Publication Cover
Molecular Physics
An International Journal at the Interface Between Chemistry and Physics
Volume 117, 2019 - Issue 20: 10th Liblice Conference on the Statistical Mechanics of Liquids
1,734
Views
10
CrossRef citations to date
0
Altmetric
Liblice 2018 Special Issue

Revisiting the Meyer-Overton rule for drug-membrane permeabilities

ORCID Icon & ORCID Icon
Pages 2900-2909 | Received 15 Dec 2018, Accepted 16 Mar 2019, Published online: 07 Apr 2019

ABSTRACT

Assessing the permeation rate of drug-like molecules across a lipid membrane is of paramount importance for pharmaceutical applications. While the Meyer-Overton rule relates the permeability coefficient to a few thermodynamic properties, its accuracy is limited by the homogeneity assumption. In this work, we extract an analogous relation for the permeation of a small solute through a lipid bilayer. This is obtained by systematically screening a subset of chemical space by means of high-throughput coarse-grained simulations, and relying on the accurate inhomogeneous solubility-diffusion model (ISDM). We connect the permeability coefficient of a compound to two molecular descriptors: partitioning free-energy and acid dissociation constant. In the ISDM the permeability is a functional of the potential of mean force. We discuss how – and when – combining together these profiles for a large variety of compounds can result in a smooth dependence of the permeability on the molecular descriptors. We focus on acidic molecules: for weak acids, the permeability largely depends on the main free-energy differences of the problem, which directly link to our molecular descriptors. For strong acids, the multivalued nature of the permeability requires to introduce a third variable, characterising the amphiphilicity of the small molecule.

Abbreviations: CG: coarse-grained; HTCG: high-throughput coarse-grained; PMF: potential of mean force; ISDM: inhomogeneous solubility-diffusion model; DOPC: 1; 2-Dioleoyl-sn-glycero-3-phosphocholine

GRAPHICAL ABSTRACT

1. Introduction

The strong binding affinity of a small drug-like candidate to the specific intracellular target, alone, does not guarantee its efficacy. In addition to toxicity issues, the efficacy can be severely compromised by the presence of prohibitively large time scales for the compound to cross the lipid barrier enclosing the cell environment [Citation1]. Understanding the transport mechanism of small molecules through lipid bilayers is therefore of fundamental importance for pharmaceutical applications.

Transmembrane transport processes are usually divided in two classes [Citation2]: (i) carrier-mediated processes, and (ii) passive permeation, i.e. diffusion of the permeant driven only by its concentration gradient and thermal fluctuations. Although both classes can contribute to the total transport of a solute, passive permeation is believed to play a dominant role in the case of small drug-like molecules [Citation3]. The permeability coefficient, P, quantifies the rate of passive permeation of a compound across the membrane. P is defined as [Citation4] (1) P=J/Δu,(1) where J is the steady-state flux per unit area of the solute arising from its concentration difference Δu between the two water compartments adjacent to the membrane. Equation (Equation1) makes it apparent that a compound characterised by a low permeability coefficient will experience more difficulties in crossing the lipid membrane, in this way reducing the amount of substance delivered to the intended intracellular target.

Pharmacokinetic issues associated to insufficient permeation are among the major causes of high attrition rates in drug discovery [Citation1]. Given the complexity of in vivo experiments, several in vitro methods have been developed to systematically estimate the permeability coefficient of small molecules, in order to screen them at early stages in the discovery process [Citation5–7]. These systems have contributed to the expansion of permeability databases due to the possibility of performing high-throughput experiments. At the same time, they typically provide limited insight into the microscopic origin of permeation.

From a theoretical point of view, the first model of solute transport across a lipid bilayer is the Meyer-Overton rule [Citation8,Citation9]. It approximates the membrane as a homogeneous slab, for which an application of Fick's first law of diffusion provides (2) P=KD/σ.(2) In Equation (Equation2), σ is the thickness of the bilayer core, while D and K are the diffusivity and partitioning coefficient of the compound, respectively. In practice, K is usually further approximated by the partitioning coefficient between water and a bulk apolar solvent (e.g. octanol).

Despite its simplicity, the Meyer-Overton rule identifies key properties associated with the permeability: due to the dependence on the partitioning coefficient, apolar compounds are characterised by higher permeation rates [Citation10]. However, its main limitation stems from the homogeneity assumption, at odds with the inherent asymmetry of a lipid bilayer. Accounting for the heterogeneity of the medium results in the inhomogenous solubility-diffusion model (ISDM) [Citation11,Citation12], which describes the diffusion process in terms of a one-dimensional Smoluchowsky equation along z, the normal distance of the compound from the bilayer midplane [Citation13]. From the ISDM, the permeability coefficient of a compound can be calculated as an integral over the membrane extension of its potential of mean force (PMF), G(z), and local diffusivity, D(z), as (3) P1=dzR(z)=dzexp[βG(z)]D(z),(3) with β=1/kBT. Equations (Equation2) and (Equation3) become equivalent only in the case of a flat potential of mean force and constant diffusivity – i.e. if homogeneity holds. The ISDM paved the way for an in silico estimation of permeability coefficients, where G(z) and D(z) are determined by means of enhanced-sampling classical molecular dynamics simulations [Citation13–19]. Predictions were shown to exhibit extremely high correlation with either in vitro or in vivo permeability measurements [Citation15,Citation20], making in silico methods a robust, complementary approach to experimental investigations. Unfortunately, the extensive computational resources required for convergence of the atomistic conformational sampling – roughly 105 CPU hours per chemical compound – have so far limited these calculations to few small molecules with very specific chemistries, see e.g. Ref. [Citation15,Citation20].

The importance of the Meyer-Overton rule lied in its ability to relate the permeability of a compound to a key molecular descriptor directly linked to the thermodynamics of the process. However, the rule has been recently shown to overestimate the permeabilities calculated through the ISDM, which were observed to be nonlinear in the partitioning coefficient [Citation19]. Moreover, in contrast to the ISDM the Meyer-Overton rule neglects the impact of the protonation state of a compound on its permeation rate [Citation15].

Given the accuracy of the ISDM, being able to connect its predictions to a set of simple thermodynamic properties would be of fundamental importance to obtain rapid estimates of permeability coefficients. Equation (Equation3) highlights that P is a functional of the diffusivity and potential of mean force. For small molecules D(z) was shown to be largely insensitive to the chemical detail [Citation15], so that the permeation rate is mainly affected by variations in the PMF. Extracting the link between thermodynamics and permeability in the ISDM would require to extensively probe at least a representative subset of the spectrum of possible drug-like molecules, to systematically investigate the impact of the PMF on the permeability coefficient. The computational resources associated to PMF calculations at an atomistic resolution, combined with the estimated size of drug chemical space – 1060 compounds [Citation21] – make this approach unfeasible in practice.

The overwhelming number of possible compounds thus calls for the introduction of new strategies that allow to more comprehensively cover large regions of chemical space. In this perspective, we recently proposed the use of coarse-grained (CG) models as a powerful instrument to preserve accurate thermodynamics while ‘clustering’ chemical space, in this way enabling a more systematic and efficient in silico exploration of chemical diversity [Citation22,Citation23].

In CG models, subsets of atoms are grouped together into elementary units (beads), with interactions tuned to capture a set of structural and/or thermodynamic properties of the original system [Citation24]. In the Martini CG force field, the fundamental building blocks are a few bead types that aim at reproducing the partitioning free energy of small organic fragments between solvents of different polarity [Citation25,Citation26]. Due to the nature of the properties chosen as target, Martini proved to be very successful in reproducing the thermodynamics of insertion of small molecules in lipid bilayers [Citation27,Citation28], at the same time significantly reducing the computational investment [Citation29].

The finite number of interaction types in the model is such that many chemically different small molecules with similar shape and polarity are mapped onto the same CG representation [Citation22]. In contrast to traditional in silico methods at an atomistic resolution, the small number of bead types further enables a more systematic exploration of the possible variations in the thermodynamics of a small molecule: this can be obtained by combinatorially constructing all coarse-grained compounds up to a certain size (Figure (a)).

Figure 1. (a): we employ high-throughput coarse-grained simulations to screen a representative subset of the chemical space of small organic molecules. The systematic insertion of this large variety of compounds with very different polarities in a lipid bilayer results in an ensemble of PMFs, which dictate their permeability coefficient. (b): we highlight PMFs from the two main classes of compounds analyzed in this work: polar (top) and apolar (bottom). (c): Construction of the effective PMF GT(z) (yellow area) contributing to the total resistivity RT(z) starting from the neutral (red) and charged (blue) PMFs of a compound. We highlight the main free-energy differences of the problem: the water membrane free-energy ΔGWM of the neutral compound, and the acid/base free energy difference ΔGba dictating the shift between the two PMFs, directly linked to the apKa.

Figure 1. (a): we employ high-throughput coarse-grained simulations to screen a representative subset of the chemical space of small organic molecules. The systematic insertion of this large variety of compounds with very different polarities in a lipid bilayer results in an ensemble of PMFs, which dictate their permeability coefficient. (b): we highlight PMFs from the two main classes of compounds analyzed in this work: polar (top) and apolar (bottom). (c): Construction of the effective PMF GT(z) (yellow area) contributing to the total resistivity RT(z) starting from the neutral (red) and charged (blue) PMFs of a compound. We highlight the main free-energy differences of the problem: the water membrane free-energy ΔGW→M of the neutral compound, and the acid/base free energy difference ΔGb→a dictating the shift between the two PMFs, directly linked to the apKa.

We recently took advantage of these features to shed new light on the thermodynamics of insertion of a small solute in a lipid membrane. In Ref. [Citation22] we employed high-throughput coarse-grained (HTCG) simulations to sample a representative subset of the chemical space of small organic molecules in the range 30−160 Da. Our comprehensive investigation highlighted that the overall shape of G(z) can be reconstructed only from the partitioning free energy ΔGWM – i.e. the free-energy difference of transferring the solute from bulk water to the bilayer midplane, see Figure (c).

In Ref. [Citation23], by feeding HTCG results into Equation (Equation3) and accounting for the possible protonation states of a compound we identified a smooth relation expressing the permeability coefficient as a function of two molecular descriptors: ΔGWM and acid dissociation constant, pKa. This enabled permeability predictions for an unprecedented 500,000 small molecules including acidic, basic, and zwitterionic compounds, whose accuracy was validated against experimental as well as atomistic simulation results.

Importantly, the relation in Ref. [Citation23] connects the permeation rate to only two thermodynamic properties, providing an alternative to the Meyer-Overton rule for predicting permeability coefficients in the framework of the accurate ISDM. Results were obtained by combining together permeabilities for compounds with highly different polarities, characterised by a large variety of neutral and charged potentials of mean force (Figure (a)). In this work, we analyze in more detail the impact of the molecular descriptors on the permeation rate, resulting – when possible – in a smooth dependence of the permeability coefficient on these thermodynamic variables. We here focus on acidic compounds, described by the acidic pKa or apKa, separately discussing neutral compounds (apKa8), weak acids (4apKa8) and strong acids (apKa4). For neutral compounds, the permeability is largely insensitive to the detailed shape of the underlying PMF, but mainly dictated by ΔGWM (Figure (c)) [Citation16]. This free-energy difference is not correctly accounted for in the Meyer-Overton rule, thus generating the observed discrepancies with ISDM predictions [Citation19]. Moving towards weak acids, the apKa comes into play. However, the insensitivity of the permeability on the shape of the PMF suggests that the permeation rate mainly depends on a combination of two free-energy differences: ΔGWM and ΔGba, where the latter is the free energy of neutralising the compound in bulk water, directly linked to the apKa (Figure (c)). This results in the smoothness of the permeability coefficient observed as a function of our thermodynamic variables. Finally, for strong acids, at fixed apKa the permeability coefficient does no longer smoothly depend on ΔGWM. This requires to introduce a third variable characterising the degree of amphiphilicity of the small molecule. In this case, our results provide a lower bound for the permeation rate of a compound: at a fixed apKa, the permeability coefficient cannot be lower than a value dictated by the combination of the two free-energy differences, ΔGWM and ΔGba.

2. Methods

2.1. Molecular dynamics simulations

Molecular dynamics simulations of the Martini force field [Citation25,Citation26,Citation30] were performed in Gromacs 4.6.6 [Citation31]. We employed an integration timestep δt=0.02τ, where τ is the model's natural unit of time. Control over the system temperature (T=300K) and pressure (P=1bar) was obtained by means of a Parrinello-Rahman barostat [Citation32] and a stochastic velocity rescaling thermostat [Citation33], with coupling constant τP=12τ and τT=τ respectively. A membrane of 36nm2 containing N=128 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) lipids (64 per layer), N=1890 water molecules, N=190 antifreeze particles [Citation26] and enough counterions to ensure charge neutrality was generated by means of the Insane building tool [Citation34], and subsequently minimised, heated up, and equilibrated.

We determined potentials of mean force G(z) by means of umbrella sampling simulations [Citation35]. For each analyzed compound, we employed 24 windows with harmonic biasing potentials (k=240kcal/mol/nm2) centred every 0.1nm along the normal to the bilayer midplane. In each window, two solute molecules were placed in the membrane in order to increase sampling and alleviate leaflet-area asymmetry [Citation36,Citation37]. Each production simulation was run for 1.2105τ, and potentials of mean force were extracted by means of the weighted histogram analysis method [Citation38–40].

2.2. Permeability coefficients

The permeability coefficient of a compound is obtained as an integral of its resistivity R(z) over the membrane normal, see Equation (Equation3). While CG simulations provide access to the potential of mean force G(z), the acceleration of the dynamics introduced by CG models [Citation41] makes the estimation of the inhomogenous diffusivity D(z) a priori difficult. However, atomistic simulations of several small molecules in a DOPC membrane showed that D(z) is largely insensitive to the chemical detail [Citation15]. Moreover, D(z) only provides a logarithmic correction to the log10P of a compound, see Equation (Equation3). In our calculations we thus relied on a unique diffusivity profile, and parametrised it as (4) D(z)=α+βeγ(xδ)+1,(4) with α=0.85106cm2/s, β=9.15106cm2/s, γ=7.5nm1, δ=3nm, which correctly captures the main features of its atomistic counterpart [Citation15].

In the case of acidic compounds, accounting for deprotonation requires to slightly modify Equation (Equation3) to include the flux of both neutral and charged species. The total resistivity reads [Citation15] RT(z)1=RN(z)1+RC(z)1, where RN and RC are the resistivities of the neutral and charged forms of the compound, respectively. The corresponding potentials of mean force GN(z) and GC(z) must be offset by the acid/base free-energy difference in bulk water [Citation36] (5) ΔGba=GacidGbase=kBT(pHapKa)ln10,(5) where we consider the case of neutral pH=7.4. Assuming that the diffusivity does not depend on the protonation state of the compound, the combination of the two PMFs in the total resistivity RT can be represented by an effective potential GT(z)=min[G¯N(z),G¯C(z)] for all z, see Figure (c), from which RT(z)exp[βGT(z)]. G¯i(z), i=N,C indicate the shifted counterparts of the neutral and charged PMFs, respectively.

Permeability coefficients predicted by the CG model were validated in Ref. [Citation23] against atomistic simulations results as well as experimental measurements for several chemically different small molecules covering a wide range of pKa. In both cases we observed a mean absolute error well within one log10 unit, demonstrating the accuracy of the CG model in reproducing the permeation rate of a compound.

3. Results

We considered all possible CG compounds consisting of a single Martini bead (‘unimers’) as well as those generated by binding together all possible combinations of two beads (‘dimers’), representative of a subset of the chemical space of small organic molecules in the range 30–160 Da [Citation22]. We calculated the PMF GN(z) for all neutral unimers and dimers, 119 in total. For each neutral compound, we then determined GC(z) for its deprotonated counterpart. At the CG level, deprotonating a chemical group amounts to replacing the original bead type with a negatively charged one. We always represent deprotonated fragments by means of a Qa type, following the standard Martini notation. In the case of unimers, this choice automatically determines the CG representation of the deprotonated form of a compound. In the case of dimers, we initially assumed that deprotonation always occurs in the chemical fragment represented by the bead of higher polarity, but relax this assumption along the discussion, vide infra. By combining GN(z) and GC(z) in the total resistivity RT(z), we calculated the permeability coefficient of every compound as a function of the apKa every 0.2 apKa unit. The set of permeability data was then projected on the (apKa,ΔGWM) plane.

To analyze how the resulting permeabilities are affected by a change in the two molecular descriptors, we separately discuss the case of neutral compounds/weak acids (apKa4) and strong acids (apKa4).

3.1. Neutral compounds/weak acids: revisiting the Meyer-Overton rule

Figure  displays a two-dimensional projection of the permeability data on the (log10P,ΔGWM) plane for neutral compounds and weak acids, apKa4. We present results for a subset of apKa values, and colour the data accordingly. Panels (a) and (b) distinguish between the permeabilities obtained for Martini unimers and dimers, effectively amounting to a segregation in molecular weights [Citation22].

Figure 2. Two-dimensional projection of the permeability data on the (log10P,ΔGWM) plane in the case of neutral compounds and weak acids, apKa4. Data are coloured according to the value of the apKa. Panels (a) and (b) distinguish between results obtained for Martini unimers and dimers. For neutral compounds with apKa8, the permeability becomes independent on the value of the acid dissociation constant. Accordingly, we only show data up to apKa=10. We present permeability coefficients obtained from HTCG simulations (points) and the corresponding interpolated permeability profiles (lines). For neutral compounds, we include predictions from the Meyer-Overton rule in the presence of flat potentials of mean force with heights equal to ΔGWM (‘Overton-mem’, dashed lines), see text.

Figure 2. Two-dimensional projection of the permeability data on the (log10⁡P,ΔGW→M) plane in the case of neutral compounds and weak acids, apKa⪆4. Data are coloured according to the value of the apKa. Panels (a) and (b) distinguish between results obtained for Martini unimers and dimers. For neutral compounds with apKa⪆8, the permeability becomes independent on the value of the acid dissociation constant. Accordingly, we only show data up to apKa=10. We present permeability coefficients obtained from HTCG simulations (points) and the corresponding interpolated permeability profiles (lines). For neutral compounds, we include predictions from the Meyer-Overton rule in the presence of flat potentials of mean force with heights equal to ΔGW→M (‘Overton-mem’, dashed lines), see text.

Due to the finite number of interaction types in the Martini model, simulation results are located at a discrete set of partitioning free energies. In both cases, the regularity of the points at fixed apKa as a function of ΔGWM allows to reconstruct continuous permeability profiles. We include these interpolations in the corresponding panels of Figure . The independence of the interpolated curves on the CG representation, together with the dense covering in ΔGWM generated by CG dimers, highlight the smoothness of the permeability coefficient as a function of apKa and ΔGWM. These results enable the calculation of the permeation rate only starting from two thermodynamic properties, providing a phenomenological alternative to the Meyer-Overton rule in the framework of the accurate ISDM.

The permeability coefficient weights the contributions of both the charged GC(z) and neutral GN(z) of a compound in an effective potential GT(z) which approximately dictates the shape of the total resistivity, lnRT(z)βGT(z) (Figure (c)). Understanding the smooth dependence of the permeability on both descriptors thus requires to more carefully analyze how the underlying potentials of mean force combine in GT(z). For this purpose, within the class of weak acids we separately discuss the case of neutral compounds – or ‘extremely weak’ acids – with apKa8, and weak acids, 4apKa8.

3.1.1 Neutral compounds: apKa8

For neutral compounds, the permeability is independent of the acid dissociation constant. This stems from the fact that by increasing the apKa beyond apKa8 the charged repulsive PMF GC(z) is shifted towards increasingly higher values, and does not contribute to the effective PMF (Figure (c)). One obtains GT(z)GN(z), such that RT(z)exp[βGN(z)].

For ΔGWM0, Figure  shows that log10P roughly plateaus. The reason is straightforward: apolar compounds are characterised by a negative, attractive well of GN(z) within the membrane core (Figure (b)). This well gives a negligible contribution to the integral in Equation (Equation3) compared to the regions in which the PMF reaches saturation – i.e. as the compound approaches the bulk water environment.

Moving towards ΔGWM0 the repulsive nature of GN(z) (Figure (b)) increases the resistivity RT(z), resulting in a reduction of the permeation rate. In this case, Figure  highlights a linear dependence of log10P as a function of ΔGWM. It was recently observed that for polar compounds the permeability is mainly dictated by the height of the repulsive PMF (i.e. ΔGWM) irrespective of its detailed shape – e.g. width, presence of attractive tails [Citation16]. This suggests the linearity in Figure  to be a consequence of this insensitivity.

We thus approximated the membrane as a bulk homogeneous solution, and employed the Meyer-Overton rule to calculate the corresponding permeabilities. In Equation (Equation2), we assumed K=KWM=exp[βΔGWM] and set σ,D equal to the average thickness and diffusivity values. The use of KWM amounts to replacing an ensemble of repulsive PMFs smoothly approaching zero by square-well potentials with heights equal to ΔGWM. We stress that KWM differs from the partitioning coefficient K in Equation (Equation2), which is instead calculated accounting for the inhomogeneous nature of the membrane, vide infra.

Permeabilities obtained from this homogenous model are included in both panels of Figure  (dashed lines). In the case of polar compounds results only slightly underestimate ISDM ones, the maximum deviation being roughly one log10 unit. This confirms that for ΔGWM0 the permeability mainly depends on the height of the repulsive PMF [Citation16], and explains the linearity of log10P as a function of ΔGWM. On the contrary, for ΔGWM0 the homogeneous model strongly overestimates the permeability coefficient: this stems from the fact that the square-well potentials neglect the saturation of the PMFs that occur by approaching the bulk water environment – the only region that for apolar compounds contributes to the resistivity.

This homogeneous model differs from the Meyer-Overton rule, in which the partitioning coefficient K of a compound reads (6) K=1σdzexp[βGN(z)].(6) Predictions from the rule were recently shown to overestimate ISDM ones [Citation19], and the permeability coefficient was observed to be multivalued in K. By approximating D(z) in Equation (Equation3) with its average value, a comparison with Equation (Equation2) highlights that in the latter we are replacing the inverse of the integral of exp[βGN(z)] with the integral of its inverse. The regions contributing to the two integrals are essentially swapped: for apolar compounds the negative well becomes the major contributor to Equation (Equation6), while in the ISDM its contribution is negligible compared to the region of saturation of the PMF. The opposite occurs in the case of polar compounds, where the free-energy barrier ΔGWM that dictates permeation rate in the ISDM does now not contribute to Equation (Equation6), which is instead governed by the regions where the PMF reaches saturation or exhibits attractive tails [Citation19]. This lies at the origin of the failure of the Meyer-Overton rule in reproducing ISDM permeabilities for neutral compounds.

3.1.2 Weak acids: 4apKa8

At fixed ΔGWM, Figure  shows that weak acids with 4apKa8 are characterised by a reduction in the permeability coefficient with respect to neutral compounds. In addition to polarity, the deprotonation of a small molecule affects its permeation rate.

By decreasing the apKa, both the neutral and charged PMFs of a compound contribute to the total resistivity. The neutral PMF GN(z) is shifted towards increasingly higher values by a factor of ΔGba, and the minimum between the two profiles dictates the shape of GT(z) (Figure (c) and Section 2.2), with RT(z)exp[βGT(z)]. We performed this construction for all compounds investigated in this work, resulting in the combination of a large variety of PMFs. However, Figure  highlights that at fixed apKa the only differences between the permeabilities of acidic and neutral compounds as a function of ΔGWM consist in the length of the initial plateau and in the presence of a negative offset for the linear regime.

To understand the origin of this behaviour, Figure  displays two examples of the construction of GT(z) for a polar and an apolar compound with apKa=5. For ΔGWM0 (main plot in Figure ), combining the neutral and charged PMFs results in a repulsive GT(z) with an effective barrier given by the sum of ΔGWM and ΔGba. As for polar compounds the permeability coefficient is mainly dictated by the height of the PMF, increasing this by a factor ΔGba – which is constant at fixed apKa and independent of ΔGWM, see Figure (c) – preserves the linearity of log10P in ΔGWM but generates the offsets observed in Figure .

Figure 3. Construction of GT(z) (blue) for an acidic compound from its charged GC(z) (red) and neutral GN(z) (orange) PMFs. In calculating GT(z), GN(z) is vertically shifted by ΔGba=kBT(pHpKa)ln10, and we consider apKa=5. We present results for a polar (main plot) and an apolar (inset) dimer.

Figure 3. Construction of GT(z) (blue) for an acidic compound from its charged GC(z) (red) and neutral GN(z) (orange) PMFs. In calculating GT(z), GN(z) is vertically shifted by ΔGb→a=kBT(pH−pKa)ln⁡10, and we consider apKa=5. We present results for a polar (main plot) and an apolar (inset) dimer.

For ΔGWM0 (inset in Figure ), the effect of a small vertical shift of GN(z) for a fixed apKa is to reduce the attractive well that in the neutral case was not contributing to the resistivity. By increasing ΔGWM, the well crosses zero and GT(z) becomes repulsive. At this stage, the permeability starts following the linear regime. This explains the reduction in the length of the plateau observed in the permeability profiles for ΔGWM0 by increasing the strength of the acid.

Overall, our results suggest that in the case of weak acids the permeability is mainly dictated by the combination of two free-energy differences: the water/membrane partitioning free energy ΔGWM, and the acid/base free-energy difference ΔGba, which directly links to the apKa (Figure (c)). This is observed irrespective of the detailed shape of the underlying PMFs, and explains the smooth dependence of the permeability observed as a function of the two molecular descriptors.

The construction of GT(z) allows for a further discussion about the role of the charged PMF of a compound in the permeation rate. Figure  shows that if GC(z) has reached the shifted neutral PMF within the membrane core – a saturation condition always satisfied for apKa4 – an increase in its amplitude does not affect GT(z). It follows that the permeability becomes largely independent on the polarity of the deprotonated form of the compound.

For a small molecule mapping to a CG dimer we originally assumed that deprotonation always occurred in the chemical fragment represented by the bead of higher polarity. We thus recalculated permeability coefficients relying on the opposite assumption – i.e. deprotonating the less polar bead. The resulting charged PMFs are characterised by a stronger repulsion within the membrane core, an example being shown in the inset of Figure .

Figure 4. Main plot: Permeability data of CG dimers (log10 scale) calculated by relying on different CG representations of the deprotonated form of a compound. Starting from a neutral dimer, coloured points are obtained by assuming that deprotonation always occur in the chemical fragment encapsulated in the bead of higher polarity. In dashed lines (‘Apolar-dep’) we always deprotonate the bead of lower polarity. Inset: Comparison of the charged PMFs GC(z) arising from the two different CG representations of the deprotonated form of the neutral polar dimer in Figure . The PMF of the compound obtained by deprotonating the bead of lower polarity is characterised by a stronger repulsion.

Figure 4. Main plot: Permeability data of CG dimers (log10 scale) calculated by relying on different CG representations of the deprotonated form of a compound. Starting from a neutral dimer, coloured points are obtained by assuming that deprotonation always occur in the chemical fragment encapsulated in the bead of higher polarity. In dashed lines (‘Apolar-dep’) we always deprotonate the bead of lower polarity. Inset: Comparison of the charged PMFs GC(z) arising from the two different CG representations of the deprotonated form of the neutral polar dimer in Figure 3. The PMF of the compound obtained by deprotonating the bead of lower polarity is characterised by a stronger repulsion.

Permeability coefficients obtained from the two deprotonation prescriptions are presented in Figure  (main plot). Results are essentially indistinguishable. Slight deviations are observed for apKa=4, in which the condition of saturation of GC(z) is not always satisfied as we are moving towards strong acids – vide infra. Importantly, this analysis shows that while the permeation rate could in principle depend on which chemical fragment of the small molecule is subject to deprotonation, this does not matter in practice.

3.2. Strong acids: the role of amphiphilicity

Panel (a) of Figure  displays permeability data of CG dimers as a function of ΔGWM for apKa=2, together with some previously analyzed results for apKa4. At apKa=2, data initially follow a linear trend analogous to the one of weak acids due to the saturation of GC(z) in the effective PMF. However, for ΔGWM0 we observe a spread of the points over the (log10P,ΔGWM) plane: the permeability coefficient at fixed apKa becomes multivalued in ΔGWM. This does not occur in the case of CG unimers, whose log10P continue to behave linearly as ΔGWM increases (data not shown).

Figure 5. Panel (a): Two-dimensional projection of the permeability data on the (log10P,ΔGWM) plane in the case of strong acids with apKa=2. We include as reference a subset of the data obtained for weak acids, apKa4. Points were obtained by assuming that starting from a neutral dimer, deprotonation always occur in the chemical fragment encapsulated in the bead of higher polarity. On the contrary, in dashed lines (‘Apolar-dep’) we always deprotonate the bead of lower polarity. Panel (b): Comparison of the neutral (main plot) and charged (inset) PMFs GN(z) and GC(z) of a P5C1 and a P2C5 dimers. In both cases, charged PMFs are obtained by assuming that deprotonation always occurs in the bead of higher polarity.

Figure 5. Panel (a): Two-dimensional projection of the permeability data on the (log10⁡P,ΔGW→M) plane in the case of strong acids with apKa=2. We include as reference a subset of the data obtained for weak acids, apKa⪆4. Points were obtained by assuming that starting from a neutral dimer, deprotonation always occur in the chemical fragment encapsulated in the bead of higher polarity. On the contrary, in dashed lines (‘Apolar-dep’) we always deprotonate the bead of lower polarity. Panel (b): Comparison of the neutral (main plot) and charged (inset) PMFs GN(z) and GC(z) of a P5C1 and a P2C5 dimers. In both cases, charged PMFs are obtained by assuming that deprotonation always occurs in the bead of higher polarity.

In strong acids, ΔGba is such that neutral PMFs are shifted towards high values with respect to their charged counterparts (Figure (c)): only GC(z) contributes to the effective PMF, and RT(z)exp[βGC(z)]. Different permeabilities for a given ΔGWM thus arise from compounds presenting very different polarities in their depronotonated forms, despite having similar ones in the neutral case – i.e. similar ΔGWM.

Martini assumes additivity in the partioning free energy of its constituting beads. Many possible combinations of two neutral beads can result in the same ΔGWM: while these will lead to roughly the same free-energy difference for the neutral PMFs GN(z), this does not necessarily hold for their deprotonated counterparts. As an example we consider a P5C1 dimer, composed by the most polar (P) and most apolar (C) Martini bead types, and a P2C5 dimer, in which we reduce the asymmetry in polarity between the two beads. Due to additivity, the compounds are characterised by similar neutral PMFs GN(z) (main plot in Figure (b)). Assuming that deprotonation always occurs in the bead of higher polarity, however, the charged form of the first compound shows a weaker repulsion from the membrane core (inset of Figure (b)). As RT(z)exp[βGC(z)], P5C1 is characterised by a higher log10P: 4.20 compared to 6.0 for P2C5.

The sensitivity of the permeability coefficient on the asymmetry in polarity between the chemical fragments composing the molecule suggests to introduce a third molecular descriptor. The most natural choice is the difference between the partitioning free energies of the two beads, ΔGasy=ΔG2ΔG1, where beads are arranged in order of decreasing polarity. For a given ΔGWM, increasing ΔGasy amounts to decreasing/increasing the polarity of the more apolar/polar chemical group. As such, ΔGasy is essentially quantifying the amphiphilicity of the small molecule.

Permeability data for apKa=2 as a function of ΔGWM and ΔGasy are presented in Figure . The introduction of ΔGasy completely solves the degeneracy in ΔGWM observed in Figure (a): for a given ΔGWM of the neutral form of the acid, increasing ΔGasy generates an increase in the permeation rate. On the one hand, this is not surprising, meaning that the reduction in permeability generated by the deprotonation of a functional group can be mitigated by coupling it with a highly apolar one. However, in analyzing these results one has to account for the tendency of Martini to underestimate the PMF of charged compounds [Citation27,Citation42]. We thus repeated our calculations by assuming that deprotonation always occur in the bead of lower polarity, which as discussed in Section 3.1 generates an increase in the repulsive barrier of GC(z). Results are included in Figure (a) (dashed line). In this case, the permeability coefficient follows a linear regime analogous to the case of weak acids due to saturation of the charged PMFs in GT(z). We conclude that if the permeability coefficient depends on the amphiphilicity of the small molecule, due to the underestimation of the PMF of charged compounds in Martini this will be observed at extremely low values of the apKa. For strong acids, however, our results provide a lower bound for the permeation rate: at a fixed apKa, the permeability coefficient cannot be lower than the linear relation dictated by combining two free-energy differences, ΔGWM and ΔGba.

Figure 6. Permeability coefficients (log10 scale) for CG dimers in the case of strong acids with apKa=2. We display results as a function of ΔGWM and the asymmetry ΔGasy. ΔGasy is calculated as the difference between the water/membrane partitioning free energy of the beads composing the dimer, sorting them in order of decreasing polarity. In all cases, the charged form of the compound was obtained by assuming that deprotonation always occur in the bead of higher polarity.

Figure 6. Permeability coefficients (log10 scale) for CG dimers in the case of strong acids with apKa=2. We display results as a function of ΔGW→M and the asymmetry ΔGasy. ΔGasy is calculated as the difference between the water/membrane partitioning free energy of the beads composing the dimer, sorting them in order of decreasing polarity. In all cases, the charged form of the compound was obtained by assuming that deprotonation always occur in the bead of higher polarity.

4. Conclusions

While the Meyer-Overton rule links the permeability of a compound across a lipid bilayer to the partitioning coefficient, its validity is limited by the homogeneity assumption. An analogous relation in the framework of the accurate inhomogeneous solubility-diffusion model (ISDM) was missing up to now. In the ISDM, the permeability coefficient is a functional of the potential of mean force (PMF). Identifying this relation would require to probe large regions of chemical space to systematically investigate the impact of the PMF on the permeation rate. This is hampered by the overwhelming number of possible compounds.

In this work, we rely on CG simulations to systematically explore the permeability variations in a subset of the chemical space of small organic molecules, focussing on acidic compounds. We connect permeability coefficients calculated from the ISDM to two thermodynamic properties: the water/membrane free energy, ΔGWM, and acid dissociation constant, apKa. Results are obtained by probing a large variety of small molecules with very different PMFs, both neutral and charged. We carefully analyze how – and when – combining together these profiles can result in a smooth dependence of the permeability on the two thermodynamic variables.

For neutral compounds, the permeation rate is independent on the apKa and smoothly depends on ΔGWM. This is due to the fact that the permeability coefficient is mainly dictated by ΔGWM irrespective of the detailed shape of the neutral PMF. The deviation of the Meyer-Overton rule from ISDM predictions stems from incorrectly accounting for this free-energy difference. For weak acids, the deprotonation of a chemical group affects the permeation rate of a compound. In this case, charged and neutral PMFs combine in an effective profile whose water/membrane free-energy difference is the sum of ΔGWM and the acid/base free-energy difference ΔGba, the latter being a linear function of the apKa. The insensitivity of the permeability coefficient on the detailed shape of the effective PMF results in a smooth dependence of the permeation rate on our thermodynamic variables. Interestingly, we show that in weak acids the permeability is independent of the chemical fragment subject to deprotonation. For strong acids, the permeability is no longer a smooth function of the two original molecular descriptors. This requires to introduce a third orthogonal variable characterising the amphiphilicity of a small molecule. In this case, our results provide a lower bound to the permeation rate of a compound: the permeability cannot be smaller than a linear relation generated by the combination of ΔGWM and ΔGba.

Acknowledgments

The authors thank Alessia Centi and Omar Valsson for critical reading of the manuscript and useful comments.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

This work was supported by the Emmy Noether program of the Deutsche Forschungsgemeinschaft (DFG). The authors gratefully acknowledge the computing time granted by the John von Neumann Institute for Computing (NIC) and provided on the supercomputer JURECA at Jülich Supercomputing Centre (JSC).

References