: 10.1124/mol.115.102509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Auerbach A. Pathways for nicotinic receptor desensitization. The Journal of General Physiology. 2020;152:e202012639. doi: 10.1085/jgp.202012639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Beckstein O, Sansom MSP. A hydrophobic gate in an ion channel: the closed state of the nicotinic acetylcholine receptor. Physical Biology. 2006;3:147–159. doi: 10.1088/1478-3975/3/2/007. [DOI] [PubMed] [Google Scholar]
  8. Bruhova I, Gregg T, Auerbach A. Energy for wild-type acetylcholine receptor channel gating from different choline derivatives. Biophysical Journal. 2013;104:565–574. doi: 10.1016/j.bpj.2012.11.3833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bruhova I, Auerbach A. Molecular recognition at cholinergic synapses: acetylcholine versus choline. The Journal of Physiology. 2017;595:1253–1261. doi: 10.1113/JP273291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Cecchini M, Changeux JP. Nicotinic receptors: from protein Allostery to computational Neuropharmacology. Molecular Aspects of Medicine. 2022;84:101044. doi: 10.1016/j.mam.2021.101044. [DOI] [PubMed] [Google Scholar]
  11. Changeux JP, Edelstein SJ. Allosteric mechanisms of signal Transduction. Science. 2005;308:1424–1428. doi: 10.1126/science.1108595. [DOI] [PubMed] [Google Scholar]
  12. Clements JD, Westbrook GL. Activation Kinetics reveal the number of glutamate and glycine binding sites on the N-methyl-D-aspartate receptor. Neuron. 1991;7:605–613. doi: 10.1016/0896-6273(91)90373-8. [DOI] [PubMed] [Google Scholar]
  13. Clements JD, Feltz A, Sahara Y, Westbrook GL. Activation Kinetics of AMPA receptor channels reveal the number of functional agonist binding sites. The Journal of Neuroscience. 1998;18:119–127. doi: 10.1523/JNEUROSCI.18-01-00119.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Colquhoun D. Binding, gating, affinity and efficacy: the interpretation of structure-activity relationships for agonists and of the effects of Mutating receptors. British Journal of Pharmacology. 1998;125:924–947. doi: 10.1038/sj.bjp.0702164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Cymes GD, Grosman C, Auerbach A. Structure of the transition state of gating in the acetylcholine receptor channel pore: a Phi-value analysis. Biochemistry. 2002;41:5548–5555. doi: 10.1021/bi011864f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Cymes GD, Grosman C. Signal Transduction through Cys-loop receptors is mediated by the nonspecific bumping of closely Apposed domains. PNAS. 2021;118:e2021016118. doi: 10.1073/pnas.2021016118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Delcastillo, J, Katz, B, Delcastillo, J. Interaction at end-plate receptors between different choline derivatives. Proceedings of the Royal Society of London. Series B - Biological Sciences. 1957;146:369–381. doi: 10.1098/rspb.1957.0018. [DOI] [PubMed] [Google Scholar]
  18. Dravid SM, Prakash A, Traynelis SF. Activation of recombinant Nr1/Nr2C NMDA receptors. The Journal of Physiology. 2008;586:4425–4439. doi: 10.1113/jphysiol.2008.158634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Ehlert FJ. Affinity and Efficacy: The Components of Drug-Receptor Interactions. World Scientific; 2015. [DOI] [Google Scholar]
  20. Ehlert FJ. Analysis of biased Agonism. Progress in Molecular Biology and Translational Science. 2018;160:63–104. doi: 10.1016/bs.pmbts.2018.08.001. [DOI] [PubMed] [Google Scholar]
  21. Engel AG, Lambert EH, Mulder DM, Torres CF, Sahashi K, Bertorini TE, Whitaker JN. A newly recognized congenital Myasthenic syndrome attributed to a prolonged open time of the acetylcholine-induced ion channel. Annals of Neurology. 1982;11:553–569. doi: 10.1002/ana.410110603. [DOI] [PubMed] [Google Scholar]
  22. Foreman JC, Johansen T, Gibb AJ. Textbook of Receptor Pharmacology. CRC Press; 2011. [DOI] [Google Scholar]
  23. Gharpure A, Noviello CM, Hibbs RE. Progress in nicotinic receptor structural biology. Neuropharmacology. 2020;171:108086. doi: 10.1016/j.neuropharm.2020.108086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Grewer C. Investigation of the Alpha(1)-Glycine receptor channel-opening Kinetics in the Submillisecond time domain. Biophysical Journal. 1999;77:727–738. doi: 10.1016/S0006-3495(99)76927-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Grosman C, Zhou M, Auerbach A. Mapping the conformational wave of acetylcholine receptor channel gating. Nature. 2000;403:773–776. doi: 10.1038/35001586. [DOI] [PubMed] [Google Scholar]
  26. Grosman C, Auerbach A. The dissociation of acetylcholine from open nicotinic receptor channels. PNAS. 2001;98:14102–14107. doi: 10.1073/pnas.251402498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Gupta S, Chakraborty S, Vij R, Auerbach A. A mechanism for acetylcholine receptor gating based on structure, coupling, Phi, and flip. The Journal of General Physiology. 2017;149:85–103. doi: 10.1085/jgp.201611673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Held M, Metzner P, Prinz J-H, Noé F. Mechanisms of protein-ligand Association and its modulation by protein mutations. Biophysical Journal. 2011;100:701–710. doi: 10.1016/j.bpj.2010.12.3699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Homans SW. Water, water everywhere--except where it matters? Drug Discovery Today. 2007;12:534–539. doi: 10.1016/j.drudis.2007.05.004. [DOI] [PubMed] [Google Scholar]
  30. Howard J. Mechanics of Motor Proteins and the Cytoskeleton. Sinauer Associates, Publishers; 2001. [Google Scholar]
  31. Indurthi D, Auerbach A. Agonist efficiency estimated from concentration response curve. Biophysical Journal. 2021;120:57a. doi: 10.1016/j.bpj.2020.11.577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Jackson MB. Kinetics of Unliganded acetylcholine receptor channel gating. Biophysical Journal. 1986;49:663–672. doi: 10.1016/S0006-3495(86)83693-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Jackson MB. Molecular and Cellular Biophysics. Cambridge University Press; 2006. [DOI] [Google Scholar]
  34. Jadey SV, Purohit P, Bruhova I, Gregg TM, Auerbach A. Design and control of acetylcholine receptor conformational change. PNAS. 2011;108:4328–4333. doi: 10.1073/pnas.1016617108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Jadey S, Auerbach A. An integrated catch-and-hold mechanism activates nicotinic acetylcholine receptors. The Journal of General Physiology. 2012;140:17–28. doi: 10.1085/jgp.201210801. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Jadey S, Purohit P, Auerbach A. Action of nicotine and analogs on acetylcholine receptors having mutations of transmitter-binding site residue Αg153. The Journal of General Physiology. 2013;141:95–104. doi: 10.1085/jgp.201210896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Jha A, Purohit P, Auerbach A. Energy and structure of the M2 Helix in acetylcholine receptor-channel gating. Biophysical Journal. 2009;96:4075–4084. doi: 10.1016/j.bpj.2009.02.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Kenakin T, Onaran O. The ligand paradox between affinity and efficacy: can you be there and not make a difference. Trends in Pharmacological Sciences. 2002;23:275–280. doi: 10.1016/s0165-6147(02)02036-9. [DOI] [PubMed] [Google Scholar]
  39. Kenakin T, Christopoulos A. Measurements of ligand bias and functional affinity. Nature Reviews. Drug Discovery. 2013;12:483. doi: 10.1038/nrd3954-c2. [DOI] [PubMed] [Google Scholar]
  40. Lape R, Colquhoun D, Sivilotti LG. On the nature of partial Agonism in the nicotinic receptor Superfamily. Nature. 2008;454:722–727. doi: 10.1038/nature07139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Lewis TM, Schofield PR, McClellan AML. Kinetic determinants of agonist action at the recombinant human glycine receptor. The Journal of Physiology. 2003;549:361–374. doi: 10.1113/jphysiol.2002.037796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Maconochie DJ, Zempel JM, Steinbach JH. How quickly can Gaba(A) receptors open. Neuron. 1994;12:61–71. doi: 10.1016/0896-6273(94)90152-x. [DOI] [PubMed] [Google Scholar]
  43. Masiulis S, Desai R, Uchański T, Serna Martin I, Laverty D, Karia D, Malinauskas T, Zivanov J, Pardon E, Kotecha A, Steyaert J, Miller KW, Aricescu AR. GABAA receptor signalling mechanisms revealed by structural pharmacology. Nature. 2019;565:454–459. doi: 10.1038/s41586-018-0832-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Mortensen M, Ebert B, Wafford K, Smart TG. Distinct activities of GABA agonists at Synaptic- and Extrasynaptic-type GABAA receptors. The Journal of Physiology. 2010;588:1251–1268. doi: 10.1113/jphysiol.2009.182444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Mukhtasimova N, Free C, Sine SM. Initial coupling of binding to gating mediated by conserved residues in the muscle nicotinic receptor. The Journal of General Physiology. 2005;126:23–39. doi: 10.1085/jgp.200509283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Nayak TK, Purohit PG, Auerbach A. The intrinsic energy of the gating isomerization of a neuromuscular acetylcholine receptor channel. The Journal of General Physiology. 2012;139:349–358. doi: 10.1085/jgp.201110752. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Nayak TK, Auerbach A. Cyclic activation of Endplate acetylcholine receptors. PNAS. 2017;114:11914–11919. doi: 10.1073/pnas.1711228114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Nayak TK, Vij R, Bruhova I, Shandilya J, Auerbach A. Efficiency measures the conversion of agonist binding energy into receptor conformational change. The Journal of General Physiology. 2019;151:465–477. doi: 10.1085/jgp.201812215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Nicolai C, Sachs F. Solving ion channel Kinetics with the Qub software. Biophysical Reviews and Letters. 2013;08:191–211. doi: 10.1142/S1793048013300053. [DOI] [Google Scholar]
  50. Nys M, Kesters D, Ulens C. Structural insights into Cys-loop receptor function and ligand recognition. Biochemical Pharmacology. 2013;86:1042–1053. doi: 10.1016/j.bcp.2013.07.001. [DOI] [PubMed] [Google Scholar]
  51. Phillips R. The Molecular Switch: Signaling and Allostery. Prinston University Press; 2020. [DOI] [Google Scholar]
  52. Pless SA, Lynch JW. Ligand-specific conformational changes in the Alpha1 glycine receptor ligand-binding domain. The Journal of Biological Chemistry. 2009;284:15847–15856. doi: 10.1074/jbc.M809343200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Purohit P, Auerbach A. Unliganded gating of acetylcholine receptor channels. PNAS. 2009;106:115–120. doi: 10.1073/pnas.0809272106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Purohit P, Gupta S, Jadey S, Auerbach A. Functional anatomy of an allosteric protein. Nature Communications. 2013;4:2984. doi: 10.1038/ncomms3984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Purohit P, Bruhova I, Gupta S, Auerbach A. Catch-and-hold activation of muscle acetylcholine receptors having transmitter binding site mutations. Biophysical Journal. 2014;107:88–99. doi: 10.1016/j.bpj.2014.04.057. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Qin F, Auerbach A, Sachs F. Maximum likelihood estimation of aggregated Markov processes. Proceedings. Biological Sciences. 1997;264:375–383. doi: 10.1098/rspb.1997.0054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Qin F. Restoration of single-channel currents using the segmental K-means method based on hidden Markov modeling. Biophysical Journal. 2004;86:1488–1501. doi: 10.1016/S0006-3495(04)74217-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Qin F. Hill coefficients of a Polymodal Monod-Wyman-Changeux model for ion channel gating. Biophysical Journal. 2010;99:L29–L31. doi: 10.1016/j.bpj.2010.05.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Rahman MM, Teng J, Worrell BT, Noviello CM, Lee M, Karlin A, Stowell MHB, Hibbs RE. Structure of the native muscle-type nicotinic receptor and inhibition by snake venom toxins. Neuron. 2020;106:952–962. doi: 10.1016/j.neuron.2020.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Rasaiah JC, Garde S, Hummer G. Water in Nonpolar confinement: from Nanotubes to proteins and beyond. Annual Review of Physical Chemistry. 2008;59:713–740. doi: 10.1146/annurev.physchem.59.032607.093815. [DOI] [PubMed] [Google Scholar]
  61. Richard JP. Enabling role of ligand-driven conformational changes in enzyme evolution. Biochemistry. 2022;61:1533–1542. doi: 10.1021/acs.biochem.2c00178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Sauguet L, Shahsavar A, Poitevin F, Huon C, Menny A, Nemecz À, Haouz A, Changeux JP, Corringer PJ, Delarue M. Crystal structures of a Pentameric ligand-Gated ion channel provide a mechanism for activation. PNAS. 2014;111:966–971. doi: 10.1073/pnas.1314997111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Schiebel J, Gaspari R, Wulsdorf T, Ngo K, Sohn C, Schrader TE, Cavalli A, Ostermann A, Heine A, Klebe G. Intriguing role of water in protein-ligand binding studied by neutron crystallography on Trypsin complexes. Nature Communications. 2018;9:3559. doi: 10.1038/s41467-018-05769-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Shi S, Lefebvre SN, Peverini L, Cerdan AH, Milán Rodríguez P, Gielen M, Changeux JP, Cecchini M, Corringer PJ. Illumination of a progressive allosteric mechanism mediating the glycine receptor activation. Nature Communications. 2023;14:795. doi: 10.1038/s41467-023-36471-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Sykes DA, Dowling MR, Charlton SJ. Exploring the mechanism of agonist efficacy: a relationship between efficacy and agonist dissociation rate at the muscarinic M3 receptor. Molecular Pharmacology. 2009;76:543–551. doi: 10.1124/mol.108.054452. [DOI] [PubMed] [Google Scholar]
  66. Traynelis SF, Wollmuth LP, McBain CJ, Menniti FS, Vance KM, Ogden KK, Hansen KB, Yuan H, Myers SJ, Dingledine R. Glutamate receptor ion channels: structure, regulation, and function. Pharmacological Reviews. 2010;62:405–496. doi: 10.1124/pr.109.002451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Tripathy S, Zheng W, Auerbach A. A single molecular distance predicts agonist binding energy in nicotinic receptors. The Journal of General Physiology. 2019;151:452–464. doi: 10.1085/jgp.201812212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Unwin N. Acetylcholine receptor channel imaged in the open state. Nature. 1995;373:37–43. doi: 10.1038/373037a0. [DOI] [PubMed] [Google Scholar]
  69. Yazdani M, Jia Z, Chen J. Hydrophobic Dewetting in gating and regulation of Transmembrane protein ion channels. The Journal of Chemical Physics. 2020;153:110901. doi: 10.1063/5.0017537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Young T, Abel R, Kim B, Berne BJ, Friesner RA. Motifs for molecular recognition exploiting hydrophobic enclosure in protein-ligand binding. PNAS. 2007;104:808–813. doi: 10.1073/pnas.0610202104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Zarkadas E, Pebay-Peyroula E, Thompson MJ, Schoehn G, Uchański T, Steyaert J, Chipot C, Dehez F, Baenziger JE, Nury H. Conformational transitions and ligand-binding to a muscle-type nicotinic acetylcholine receptor. Neuron. 2022;110:1358–1370. doi: 10.1016/j.neuron.2022.01.013. [DOI] [PubMed] [Google Scholar]

Editor's evaluation

László Csanády 1

This valuable work investigates the fundamental concept of how the energy of agonist binding is converted into the energy of the conformational change that opens the pore of a ligand-gated ion channel. The conclusions are based on analysis of solid data in terms of a mechanistic model. The findings will be interesting to biophysicists working on ligand-gated ion channels and, more generally, to enzymologists focused on allosteric enzyme regulation.

Decision letter

Editor: László Csanády1
Reviewed by: László Csanády2, Andrew JR Plested3

Our editorial process produces two outputs: (i) public reviews designed to be posted alongside the preprint for the benefit of readers; (ii) feedback on the manuscript for the authors, including requests for revisions, shown below. We also include an acceptance summary that explains what the editors found interesting or important about the work.

Decision letter after peer review:

Thank you for submitting your article "Agonist efficiency links binding and gating in a nicotinic receptor" for consideration by eLife. Your article has been reviewed by 3 peer reviewers, including László Csánady as the Reviewing Editor and Reviewer #1, and the evaluation has been overseen by Kenton Swartz as the Senior Editor. The following individual involved in review of your submission has agreed to reveal their identity: Andrew J R Plested (Reviewer #2).

The reviewers have discussed their reviews with one another, and the Reviewing Editor has drafted this to help you prepare a revised submission.

Essential revisions:

1. Generality

The authors repeatedly state that the concept of efficiency is generally applicable, but this is not demonstrated. Moreover, "efficiency" (eta) is defined in multiple different ways. E.g.: (i) In the Table 1 legend: "eta=1-logKdC/logKdO". That definition is certainly applicable to any receptor, as KdC and KdO values always exist. (ii) On line 325: "efficiency is the correlation between affinity and efficacy. It is a universal parameter that applies to every agonist of every receptor." Not only is the latter definition not shown to be universally applicable, but it is not even meaningful: the data presented here shows that affinity and efficacy are not "correlated", as their ratio can scatter between <0.3 and >0.6 (Figure 7B). Thus, the authors should more exactly define what they mean by "the concept of efficiency", and clearly spell out what is generally true vs. what is shown to be true specifically for the nAChR and only assumed to be true in general.

2. Definition of eta-classes

One concern is related to the statistical treatment of the measured and calculated values.

No statistical evaluation (SD, confidence intervals, p-values) is given for the calculated metrics- KdC, KdO, c, eta. Surely, the errors of the values used in calculations should propagate to the resulting calculus results. Uncertainty of these calculated metrics matter because based on them principal conclusions are made. It is not clear at all if variability of efficiencies would affect grouping of them as shown in Figure 7. Presenting the points with the error bars would give the reader a test of how confident one could be in groups of efficiencies.

The reviewers are concerned about the clustering of the data set into exactly five classes. First, the statistical treatment of the measured and calculated values is a major concern. The symbols in the efficiency plots (Figures 5a-b, 7a-b) lack error estimates: no statistical evaluation (SD, confidence intervals, p-values) is given for the calculated metrics KdC, KdO, DDG, or eta. Surely, the errors of the values used in the calculations should propagate to the resulting calculated variables. Uncertainty of these calculated metrics matter because based on them principal conclusions are made. It is not clear if the variability of efficiencies would affect grouping of them as shown in Figure 7. Presenting the points with the error bars would give the reader a taste of how confident one can be in the grouping of efficiencies. Second, even without error estimates, if the dots in Figure 5b were not color coded they would not appear to be segregated into exactly five clusters each of which fall on distinct straight lines. Third, there is no theoretical foundation for the existence of a discrete number of eta-classes as opposed to a continuum of possible eta values: the catch&hold LFER model presented in Figure 2 does not predict this.

3. Uniqueness of the catch&hold LFER model

The authors do not discuss the uniqueness of the proposed catch&hold LFER model (Figure 2) used for data interpretation. It seems that the existence of eta-classes might be explained just as well by an alternative model which assumes a single gating mechanism for the receptor, but distinct patterns of ligand-protein interactions for the different agonists (see Review Figure linked to the decision letter). This fact should be acknowledged, or if data exist to differentiate between these possibilities they should be presented and discussed.

4. Differentiating new from old data

The authors should clearly indicate what new data and what old data are included in each figure so the readers can judge the claimed advance. Replotting old data in a new way is acceptable, but throughout the manuscript it should be clearly spelt out what are actual new data and what are published data that have been replotted.

5. Clarity of presentation

All three reviewers have pointed out numerous places where the presentation is unclear. The paper would benefit from a major re-writing along the following specific guidelines: the authors should (i) provide only a single definition for a given variable (e.g., eta), (ii) avoid introduction of unnecessary parameters (e.g., kappa), (iii) expose the theory in a single block in the main text (e.g., in Intro or beginning of Results), and (iv) address specific comments regarding lack of clarity raised in all three reviews.

Finally, the reviewers feel that the attached movie does not provide additional information relative to that presented in the paper, and it is also much longer than a typical concise summary video. Thus, linking it to the manuscript seems unnecessary.

Reviewer #1 (Recommendations for the authors):

1. What exactly is plotted in Figure 7b? What does the x axis represent?

2. Normalization (line 577): "The free energies DeltaG(LA) and DeltaG(HA) (Figure 1B), proportional to logKdC and logKdO, are each a sum of a ligand-protein binding energy and a chemical potential that incorporates the energy consequence of removing the ligand from solution".

Shouldn't DeltaG(LA) and DeltaG(HA) actually represent standard free energies of binding, which are devoid of the log-concentration term? (The expression -RT*lnKeq gives standard free energy change.)

Reviewer #2 (Recommendations for the authors):

There are many interesting points made – it is not necessarily new to think of agonists like mutations but this paper underlines the similarity.

1. One criticism I have is exemplified by statements like the following:

line 214 "Every agonist of every receptor has an affinity, efficacy and efficiency."

line 39 "Recently, a third universal agonist property, efficiency (; eta), was introduced as the correlation between affinity and efficacy"

I know that these assertions make sense to the authors. On paper, they seem reasonable. But they are not tested. There is no general demonstration. There is no information that they apply to any other receptor. Of course, the senior author's lab has developed tools over decades to test these ideas in muscle nicotinic receptors. And this is a totally valid approach. But we have no information about whether this concept is universal or special to the nAChR.

I am not expecting the authors to go back and prove that this idea is general for this paper. But the statements could be toned down because they lack evidence.

As a side note, how would this apply for example to g-protein coupled receptors? Is it generally applicable to receptors with co-ligands, co-factors? Is it applicable for other types of receptors? Or is it limited to (in as much as I mean, only useful for) ion channel receptors with nice gating properties. I think the reader deserves an honest assessment. Can it be reasonably concluded that it is too difficult to assess these properties in other systems?

A related problem is that, as was discussed in the past, the citations are overwhelmingly papers from the senior author's own lab. I did not count the percentage. It's close to half, I think. I know that there is some justification for this, but it speaks against the generality of the work. It suggests that this concept is in fact rather narrow.

2. Overall I am not sure why the treatment oscillates in the paper between kappa and eta. It took me quite a long time reading to recognise that kappa = 1 – eta and this makes understanding this technical work a tick harder.

I would stick to one and only mention the other in exceptional circumstances.

3. line 202 We measured kappa and eta in AChRs having wild-type (wt) binding sites for 5 previously unstudied agonists

I think this is the wrong way to put it. Decamethonium has been studied a great deal (e.g. https://www.pnas.org/doi/pdf/10.1073/pnas.75.6.2994), as has SCh. Do you mean, comparatively understudied by single channel recording? But there were already a few papers…

4. It would be nice in Figure 3 to note the voltage, and the mutants that are employed. It's not the same mutant for each curve. I think these records should have been recorded at +70 mV (as described in the methods, to avoid block) but the openings are downwards. Is this right? Inward currents that were flipped? Sorry, I think I have something backwards but I don't understand.

5. line 341 The main results are as follows. (1) Empirically, efficiency is the log-log correlation between affinity and efficacy. It is a universal parameter that applies to every agonist of every receptor.

Is that a result that is shown in this paper? Same goes for result (2) in this list, which seems not to be a result from this paper but rather a postulate taken on the basis of existing results.

6. line 317 "Figure 7A shows a kappa-plot for all mutations that have been measured so far."

I'm confused. This panel includes results from previous work. There are >50 measurements of Kappa in Figure 7. But there are not 50 measurements of mutants in Table 2-supplement 1 and anyway nearly half of those in the table are repeats on different agonists. Can you be explicit about the mutants that are in Figure 7? The strength of the conclusion that the groups are the same for binding site mutations and for agonists rests on a more precise report of what is plotted here.

Reviewer #3 (Recommendations for the authors):

First of all, I would like to compliment the Authors for producing an impressive set of experimental data. The amount of single channel recordings produced is truly formidable.

One concern is related to the statistical treatment of the measured and calculated values. Only the EC50 and Po estimates have standard errors of the mean presented in Table 1. However, no statistical evaluation (SD, confidence intervals, p-values) is given for the calculated metrics- KdC, KdO, c, nu. Surely, the errors of the values used in calculations should propagate to the resulting calculus results. Uncertainty of these calculated metrics matter because based on them principal conclusions are made.

It is not clear at all if the variability of efficiencies would affect the grouping of them as shown in Figure 7. Presenting the points with the error bars would give the reader a test of how confident one could be in groups of efficiencies.

One example of the problems related to the statistical treatment is the following. A dissociation rate constant, koff, value used in calculations is 15,000 1/s. The Authors accept that it is approximate (lines 556 and 557), however, no standard deviation or confidence intervals are given. Surely, the error of this value should propagate into all calculated values where koff is used. On the other hand, koff was estimated and is approximately the same for few agonists tested in Jadey & Auerbach 2012 but is it the same for many more agonists in this study?

Another concern about this manuscript is that it is written in a very confused and convoluted style. The manuscript is very difficult to follow. The entire section of 'Theory' in the results and a corresponding section in the Methods should be united (for fluency of reading) and use more straightforward and exact language. Indeed, I might have misunderstood some of the arguments the Authors make and I'm not sure I fully understood the theoretical background.

Here are a few comments in order of appearance I wrote down before giving up. It is in no way an exhaustive list.

Line 36: It is not clear what 'Receptor theory' refers to.

Lines 38-39: efficiency- universal agonist property. Is it really universal? There is no supporting evidence for that beyond the AChR receptor analysed in this particular framework advocated by the authors.

Lines 46-47: activation to a first approximation… What is activation to a second and further approximations?

Lines 49-50: 'regions change structure and function conjointly'. It is not a very clear expression. Do authors say that conformation of both- binding site and gate- change concertedly/instantaneously? Does the binding site change conformation as well when spontaneous opening happens?

Line 53-55: sentence is very unclear. Is reference to Figure 1 supposed to show an increase in Po and membrane current? What exactly does 'membrane current' in this sentence mean? Is the reference to Figure 1 supposed to be to Figure 1A to demonstrate 'favorable binding free energy to O…'? Yet it is not clear from Figure 1A what it is.

Lines 55-57: very confusing sentence. 'Asymptotes of CRC' come out of the blue and a nonexpert reader could be left very disconcerted.

Lines 58-59: Without finishing proper introduction (Introduction will continue in a paragraph below) the Authors declare what their intention is. The intention 'is to estimate free energy changes associated with each step in the process'. At this stage we have no idea what 'the process' is and even less idea about the steps in 'the process'. The rest of this paragraph does not get much clearer.

Line 64: 'weak/strong ratio': ambiguous.

Lines 108-109: 'two apparently independent steps'- do the Authors mean events? Surely, there are more than two steps in the scheme in Figure 1B.

Figure 1, text and several equations: it is not clear why L is used for efficacy in this manuscript despite the fact that letter E is used in the field and, indeed, in an earlier paper by the same authors.

Words 'bind' and 'gate' seem to be used in a very liberal way.

eLife. 2023 Jul 3;12:e86496. doi: 10.7554/eLife.86496.sa2

Author response


  • New statistics Ln 258 (Figure 4A), 291-297 (Figure 4B), 345-349 (Figure 7A), 353-355 (Figure 7B). We used the x-means algorithm and AICc metric to select the optimal number as n=5 classes (Ln 666).

  • Figures4B and 7B. we note that the sd of each point is smaller than the symbol.

  • Ln 429. Number of classes

Ln 241 h is not sensitive to errors in EC50 and POmax. h is precise (a ratio of logs). Agonist classes that are poorly defined in Figure 4B are supported by the mutational classes in Figure 7.

The reviewers are concerned about the clustering of the data set into exactly five classes. First, the statistical treatment of the measured and calculated values is a major concern. The symbols in the efficiency plots (Figures 5a-b, 7a-b) lack error estimates: no statistical evaluation (SD, confidence intervals, p-values) is given for the calculated metrics KdC, KdO, DDG, or eta. Surely, the errors of the values used in the calculations should propagate to the resulting calculated variables. Uncertainty of these calculated metrics matter because based on them principal conclusions are made. It is not clear if the variability of efficiencies would affect grouping of them as shown in Figure 7. Presenting the points with the error bars would give the reader a taste of how confident one can be in the grouping of efficiencies. Second, even without error estimates, if the dots in Figure 5b were not color coded they would not appear to be segregated into exactly five clusters each of which fall on distinct straight lines. Third, there is no theoretical foundation for the existence of a discrete number of eta-classes as opposed to a continuum of possible eta values: the catch&hold LFER model presented in Figure 2 does not predict this.

Correct Ln 148 403 421.

3. Uniqueness of the catch&hold LFER model

The authors do not discuss the uniqueness of the proposed catch&hold LFER model (Figure 2) used for data interpretation.

Ln 420 (see below for generality of induced fit/clamshell). In AChRs, the catch-hold LFER is correct.

It seems that the existence of eta-classes might be explained just as well by an alternative model which assumes a single gating mechanism for the receptor, but distinct patterns of ligand-protein interactions for the different agonists (see Review Figure linked to the decision letter). This fact should be acknowledged, or if data exist to differentiate between these possibilities they should be presented and discussed.

Ln 141-145 We repeat the distinction between efficacy (λ) and efficiency (eta). The cartoons only depict ‘hold’ (in Figure 2, ACLA=ACHA). We are not sure what a “single gating mechanism” means (see below). We think that the contact ratio idea is too simplistic:

  1. AChR agonists are small and the physical basis of a ‘contact’ is not clear.

  2. It is extremely unlikely that every contact has an identical deltaG.

  3. Ln 170-190 Kd is set by ligand/protein/water rearrangements (induced fits) rather than only ligand-protein contacts.

  4. The cartoons do not incorporate catch (the slow kon to C, SI Figure 2 legend).

  5. Ln 444-449 Our experimental focus in this paper is energy and without data we limit our discussion of structure to what we learn from h. We need to have structures of AC, ACLA and ACHA w/ different agonists.

4. Differentiating new from old data

The authors should clearly indicate what new data and what old data are included in each figure so the readers can judge the claimed advance. Replotting old data in a new way is acceptable, but throughout the manuscript it should be clearly spelt out what are actual new data and what are published data that have been replotted.

Ln 86-88, 229 (agonists), 304-306 (mutations), Table titles (Main and SI).

5. Clarity of presentation

All three reviewers have pointed out numerous places where the presentation is unclear. The paper would benefit from a major re-writing along the following specific guidelines: the authors should (i) provide only a single definition for a given variable (e.g., eta), (ii) avoid introduction of unnecessary parameters (e.g., kappa),

See above; eta defined in Eq 2; kappa is gone.

(iii) expose the theory in a single block in the main text (e.g., in Intro or beginning of Results), and (iv) address specific comments regarding lack of clarity raised in all three reviews.

Finally, the reviewers feel that the attached movie does not provide additional information relative to that presented in the paper, and it is also much longer than a typical concise summary video. Thus, linking it to the manuscript seems unnecessary.

Ln 110-145 single block – affinity vs efficacy vs efficiency.

Ln 170-182 “catch” and “hold” replace the confusing “bind” and “gate”.

Reviewer #1 (Recommendations for the authors):

1. What exactly is plotted in Figure 7b? What does the x axis represent?

Figure 7B has been changed to DGLA vs DGHA. The DDG plot is now in the SI Figure 7.

2. Normalization (line 577): "The free energies DeltaG(LA) and DeltaG(HA) (Figure 1B), proportional to logKdC and logKdO, are each a sum of a ligand-protein binding energy and a chemical potential that incorporates the energy consequence of removing the ligand from solution".

Shouldn't DeltaG(LA) and DeltaG(HA) actually represent standard free energies of binding, which are devoid of the log-concentration term? (The expression -RT*lnKeq gives standard free energy change.)

Ln 110-116 DG defined and examples given.

Ln 191-203, SI Figure 7. Consideration of the chemical potential (binding entropy)

Reviewer #2 (Recommendations for the authors):

There are many interesting points made – it is not necessarily new to think of agonists like mutations but this paper underlines the similarity.

1. One criticism I have is exemplified by statements like the following:

line 214 "Every agonist of every receptor has an affinity, efficacy and efficiency."

deleted (replaced by Ln 146-151).

line 39 "Recently, a third universal agonist property, efficiency (eta), was introduced as the correlation between affinity and efficacy"

This statement is true.

I know that these assertions make sense to the authors. On paper, they seem reasonable. But they are not tested. There is no general demonstration. There is no information that they apply to any other receptor. Of course, the senior author's lab has developed tools over decades to test these ideas in muscle nicotinic receptors. And this is a totally valid approach. But we have no information about whether this concept is universal or special to the nAChR.

See above. We have toned down the language re: universality.

Ln 146 Eq. 2 If KdC and KdO are universal then by definition so, too, is h. If A and B are universal, then so is A/B.

Ln 556-558 What needs to be tested by experiment are whether efficiency classes and our physiochemicial interpretation of h are general.

Ln 402-405 KdC, KdO and kon to O have not been measured in other receptors.

Ln 424, 428 Structures of the key intermediate states have not been identified.

Ln 622 L0 (also mostly unknown in other receptors) is the key to measuring eta.

I am not expecting the authors to go back and prove that this idea is general for this paper.

We think our job is to drill deeper into the AChR rather than dig wider into other receptors. We hope others will measure efficiency in their receptor.

But the statements could be toned down because they lack evidence.

Language has been toned down.

As a side note, how would this apply for example to g-protein coupled receptors? Is it generally applicable to receptors with co-ligands, co-factors? Is it applicable for other types of receptors? Or is it limited to (in as much as I mean, only useful for) ion channel receptors with nice gating properties. I think the reader deserves an honest assessment. Can it be reasonably concluded that it is too difficult to assess these properties in other systems?

Ln 419 We mention a GPCR. Insofar as Figure 1 is general, then so, too is h. Any bistable protein activated by a difference in stimulus energy – GPCR, Hb, GroEL, , LGICs w/ cofactors, VGICs… – has an associated efficiency (the ratio of stimulus energies to off/on conformations). However, the existence of efficiency classes and generality of the induced fit mechanism need to be tested by experiment. Our hypothesis is that these, too, could be general (see above).

Ln 415-423 slow kon to C and h plots from published results (Nayak et al. 2019) hint that h classes exist in other receptors, including a GPCR. There are caveats regarding testing for h classes in other receptors:

  1. Measuring the energies accurately by experiment can be problematic. Affinities switch readily and it can be hard to measure KdC and KdO cleanly. Also, state models for some receptors can be complex. The place to start is L0.

  2. There could be additional binding energy changes buried in the data. However, downstream energy changes that do not involve a ligand energy change show up as an offset to the y axis of the h plot and do not affect the h estimate. Post-hold events (post-gating in GPCRs) that are the same energetically for all agonists simply translate the plot up/ down but do not change the slope.

  3. If there is an external energy source, or if occupancy of the site by an agonist has a global effect on the protein other than gating, then Figure 1 is void. The key assumptions in Figure 1 (the thermodynamic cycle) are (1) that differential ligand binding energies simply add to an intrinsic gating energy that is the same with or without the ligand, and (2) there is no significant external energy. In AChRs, both have been proved experimentally (Nayak 2017). Moreover, mutant analyses in AChRs show that energy changes (including from agonist binding) are independent and local (additive when separated by >12 A) (Gupta 2017).

We don’t discuss these caveats because we think they would be a distraction.

A related problem is that, as was discussed in the past, the citations are overwhelmingly papers from the senior author's own lab. I did not count the percentage. It's close to half, I think. I know that there is some justification for this, but it speaks against the generality of the work. It suggests that this concept is in fact rather narrow.

We do not agree. There can be other reasons for few external citations.

2. Overall I am not sure why the treatment oscillates in the paper between kappa and eta. It took me quite a long time reading to recognise that kappa = 1 – eta and this makes understanding this technical work a tick harder.

I would stick to one and only mention the other in exceptional circumstances.

kappa is gone.

3. line 202 We measured kappa and eta in AChRs having wild-type (wt) binding sites for 5 previously unstudied agonists.

I think this is the wrong way to put it. Decamethonium has been studied a great deal (e.g. https://www.pnas.org/doi/pdf/10.1073/pnas.75.6.2994), as has SCh. Do you mean, comparatively understudied by single channel recording? But there were already a few papers…

Ln 235 wording changed.

4. It would be nice in Figure 3 to note the voltage, and the mutants that are employed. It's not the same mutant for each curve. I think these records should have been recorded at +70 mV (as described in the methods, to avoid block) but the openings are downwards. Is this right? Inward currents that were flipped? Sorry, I think I have something backwards but I don't understand.

Currents are now inverted; Vm and the background mutations are noted in the legends.

Ln 625-632 The background is discussed in the Methods The mutation eS450W compensates exactly for the effect of depolarization on gating without altering binding. It was discovered long ago (Jadey and Auerbach, 2011) has been referenced in perhaps a dozen papers. With this mutation, the duration of single-channel outward current intervals at +70 mV (no channel block) are the same as inward intervals at -100 mV, which is to say long and therefore easy to measure.

5. line 341 The main results are as follows.

The list has been deleted from the Discussion.

1) Empirically, efficiency is the log-log correlation between affinity and efficacy. It is a universal parameter that applies to every agonist of every receptor.

See above.

Is that a result that is shown in this paper? Same goes for result (2) in this list, which seems not to be a result from this paper but rather a postulate taken on the basis of existing results.

See above.

6. line 317 "Figure 7A shows a kappa-plot for all mutations that have been measured so far." I'm confused. This panel includes results from previous work. There are >50 measurements of Kappa in Figure 7. But there are not 50 measurements of mutants in Table 2-supplement 1 and anyway nearly half of those in the table are repeats on different agonists. Can you be explicit about the mutants that are in Figure 7?

All are given in Table 2 and SI Table2.

The strength of the conclusion that the groups are the same for binding site mutations and for agonists rests on a more precise report of what is plotted here.

Reviewer #3 (Recommendations for the authors):

First of all, I would like to compliment the Authors for producing an impressive set of experimental data. The amount of single channel recordings produced is truly formidable.

One concern is related to the statistical treatment of the measured and calculated values. Only the EC50 and Po estimates have standard errors of the mean presented in Table 1. However, no statistical evaluation (SD, confidence intervals, p-values) is given for the calculated metrics- KdC, KdO, c, nu. Surely, the errors of the values used in calculations should propagate to the resulting calculus results. Uncertainty of these calculated metrics matter because based on them principal conclusions are made.

It is not clear at all if the variability of efficiencies would affect the grouping of them as shown in Figure 7. Presenting the points with the error bars would give the reader a test of how confident one could be in groups of efficiencies.

See above. The error bars in Figure 4B and 7B are all smaller than the symbol.

One example of the problems related to the statistical treatment is the following. A dissociation rate constant, koff, value used in calculations is 15,000 1/s. The Authors accept that it is approximate (lines 556 and 557), however, no standard deviation or confidence intervals are given. Surely, the error of this value should propagate into all calculated values where koff is used. On the other hand, koff was estimated and is approximately the same for few agonists tested in Jadey & Auerbach 2012 but is it the same for many more agonists in this study?

This discussion has been moved to the SI Figure 2. We calculated eta from equilibrium constants estimated from a CRC (we did not measure koff). This is a thought experiment and meant only to emphasize bind-gate entanglement. We are not suggesting that kon measurements should replace actual measurements of CRCs and tau. However, we think it’s worth noting that an association rate constant can predict gating responses.

Another concern about this manuscript is that it is written in a very confused and convoluted style. The manuscript is very difficult to follow. The entire section of 'Theory' in the results and a corresponding section in the Methods should be united (for fluency of reading) and use more straightforward and exact language. Indeed, I might have misunderstood some of the arguments the Authors make and I'm not sure I fully understood the theoretical background.

Here are a few comments in order of appearance I wrote down before giving up. It is in no way an exhaustive list.

We hope that our changes (the earlier introduction of catch-hold and induced fits) alleviate some of the confusion.

Line 36: It is not clear what 'Receptor theory' refers to.

Gone.

Lines 38-39: efficiency- universal agonist property. Is it really universal? There is no supporting evidence for that beyond the AChR receptor analysed in this particular framework advocated by the authors.

See above.

Lines 46-47: activation to a first approximation… What is activation to a second and further approximations?

Ln 479-485 In the original we used “first approximation” because the results suggest there are 5 C/O structural pairs and many intermediate transitions. Perhaps some confusion is caused by taking the simple state model (Figure 1B) and single-channel current traces (on-off) too literally. Rather, schemes and patch clamp recordings are approximations that do not reflect the reality of proteins. In a scheme, each capital letter (state) represents an ensemble and each arrow (transition) contains intermediates, both of which are invisible in recordings. Figure 2A and SI Figure 2 expand the standard scheme to allow us to interpret experimental measurements that are aggregates (DGLA, DGHA, kon, F). One needs to imagine 5 stable C/O pairs and jittering intermediate transitions between states having ns-us lifetimes. These can be inferred from energy measurements even if the corresponding structures have not yet been identified. The classic example of a useful, if undetected, intermediate state is Michaelis-Menton.

Lines 49-50: 'regions change structure and function conjointly'. It is not a very clear expression.

We don’t think this phrase is vague, but “conjointly” is redundant and therefore has been deleted.

Do authors say that conformation of both- binding site and gate- change concertedly/instantaneously?

The definition of these words depends on the time scale (see above). Neither applies on the ps time scale, but both probably apply on the ms scale. We think about the AChR in ns-ms range and so avoid these words.

Does the binding site change conformation as well when spontaneous opening happens?

Ln 475-478. We don’t know.

Line 53-55: sentence is very unclear. Is reference to Figure 1 supposed to show an increase in Po and membrane current? What exactly does 'membrane current' in this sentence mean?

Changed. Single channel PO is proportional to membrane current (whole cell current = nPOi, where i is the single channel current and n is the number of channels and equal to 1 in our experiments). This relationship is so standard we don’t think it’s worth mentioning.

Is the reference to Figure 1 supposed to be to Figure 1A to demonstrate 'favorable binding free energy to O…'? Yet it is not clear from Figure 1A what it is.

Figure 1 shows the same thing – the thermodynamic cycle – in 2 different ways. A, The blue lines in the landscape represent binding energies, stronger (longer) to O versus C. B, The cycle is easier to see when written as a reaction scheme. We don’t see how we can make be any clearer.

Lines 55-57: very confusing sentence. 'Asymptotes of CRC' come out of the blue and a nonexpert reader could be left very disconcerted.

Ln 59 Changed.

Lines 58-59: Without finishing proper introduction (Introduction will continue in a paragraph below) the Authors declare what their intention is. The intention 'is to estimate free energy changes associated with each step in the process'. At this stage we have no idea what 'the process' is and even less idea about the steps in 'the process'. The rest of this paragraph does not get much clearer.

Ln 56-57 wording changed.

Line 64: 'weak/strong ratio': ambiguous.

Ln 63 changed.

Lines 108-109: 'two apparently independent steps'- do the Authors mean events? Surely, there are more than two steps in the scheme in Figure 1B.

Ln 105-109 changed.

Figure 1, text and several equations: it is not clear why L is used for efficacy in this manuscript despite the fact that letter E is used in the field and, indeed, in an earlier paper by the same authors.

Some people complained about our previous use of E (some use it for energy). L is the standard for the gating equilibrium constant in VGICs (including BK channels that are activated by ligands). In this paper we are consistent do not think that the change will generate significant confusion, especially because we mainly consider Kds.

Words 'bind' and 'gate' seem to be used in a very liberal way.

Figure 1B These words are defined explicitly in the reaction scheme.

Ln 170-182 Figure 2 and SI Figure 2. We now use ‘catch’ and ‘hold’, the microscopic subevents inside bind and gate that are germane to eta. ‘catch’ and ‘hold’ may not be as memorable as ‘flip’ (perhaps because they are more complex/abstract), but they are informative and useful.

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    Table 1—source data 1. Agonist efficiency.
    Table 2—source data 1. Mutation efficiency.
    MDAR checklist

    Data Availability Statement

    The source data is available in the Tables and table source data files and referred to in the legend of the associated figure.


    Articles from eLife are provided here courtesy of eLife Sciences Publications, Ltd

    RESOURCES