3
views
0
recommends
+1 Recommend
0 collections
    0
    shares
      • Record: found
      • Abstract: found
      • Article: not found

      Forces Determining Ion Permeation

      letter
      The Journal of General Physiology
      The Rockefeller University Press

      Read this article at

      ScienceOpenPMC
      Bookmark
          There is no author summary for this article yet. Authors can add summaries to their articles on ScienceOpen to make them more accessible to a non-specialist audience.

          Abstract

          I would like to address two key issues related to the Perspectives on Ion Permeation. The first is the estimate of the size of the physical forces relevant for ion transport. Any good physical model of ion permeation requires the identification of the dominant forces. The second subject concerns a careful definition of the terms “mean field” and “mean field approximation.” Several articles in the field of ion permeation show considerable confusion about the mean field concept. The Size of the Forces What is the relative size of the forces that ions experience while passing through a channel? To enter the channel ions have to become partly, if not completely dehydrated. Highly polar and even charged functional groups forming the channel walls compensate for this loss of hydration energy. In addition, there are two kinds of strong ion–ion forces: long range electrostatic forces and short range hard core interactions leading to volume exclusion effects. Considerable simplifications are necessary to compute current–voltage curves of physical model channels. At present only two theories of ion transport are sufficiently simple for direct comparison between theory and experiment. These two theories consider different parts of the relevant forces as the most strong ones. In reaction-rate theories, also designated barrier models, the interaction of the ion with its environment and volume exclusion effects among ions are considered to dominate. Neglecting electrostatic ion–ion interactions, the rates of barrier crossing can be computed from first principles (Laio and Torre 1999). In contrast, Poisson-Nernst-Planck (PNP) theory assumes that electrostatic forces between the ions determine ion transport. Reaction-rate theories apply only if environmental forces surpass the electrostatic forces between ions. Otherwise, environmental interactions would not determine ion transport. Consequently, typical energies of electrostatic ion–ion interactions inside the channel represent lower limits for the energy differences between barriers and wells. Using a simple Coulomb law with a dielectric constant of 10, the energy required to bring two positively charged monovalent ions as close as 0.58 nm requires at least 250 milli electron volts (10 kT). Thus, environmental forces only dominate if the barrier energies are much larger. Reaction-rate theories explain the saturation of the channel conductance with increasing external concentrations by hard core ion–ion interactions. To experience the short range volume exclusion interactions, the ions must come rather close to each other. The distances of closest approach between ions including a single intermediate water molecule are in the order of 0.3–0.5 nm. Consequently, reaction-rate theories predicting conductance saturation automatically involve strong electrostatic ion–ion interactions. Thus, even single ion channels require large barriers to dominate these electrostatic forces between ions. A more general discussion of the forces important for ion transport was published recently (Syganow and von Kitzing 1999a). The identification of the dominant forces is crucial for understanding ion transport. Only those models of ion permeation that include the strongest interactions can provide a reasonable picture of what is going on inside biological ion channels. Mean Fields and Mean Field Approximation In his editorial, Andersen 1999 questions the applicability of the mean field approximation for the situation of ions in the narrow pore of the channel: “Finally, notwithstanding the utility of the mean field approximation, is it appropriate for narrow channels that are occupied by only a few ions?” Similar doubts are mentioned in other Perspectives. Therefore, a critical inspection of the concepts of mean fields and mean field approximation in statistical physics applied to biological ion channels is timely. Whereas the introduction of mean fields generally does not involve any approximation, the mean field approximation is necessary to account for the nonlinear long-range electrostatic interaction between permeant ions. During a 10-pA, 1-s channel opening, 6.3 · 107 ions pass the channel. With a time resolution in the millisecond range, the experimental mean current samples >60,000 ions. Passing the bottleneck of the channel each of those ions will see different forces. Side chains at the channel wall may change their orientations between the passage of two ions, and sometimes they even block the path. Also, the position of other ions differs at different ion passages, resulting in different electrostatic forces. For instance, the forces seen by an ion entering a channel differ considerably whether the channel is occupied by another ion or not. However, these very different forces add up linearly. If the channel is 50% occupied by other ions, on the average the incoming ion sees a half-occupied channel. The average of 105 different configurations results in a mean force. The absolute value of the force seen at each passage is generally large compared with that of the mean force. What is important for the measured mean current is the mean force, the linear average over all those very different contributions. We do not measure mean forces, but their integrals over the paths of the ions. Current–voltage relations represent integral properties of the channel (Syganow and von Kitzing 1999b). Many particularities of the channel structure are averaged out. This explains why extremely simple theories, such as reaction-rate theories with few barriers or PNP equations without particular structural elements, often can reproduce experimental data. Since we are measuring integral channel properties averaged over many configurations, mean fields are the appropriate physical tools to mathematically describe ion transport through biological ion channels. Unfortunately, the introduction of mean forces is not sufficient to handle the strong, long-range electrostatic ion–ion interactions. To describe the behavior of plasmas, Vlasov 1938 approximated the mean conditional force seen by a single ion by the mean force, the electric force of the mean charge density. This approximation generalizes the Gouy-Chapman theory (Gouy 1910; Chapman 1913) to nonequilibrium systems. What is the difference between the mean force and the conditional mean force? One of the oldest problems in physical chemistry is the nonlinear concentration dependence of the conductivity of electrolytes. In 1926, Onsager 1926, Onsager 1927 explained this effect by the difference between the mean force (due to the electric field across the electrolyte) and the conditional mean force (the mean force seen by a single ion). The nonlinear deviation originates from the fact that the motion and distribution of ions in solution is correlated. This correlation leads to two mechanisms: electrophoresis and electropolarization. Each moving ion pushes a part of the solvent molecules, and thus induces hydrodynamic ion–ion forces. This electrophoresis results in a modified effective mobility of the ions in the solvent (Hubbard 1987). The electric field deforms the counter ion cloud around each ion. This polarized counter ion cloud produces a local electric field that shields its host ion from the external electric field. Recently, Lehmani et al. 1997 included ion–ion correlations in the conductivity of ion-exchange membranes with large pores. Also within biological ion channels, one should expect effects because of ion–ion correlations. Solvent-mediated ion–ion interactions lead to single filing. This “electrophoretic” effect may be strong in channels such as the potassium channel. Consequently, PNP theories need to implement this mechanism, as for instance suggested by Conti and Eisenman 1966. Also, electropolarization may influence ion transport. Because of long-range strong electrostatic interactions, the occupancy of the channel depends on whether an ion is placed at the channel entrance or not. Therefore, the electric field due to the other mobile ions seen by the ion ready to enter differs from the field of the mean ion distribution at the same position. How reliable are these mean field approximations? In physics, such mean field approximations are frequently employed with different success. Unfortunately, the comparison (Cooper et al. 1985) between the ion concentrations obtained from PNP and Brownian dynamics cannot be used to evaluate the PNP solution because the two ion channel models employed different forms of electrostatic interactions (Syganow and von Kitzing 1999a). However, excellent tools to test the quality of mean field approximations are the so-called sum rules (Henderson 1992). These are exact relations generally derived for equilibrium systems; any exact theory strictly obeys these rules. There are two rules that are particularly relevant for homogeneous and inhomogeneous electrolytes (Blum and Henderson 1992). The screening sum rule state that all charges, dipoles, and higher multipoles within an electrolyte system are screened by respective counter charges, dipoles, or higher multipoles. It has been shown (Blum and Henderson 1992) that the PNP theory fulfills this rule exactly. The other important sum rule is the contact theorem. It relates the electric field at a certain boundary to the respective ion concentrations. PNP follows this rule only approximately (Henderson et al. 1979; Blum and Henderson 1992). The agreement becomes almost exact in the case of strong electric fields. Because both sum rules apply to equilibrium, they justify to some extent the neglect of the counter ion cloud polarization. However, they say nothing about the importance of single filing. Thus, in accord with the estimates of the size of the forces inside the channel, the PNP theory, using the mean field approximation for the electrostatic interactions, is the method of choice for modeling ion channels dominated by strong electrostatic fields. This theory obeys two important sum rules derived in statistical physics. The fact that the current–voltage relations used to compare theory with experiment are integral properties of the channel renders the judgment of the sum rules particularly valid. In contrast, commonly used reaction-rate theories obey none of these rules. Therefore, they cannot account correctly for strong, long range electrostatic interactions. Perspectives and Outlook McCleskey 1999 and Miller 1999 discuss the value of reaction-rate and PNP2 theories according to their ability to reproduce particular experimental current–voltage curves. They neglect the question of whether the basic physical assumptions of those theories are satisfied or not. As shown above, the applicability of any of the two theories depends critically on the size of electrostatic interactions compared with environmental forces. To study the fundamentals of ion permeation, simpler channels than the potassium or calcium channel should also be considered. For instance, the kind and position of mutations in the acetylcholine receptor channel (Konno et al. 1991) has no correlation with the obtained barrier energy profile. Such correlation, however, would be expected if this profile is assumed to represent the interactions of the ions with their environment. Random changes of the parameters of reaction-rate theories generally lead to nonlinear current–voltage curves (Levitt 1986). If reaction-rate theories represented the physics of ion channels such as the acetylcholine receptor channel, most single point mutations should result in nonlinear current–voltage curves. In contrast, nearly all mutations in this channel lead to fairly linear current–voltage curves. Such behavior is characteristic for ion channels dominated by electrostatic interactions (Syganow and von Kitzing 1999b). Today, the formulation of the Gouy–Chapman theory (Gouy 1910; Chapman 1913) is considered as an important first step in understanding the behavior of strong, inhomogeneous electrolytes. Since then, electrolyte theories have considerably improved (Blum and Henderson 1992). In the same sense, PNP theory should be considered as a first step to describe strong, long range electrostatic interactions in ion channels. The inclusion of ion size effects such as single filing must be one of the next steps.

          Related collections

          Most cited references20

          • Record: found
          • Abstract: found
          • Article: not found

          Interpretation of biological ion channel flux data--reaction-rate versus continuum theory.

          D Levitt (1985)
          Although the reaction-rate theory may provide a useful mathematical description of the channel flux, it presents a misleading physical picture of the channel structure. There is a tendency to regard the barriers in the model as actual physical structures, whereas they are actually only mathematical artifacts that allow one to reduce a complicated differential equation with an infinite number of states to a finite difference equation with a minimum number of states. I argue that the energy profile in the permeation pathway of most biological channels should vary relatively smoothly with only a few localized energy barriers or wells. In these smoothly varying regions, the resistance to ion movement is similar to bulk diffusion and cannot be accurately modeled by one or two energy barriers. For the one-ion channel, the continuum approach is as general and at least as simple as the reaction-rate theory and may provide a more physical interpretation of the data. Thus for the SR K+ channel, the structure suggested by the reaction-rate theory seems inconsistent with some experimental data, while the continuum-theory model is not only consistent with, but complements, the structure suggested by other data. Multi-ion channels have such complicated kinetics that one can only expect the theories to provide a qualitative description of the experimental data. They can be modeled by either the reaction-rate model or a finite difference approximation to the continuum model.
            Bookmark
            • Record: found
            • Abstract: found
            • Article: not found

            Physical origin of selectivity in ionic channels of biological membranes.

            This paper shows that the selectivity properties of monovalent cation channels found in biological membranes can originate simply from geometrical properties of the inner core of the channel without any critical contribution from electrostatic interactions between the permeating ions and charged or polar groups. By using well-known techniques of statistical mechanics, such as the Langevin equations and Kramer theory of reaction rates, a theoretical equation is provided relating the permeability ratio PB/PA between ions A and B to simple physical properties, such as channel geometry, thermodynamics of ion hydration, and electrostatic interactions between the ion and charged (or polar) groups. Diffusive corrections and recrossing rates are also considered and evaluated. It is shown that the selectivity found in usual K+, gramicidin, Na+, cyclic nucleotide gated, and end plate channels can be explained also in the absence of any charged or polar group. If these groups are present, they significantly change the permeability ratio only if the ion at the selectivity filter is in van der Waals contact with them, otherwise these groups simply affect the channel conductance, lowering the free energy barrier of the same amount for the two ions, thus explaining why single channel conductance, as it is experimentally observed, can be very different in channels sharing the same selectivity sequence. The proposed theory also provides an estimate of channel minimum radius for K+, gramicidin, Na+, and cyclic nucleotide gated channels.
              Bookmark
              • Record: found
              • Abstract: found
              • Article: not found

              Ionic Hopping Defended

              In the whirlwind of cloning, mutagenesis, and, suddenly, structure that the ion channel field has been riding for the past 15 yr, it is easy to forget that we still don't have a satisfactory view of that most basic task carried out by these proteins: ion permeation. The diffusion of ions through the aqueous pores of ion channels (a process much simpler than gating) is being treated in two different ways by two increasingly polarized schools of thought. For want of terms that are both precise and concise, I refer to these as the “chemical- kinetic” and “continuum” descriptions of channel- mediated electrodiffusion—both of which treat ions as stumbling through one-dimensional random walks along the pore. In chemical-kinetic descriptions, ions hop along a small number of binding sites (Hille, 1991); in continuum theories, they diffuse in continuous space along the pore under the influence of local electrochemical gradients (Sten-Knudsen, 1978). This distinction may seem the stuff of academic hairsplitting, but it is not—it is fundamental, and the vituperation spent on these models in recent years attests to their current irreconcilability. For the present discussion, I assume a qualitative understanding of the basic ideas behind the two competing views of ion conduction (but not necessarily the details of their implementation) and will offer some reasons why I consider the chemical-kinetic approach to be of greater practical utility. Rather than making general arguments about electrodiffusion to defend this position, I will illustrate the current clash of these theories by examining one particular case. An Experimental Example The subject we will look upon is calcium channel permeation, models of which have been described in lucid detail (Almers and McCleskey, 1984; Hess and Tsien, 1984; McCleskey, 1997; Nonner and Eisenberg, 1998). This case serves to bring out essential differences in a palpable and physiologically important context. I deal with just one small corner of this subject—the key observation that launched calcium channel permeation as a rich area of investigation. This is the remarkable fact that under physiological conditions the channel is strongly selective for Ca2+, but when bath [Ca2+] is reduced below 1 μM this selectivity is lost and monovalent cations easily permeate. In the well-known classical experiment, inward current through calcium channels is measured as a function of external Ca2+ concentration in a conventional physiological bath medium at a fixed holding voltage of, say, −30 mV. At very low Ca2+ (<0.1 μM) there is a large inward current carried by Na+. As [Ca2+] is raised up to ∼100 μM, the current decreases to near zero. But then, as [Ca2+] increases farther into the 1–10 mM range, inward current rises again, with Ca2+ as the charge carrier, and the reversal potential shifts positive towards ECa. This nonmonotonic variation in current with external [Ca2+], sometimes termed the “anomalous mole-fraction effect” (AMFE), is explained in vastly different ways by the two opposing viewpoints. Chemical Kinetic Viewpoint: Multiple Occupancy on Discrete Sites According to the canonical model, the observed AMFE is a direct reflection of the binding of two Ca2+ ions in a single-filing pore. The idea is simple, proceeding from the postulate that the channel is designed to coordinate Ca2+ at specific anionic sites. In the absence of Ca2+, when these are electrostatically hungry, the pore is merely charge selective, allowing virtually any monovalent cation to permeate as long as it is physically small enough to squeak through. Thus, at low [Ca2+], the Na+ conductance is high. In the presence of micromolar Ca2+ concentrations, the pore's selectivity region now becomes occupied by a single Ca2+ a significant fraction of the time (which varies according to the bath concentration). Because of its intimate coordination by protein groups, this bound ion's dissociation rate from the channel is low, ∼103 s−1, some three to four orders of magnitude slower than the throughput of Na+ ions. Under these conditions, Na+ roars through the pore when Ca2+ is absent; but whenever a Ca2+ binds, the flow of Na+ current is fully blocked. This block lasts on the order of 0.1–1 ms, and it is due directly to the single-filing property: the impossibility of a Na+ ion diffusing “around” a bound Ca2+. Only after the Ca2+ vacates the binding site can the flow of Na+ through the channel resume. This effect, averaged over many channels in a macroscopic experiment (or over time in a single-channel experiment), leads to the “falling phase” of the AMFE; i.e., the decrease of inward current as Ca2+ increases through the micromolar range. If Ca2+ concentration is pushed up into the millimolar range, a new phenomenon appears. Now a second Ca2+ can bind and, as a result of this double occupancy, the exit rate of Ca2+ from the pore increases ∼1,000-fold. This huge increase in Ca2+ off rate is usually explained by invoking electrostatic repulsion between the two ions, but other mechanisms could be involved (McCleskey, 1997). In any case, as a result of this double occupancy, Ca2+ now flows through the channel at rates high enough to show up as current, which increases with Ca2+ concentration to produce the “rising phase” of the AMFE. To quantify these effects, the chemical-kinetic approach makes an explicit distinction between four different occupancy forms of the channel: a Na+-conducting form with no bound Ca2+ [O, O], two nonconducting forms, each with one Ca2+ bound on either side [O, X] and [X, O], and a Ca2+-conducting form with two Ca2+ bound [X, X]. The average current and Ca2+/Na+ selectivity are given by the kinetic transitions among these various forms of the channel; i.e., by a set of rate constants between explicit chemical intermediates, exactly as in any conventional chemical kinetic problem. The values of the rate constants cannot be estimated from first principles, but must be derived by fitting experimental data to a kinetic model, a straightforward but rarely unambiguous procedure. Continuum Viewpoint: A Nanoscale Ion Exchanger Much of the present controversy centers on the use of Poisson-Nernst-Planck electrodiffusion models in biological channels. Such models have been in use for a long time, both as qualitative handles for the classical squid axon channels and as more intricate frameworks for permeation in ion-selective channels with firm structural foundations (Levitt, 1982, 1986). For this discussion, I will focus on a recent application to calcium channels termed PNP2 (Nonner and Eisenberg, 1998), which views the pore as a continuum containing several negatively charged groups smeared out over a reasonable pore volume; this represents a very high concentration of fixed charge (∼10 M). Both Ca2+ and Na+ have free access to this forest of negative charge, where they act as gegenions, cations that are not chemically coordinated by the fixed charge, but rather are held nonspecifically by the demands of electroneutrality (or, more properly, the Poisson equation), as in an ion- exchange resin. The system is described by the simultaneous solutions of three equations in which the three crucial variables, ion concentration, electrical potential, and distance along the pore, are nonlinearly entwined. The solutions lead to mathematically self-consistent predictions of ionic current as a function of transmembrane voltage and bath ion concentrations. In modeling the AMFE, PNP2 is a theory of ionic cleansing. With no Ca2+ present, the pore conducts well because Na+ ions dwell there at high concentration, this being the only cation available for electroneutrality. But when a little Ca2+ is added to the bath, these divalent intruders, with their heavy artillery in the form of a +2 valence, take over, displacing the numerous but poorly armed Na+ ions. In electrostatics, divalents always beat monovalents. Thus, as Ca2+ is increased, the pore, initially Na+ rich, becomes loaded with Ca2+, and the conductance goes down because the Ca2+ diffusion coefficient is assumed to be lower than that of Na+. By the time bath Ca2+ concentration reaches, say, 100 μM, all the Na+ has been expelled from the pore, and the current is carried solely by Ca2+. Thus, the falling phase of the AMFE. But why does the channel conductance rise again as Ca2+ is raised further? The surprising answer provided by the PNP2 treatment is that it doesn't! The conductance is predicted to remain essentially flat as Ca2+ rises to high levels because electroneutrality forbids admittance to additional Ca2+ over and above the fixed negative charge; but the current measured at a given voltage (e.g., −40 mV) does increase to produce the AMFE for a simple reason: the reversal potential keeps moving positive as external Ca2+ is increased. It's the driving force that goes up, not the conductance. In other words, this analysis asserts, everyone in the field has been dunderheaded all these years on a most elementary point, having apparently forgotten that current equals the product of conductance and driving force! (I am oversimplifying a little here; a small rise in conductance with [Ca2+] is predicted by PNP2, but this is a second- order effect having to do with surface polarization.) Evaluation and Conclusions So here we have two very different ways of interpreting a fundamental set of facts about ion permeation in calcium channels. I will state my opinion bluntly. First, no theory, however mathematically sophisticated, that rejects specific ionic coordination by protein moieties, dismisses the finite size of ions, and ignores the single-filing effects necessarily arising from the small spaces in the molecular structures of ion channels can have much worthwhile to say about selective ion permeation. Second, a ubiquitous feature of continuum theory— the mean-field assumption—invalidates, or at least greatly vitiates, its application to channels in which only a small number of ions reside at any one time. Third, PNP2 is inadequate to understand the particular calcium channel problem under examination here. Fourth, the undoubted quantitative weaknesses of the chemical-kinetic approach do not undercut its value in capturing the mechanistic essence of permeation in ion- selective channels. (1) The continuum approach ignores ion channel chemistry. For many years, indirect experiments have suggested that ions permeate selective channels by binding to localized sites at which protein functional groups replace waters of hydration, and that ion selectivity mainly reflects the energetics of the switch from water solvation to protein coordination (Hille, 1991). In soluble proteins, it is hardly a radical notion that binding of dehydrated inorganic ions lies at the basis of a multitude of functions (Falke et al., 1994), and now, with the structure of KcsA (Doyle et al., 1998), this idea has been confirmed directly for a strongly selective ion channel. The KcsA structure dramatically confirms for a K+ channel the multi-ion single-filing assumption, long known also to be valid for the peptide channel gramicidin A (Finkelstein and Andersen, 1981). For permeation, the qualitative consequences of localized, structured binding sites and single-filing are profound; they lead naturally and necessarily to familiar phenomena seen in many channels: strong, concentration- dependent selectivity, discrete ionic block of permeation, and anomalously high ratios of unidirectional ionic fluxes. It is not surprising that the continuum theories presently under discussion have been unable to satisfactorily reproduce these “enzyme-like” phenomena, since they (a) disregard the close-up chemistry of ionic coordination, (b) explicitly permit ions to move through one another within the pore, and (c) treat permeation mainly in terms of electric fields acting at a distance. This approach asserts the virtue of pristine, mathematically tractable physical principles, but it commits the vice of ignoring the messy parts: the prominent, obvious structural characteristics of channel proteins. To be sure, continuum theories have traditionally endeavored to include chemistry by superimposing upon the electrical potential a position-dependent free energy profile that may differ for different ions (Levitt, 1986) or by assigning to each ion its own diffusion coefficient. These are worthy additions to otherwise “featureless” electrodiffusion theories, but they are simply not enough; nobody has yet figured out how to weld single-file arrangements of binding sites to continuum theories in a general manner, although Levitt (1982, 1986, 1991a,b) has achieved impressive success in incorporating these features in particular cases, and Nonner et al. (1998) similarly have modeled a subset of K+ channel behaviors with a selective binding “region.” Conquest of this analytical impediment would represent a major advance in modeling permeation; in such a case, the entire channel field would unhesitatingly embrace continuum electrodiffusion as the preferred approach to the problem. (2) The mean-field assumption is inapplicable in small spaces. To obtain solutions for the ionic fluxes, continuum treatments must use concentration and electrical potential as continuous spatial variables in the coupled differential equations. The concentration at a given position determines the net charge density, which in turn influences the value of potential at that and nearby positions. Concentration is an intrinsically probabilistic quantity—the average number of ions per unit volume. For a macroscopic object such as a worm of ion-exchange gel, the number of ions present is sufficiently large that this average can be taken within each cross-sectional slab of the object at each moment in time. The concentration at each position will fluctuate with time, but, if the object is large enough, these fluctuations will be negligible and the concentration, and therefore the potential, will be a time-invariant spatial average. This is the “mean-field” assumption: this average potential may be validly used in the three crucial equations. In an object of molecular dimensions, however, a huge problem arises. For something the size of a calcium channel selectivity filter, a concentration of 10 M represents on average only one or two ions in the entire volume. “Concentration” is still defined as a statistical average, but in this case the average must be taken over time; i.e., by sitting at a given position in the pore and asking what fraction of the time an ion is present. This is a perfectly good stochastic definition of concentration, but when you try to use it to relate concentration to potential via the Poisson equation, a fundamental difficulty asserts itself. A channel containing, say, one Ca2+ on average will be fluctuating in occupancy among 0, 1, or 2 ions (0, 5, and 10 M concentration). The mathematics represents the channel as having a time-invariant potential equivalent to the average situation: single Ca2+ occupancy. But this is a severe misrepresentation of the potential that a Ca2+ approaching an empty channel, or a Ca2+ about to leave a doubly occupied channel, actually sees, and these events are often rate determining for permeation. It is as if the I.R.S. applied to every taxpayer a uniform exemption calculated for 2.6 children, the mean number of children per American family. Described another way, a Ca2+ aspiring to enter an empty channel at a given moment is treated by the electrodiffusion equations as though it experiences the repulsive electric field that existed, say, a microsecond before this moment, when the channel had one ion in residence; since occupancy-dependent changes in field are enormous and are established instantaneously, large errors in predicted behavior will arise from using an average potential. Thus, while valid for macroscopic objects and large, wide channels, the mean-field assumption applied to physically small channels yields solutions to the electrodiffusion equations that are mathematically chaste but physically debauched. The chemical-kinetic treatment avoids this problem by explicitly assigning distinct properties to the different occupancy forms of the channel. It asserts, for example, that the probability per unit time (i.e., the rate constant) of a Ca2+ entering an unoccupied channel is very different from the probability of entering a singly or doubly occupied channel precisely because of the very different electric fields in the three situations. This description deliberately avoids doing what continuum theory, for mathematical reasons, must do: treating the channel as a single entity with properties averaged over the different occupancy forms. (3) PNP2 misrepresents calcium channel behavior. The single example in the literature of a continuum treatment of calcium channel behavior, PNP2 (Nonner and Eisenberg, 1998), does not achieve the goal it sets for itself. The analysis is very similar to that of the classical macroscopic ion-exchange membrane, where electrodiffusion is well understood (Teorell, 1953). Emphasized in the analysis is that only the channel current at a fixed voltage, and not the conductance, is expected to show AMFE. Nonner and Eisenberg (1998) claim that published calcium channel experiments have demonstrated an AMFE only in current at fixed voltage, and that proponents of the standard view have merely assumed without evidence that the AMFE also applies to conductance, as chemical-kinetic theory says it must. This claim, if correct, would be a deadly criticism of the chemical-kinetic approach. But the claim is false. The original papers on calcium channel permeation (Figure 7 in Kostyuk et al., 1983; Figure 3 in Almers et al., 1984; Figure 2 in Almers and McCleskey, 1984) reported strong AMFE in current at fixed voltage as well as in conductance, based on macroscopic I–V curves over a [Ca2+] range from 60 nM to 10 mM. The conductance minimum is unambiguously observed at the single-channel level as well, in both Ca2+ and Ba2+ (Lansman et al., 1986; Friel and Tsien, 1989; Yue and Marban, 1990; Kuo and Hess, 1993). These elementary facts, well-known to the channel community, contributed mightily to the swift and widespread acceptance of the chemical-kinetic view of permeation. The PNP2 analysis proceeds as though these facts do not exist, and it accordingly fails to explain the most basic hallmark of calcium channel permeation. This incorrect prediction of a Ca2+-independent conductance at physiological concentrations illustrates how badly a continuum theory that uses the mean-field assumption and ignores coordination chemistry can falter. (4) Chemical kinetics preserves the basics. As for the weaknesses of the chemical-kinetic view, they are certainly prominent and well-appreciated (Cooper et al., 1985; Levitt, 1986; Dani and Levitt, 1990). It is impossible to predict a priori what the absolute values of the rate constants should be or how to relate rate constants to transition-state free energies. Likewise, the use of Eyring-like exponential voltage dependence to the rate constants is theoretically unjustified and always leads to incorrect I–V curve shapes. And physical space inside of channels is in fact continuous, not a lattice of sites. But so what? Most channel researchers don't really care about predicting absolute values of currents, just as enzymologists don't feel the need to calculate the k cat of an ATPase from quantum mechanics; it's the patterns of permeation behavior that count, not the absolute rates. As for the precise shapes of open-channel I–V curves, this is not a particularly compelling issue in channel physiology; the examples of unusual I–V shapes encountered in biologically meaningful contexts are invariably due not to intrinsic ionic diffusion properties, but rather to specific block (on discrete binding sites) by exogenous molecules (e.g., polyamine-induced inward rectification in K ir channels, or Mg2+-induced outward rectification by NMDA-receptor channels). And chemical kineticists don't believe that ions leap over tens of angstroms of pore length in a single bound; we do posit, however, in analogy to chemical reaction mechanisms, that sojourns on binding sites represent the preponderance of time the ion spends within the pore, and thus define the important rate- determining steps of ion permeation. Finally, there is a particularly compelling reason not to reject chemical kinetics in spite of its formal flaws: when used with an understanding of its limitations, it works. Its track record is excellent. It is primarily by chemical-kinetic analysis of ionic permeation over the past two decades that we have achieved physical pictures of ion channel proteins in the complete absence of direct structural information. It was chemical-kinetic analysis that told us that channels are built as axially symmetric structures with discrete selectivity filters and ion-binding sites at which ions are largely dehydrated, with narrow regions where ions and water lie in single file, with wide vestibules where drugs bind, and with enzymologically unprecedented regions where multiple ions bind simultaneously in close proximity. All of these features, which underpin the mechanisms by which ion channels achieve their paradoxical combination of selectivity and high transport rate, have now been observed directly in the first structure of a selective channel protein.
                Bookmark

                Author and article information

                Journal
                J Gen Physiol
                The Journal of General Physiology
                The Rockefeller University Press
                0022-1295
                1540-7748
                1 October 1999
                : 114
                : 4
                : 593-596
                Article
                8043
                2229472
                10577025
                2466947c-d88b-480f-a59b-a7ce56c632de
                © 1999 The Rockefeller University Press
                History
                Categories
                Letter to the Editor

                Anatomy & Physiology
                Anatomy & Physiology

                Comments

                Comment on this article