Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2012 Oct 11;109(45):18453–18458. doi: 10.1073/pnas.1208771109

Kinetics and thermodynamics of binding reactions as exemplified by anthrax toxin channel blockage with a cationic cyclodextrin derivative

Ekaterina M Nestorovich a, Vladimir A Karginov b, Alexander M Berezhkovskii c, V Adrian Parsegian d, Sergey M Bezrukov e,1
PMCID: PMC3494913  PMID: 23100532

Abstract

The thermodynamics of binding reactions is usually studied in the framework of the linear van’t Hoff analysis of the temperature dependence of the equilibrium constant. The logarithm of the equilibrium constant is plotted versus inverse temperature to discriminate between two terms: an enthalpic contribution that is linear in the inverse temperature, and a temperature-independent entropic contribution. When we apply this approach to a particular case—blockage of the anthrax PA63 channel by a multicharged cyclodextrin derivative—we obtain a nearly linear behavior with a slope that is characterized by enthalpy of about 1 kcal/mol. In contrast, from blocker partitioning between the channel and the bulk, we estimate the depth of the potential well for the blocker in the channel to be at least 8 kcal/mol. To understand this apparent discrepancy, we use a simple model of particle interaction with the channel and show that this significant difference between the two estimates is due to the temperature dependence of the physical forces between the blocker and the channel. In particular, we demonstrate that if the major component of blocker–channel interaction is van der Waals interactions and/or Coulomb forces in water, the van’t Hoff enthalpy of the binding reaction may be close to zero or even negative, including cases of relatively strong binding. The results are quite general and, therefore, of importance for studies of enzymatic reactions, rational drug design, small-molecule binding to proteins, protein–protein interactions, and protein folding, among others.

Keywords: protein-ligand binding, molecular docking


Motivated by the search for efficient small-molecule blockers of “virulent” transmembrane channels (13), we investigated the temperature behavior of the blockage reaction. The temperature dependence of the equilibrium constant of blocker association with the pore, that is, the reaction blocker(bulk) + channel(empty) Inline graphic channel Inline graphic blocker, was found to be surprisingly weak, with the temperature coefficient Inline graphic 1.08. Applying the standard linear van’t Hoff analysis to evaluate the enthalpy of an association reaction (4) to that of channel blockage (5), we arrive at an estimated enthalpy for the channel–blocker interaction of about 1 kcal/mol. However, examination of blocker partitioning between the bulk and the channel interior gives an estimate for the depth of the potential well for the blocker in the pore of at least 8 kcal/mol. Although these two estimates generally do not have to coincide, this impressive difference compelled us to take a detailed look at the assumptions that are often used in the thermodynamic analysis of blocking reactions as well as at the possible physical forces involved in the channel–blocker interaction. Our analysis is based on consideration of a simple model of particle interaction with the channel (68), which allows an explicit calculation of the involved thermodynamics. As might be expected, in the case of the temperature-independent flat potential well, the enthalpy of the binding reaction is equal to the depth of the potential well. However, in the realistic case of temperature-dependent interactions, the situation may be drastically different. In particular, we demonstrate that accounting for the temperature dependence of van der Waals and/or Coulomb forces may lead to very small or even negative enthalpy values. These conclusions go beyond the particular case of channel blockage to being helpful in the analysis of other problems such as protein–ligand binding, protein–protein interaction, and molecular docking for drug design, to mention just a few.

Results

The blockage of the anthrax PA63 channel by per-6-S-(3-amino)propylthio-β-cyclodextrin (AmPrβCD), a derivative of β-cyclodextrin carrying seven positive charges (13), is temperature-dependent. Fig. 1 gives three representative records of the ion current through a single channel at 15, 23, and 40 °C (Top to Bottom, respectively) in the presence of 0.3 μM AmPrβCD. It is seen that higher temperatures (i) speed up the kinetics of the blockage process and (ii) increase channel conductance. The increase in the on- and off-rates of the reaction manifests itself in the decrease of the average times spent by the channel in both open and blocked states. As temperature is elevated from 15 to 40 °C the conductance of the channel grows by a factor of 2, which is close to the effect of the temperature on ionic conductance of other large channels (9).

Fig. 1.

Fig. 1.

Typical recordings of an ionic current through a single PA63 channel in the presence of 0.3 μM AmPrβCD on the cis side of the membrane demonstrate temperature-dependent amplitudes and kinetics. Measurements were performed in 1 M KCl (pH 6.6) bathing solution at 50-mV transmembrane voltage and are presented at 10-ms time resolution. It can be seen that the current through the open channel and the frequency of blockage events increase with solution temperature, whereas the AmPrβCD binding time decreases.

Fig. 2 demonstrates the temperature dependence of the characteristic on- and off-times of the blockage reaction. Both the average time between successive blockages, Inline graphic, measured at the fixed blocker concentration of 0.3 μM, and the average time the blocker molecule spends in the channel, Inline graphic, decrease with increasing temperature. The temperature-induced drop in Inline graphic only slightly exceeds that in Inline graphic. The times were calculated from the characteristic corner frequencies of Lorentzian spectra and the independently measured probability of finding the channel in the blocked state, as described earlier (3).

Fig. 2.

Fig. 2.

Kinetic parameters of AmPrβCD-induced current blockage depend on the temperature of the bathing solutions. Both Inline graphic (time between successful blockages) and Inline graphic (binding time) decrease exponentially when the temperature increases. Error bars show SDs.

The plot of the natural logarithm of the equilibrium constant of the blocking reaction as a function of the inverse temperature, 1,000/T, the linear van’t Hoff plot (4), is given in Fig. 3. The equilibrium constant is calculated from the probability of finding the channel in the blocked state at a given blocker concentration, as described earlier (3). It is seen that within the accuracy of our measurements the dependence is linear and very weak. To illustrate this point, the figure also shows the conductivity of salt solution in the same temperature range, with Inline graphic 1.18.

Fig. 3.

Fig. 3.

The equilibrium binding constant of AmPrβCD to the channel shows only slight temperature dependence (solid circles), which is even weaker than the temperature dependence of salt solution conductivity (open squares). Error bars show SDs.

Discussion

Temperature Dependence of the Equilibrium Binding Constant.

Our analysis of the data in Fig. 3 gives Inline graphic 1.38 for the change of the equilibrium constant when the temperature of the system is varied from 10 to 50 °C. This leads to a Inline graphic value of about (1.38)1/4 Inline graphic1.08 that is significantly smaller than even the Inline graphic for solution conductivity. This weak dependence is usually interpreted as an anomalously low binding energy, which would correspond to a very shallow potential well for the blocker–channel attraction.

Indeed, if we fit the equilibrium constant data in Fig. 3 with the linear van’t Hoff equation widely used for the thermodynamic analysis of binding reactions (4, 10, 11),

graphic file with name pnas.1208771109eq1.jpg

where Inline graphic and Inline graphic are the absolute temperature and universal gas constant, respectively, and, if both enthalpy Inline graphic and entropy Inline graphic are assumed to be temperature-independent, we obtain Inline graphic 1 kcal/mol or about 2 Inline graphic per molecule, where Inline graphic is the Boltzmann constant. The enthalpy term is supposed to characterize the strength of the blocker–channel interaction and be equal to the depth Inline graphic of the temperature-independent flat potential well for the blocking particle inside the channel (see model considerations below).

This energy estimate is drastically smaller, by more than 10 Inline graphic, than the depth of the potential well estimated from blocker partitioning between the channel lumen and the bulk. For the depth of the potential well describing blocker partitioning under the simplifying assumptions detailed below, we have

graphic file with name pnas.1208771109eq2.jpg

where Inline graphic is the blocker concentration in the channel and Inline graphic is the blocker concentration in the bulk.

From the very beginning it is necessary to point out that Eq. 2 provides only the lower limit for the depth of the potential well describing the blocker–channel interaction. The main reason, as explained below, is that the notion of “blocker concentration in the channel” is a simplification based on the assumption of a homogeneous, square-well interaction potential for the blocker within the whole length of the blocker-accessible part of the channel. This concentration could be estimated by combining the data in Fig. 1, showing that channel occupancy by the blocker molecule is of the order of 1, with the existing predictions for the channel structure (12) illustrated in Fig. 4. However, even with the accepted simplifications, the estimate in Eq. 2 is inaccurate in the case of high occupancy, because of the strong repulsion between blocker molecules. To escape this additional complication, we make use of the data obtained at an order of magnitude lower blocker concentration in the bulk, Inline graphic 3 × 10−8 M (3). In this case the channel occupancy by the blocker, Inline graphic, which is equal to the average number of blocker molecules in the channel, Inline graphic, is much smaller than 1. Taking the values of Inline graphic 1.2 s and Inline graphic 0.11 s from figure 3 b and c of ref. 3, for this blocker concentration we have Inline graphic 0.084. To calculate the blocker concentration in the channel, Inline graphic, we need an estimate for the volume Inline graphic available for the center of the blocker molecule inside the channel lumen, Inline graphic, where Inline graphic is the pore radius, Inline graphic is the blocker radius, and Inline graphic is the blocker-accessible channel length; we assume that this part of the channel has a cylindrical shape. Based on the structure of the blocker (2) and structural predictions of ref. 12 for the PA63 pore reproduced in Fig. 4, we accept the following values for the parameters: Inline graphic 2 nm, Inline graphic 1 nm, Inline graphic 4 nm. This leads to Inline graphic 0.7 × 1025 m−3, which translates into the molar blocker concentration in the channel Inline graphic 1.2 × 10−2 M at Inline graphic 3 × 10−8 M in the bulk. Substituting these concentrations into Eq. 2, we obtain Inline graphic 13 Inline graphic or about 8 kcal/mol. It has to be noted here that this value gives only a lower estimate for the depth of the potential well, because the effective “binding volume” (13) of the blocker molecule can indeed be much smaller than the volume anticipated on the basis of the structure shown in Fig. 4.

Fig. 4.

Fig. 4.

Schematic representation of the PA63 channel and AmPrβCD (β-CD) drawn in CorelDRAW 11 (Corel Corporation) after the structure published in ref. 12. The rectangular well, two-rectangular well, and “binding site” and tilted potentials are plotted as explained in the Discussion. The vertical dashed lines indicate the part of the channel lumen accessible to the blocker molecule.

Thermodynamics of the Continuum Diffusion Model.

The huge difference in the two estimates above suggests that the physical forces responsible for the blocker interaction with the channel are temperature-dependent. Our further analysis is based on a simple model of particle interaction with the channel (68). We assume that (i) the maximal occupancy of the channel by a blocker molecule is one—to account for strong repulsion of the molecules carrying seven positive charges—and (ii) the blockage is a two-state Markov process, as is strongly supported by the simple Lorentzian shape of the spectral density reported earlier (3).

The equilibrium probability of finding the blocker in the channel, Inline graphic, can be written as a function of blocker concentration in the bulk, Inline graphic, and the equilibrium binding constant, Inline graphic, Inline graphic. Therefore,

graphic file with name pnas.1208771109eq3.jpg

where Inline graphic is the probability of finding the channel empty. On the other hand, Inline graphic, so that

graphic file with name pnas.1208771109eq4.jpg

where Inline graphic and Inline graphic are the on- and off-rate constants of the blockage, respectively, Inline graphic and Inline graphic.

The equilibrium constant can also be expressed through the partition functions in the blocker-occupied, Inline graphic, and blocker-empty, Inline graphic, states of the channel. Taking advantage of the fact that the ratio of the residence times entering into Eq. 4 is equal to the ratio of the partition functions of the occupied and empty states, Inline graphic, the equilibrium constant in Eq. 4 can be written as

graphic file with name pnas.1208771109eq5.jpg

The partition function for the empty state is given by

graphic file with name pnas.1208771109eq6.jpg

where Inline graphic is the partition function of a single particle (blocker molecule) in the system of volume Inline graphic (14) and Inline graphic is the number of particles in this system, so that the bulk blocker concentration is Inline graphic, Inline graphic is the blocking particle mass, and Inline graphic is the Planck constant. For the occupied state, the partition function is

graphic file with name pnas.1208771109eq7.jpg

Here Inline graphic is the configuration integral (14),

graphic file with name pnas.1208771109eq8.jpg

where Inline graphic, Inline graphic is a coordinate measured along the channel axis, Inline graphic, Inline graphic is the coordinate-dependent cross-sectional area of the channel, and Inline graphic is the coordinate-dependent potential of mean force describing particle interaction with the channel (68). Both the cross-sectional area and the potential of mean force are, in general, functions of temperature. It should be noted here that a rigorous evaluation of the configuration integral for a particular system must include the effects of the changes in all internal degrees of freedom, including those of the water molecules that are pushed out by the blocker molecule, the change in blocker and channel hydration, and so forth. However, for the purpose of the present analytical consideration, all these effects can be packed into the cross-sectional area Inline graphic and the potential of mean force Inline graphic, whose temperature dependences would then account for the above-mentioned effects. From Eqs. 57 it follows that

graphic file with name pnas.1208771109eq9.jpg

and the van’t Hoff equation is

graphic file with name pnas.1208771109eq10.jpg

where Inline graphic is the Avogadro number and the angular brackets denote Boltzmann averaging over the channel,

graphic file with name pnas.1208771109eq11.jpg

Temperature-Independent Rectangular Potential.

We first consider the case of a cylindrical channel of radius Inline graphic and length Inline graphic, assuming that the interaction with the blocker particle is described by a square-well potential of depth Inline graphic, which is temperature-independent (Fig. 4A). For such a channel, Eq. 10 takes the form Inline graphic, leading to

graphic file with name pnas.1208771109eq12.jpg

where Inline graphic is a reference temperature. It is seen that Eq. 12 yields a linear dependence of Inline graphic on the inverse temperature with the slope equal to the well depth that characterizes the strength of blocker–channel interaction, measured in Inline graphic units. Importantly, in this case, Inline graphic, that is, the depth of the potential well equals the enthalpy of the binding reaction.

Arbitrary Temperature-Independent Potential.

The square-well potential of mean force occupying the entire blocker-accessible part of the channel represents a grossly simplified situation. We will now consider a cylindrical channel with an arbitrary coordinate-dependent potential. In this case, it is easy to show that Eq. 1 does not hold true even for temperature-independent Inline graphic and a constant cross-sectional area. Indeed, from Eq. 10 it follows that the derivative Inline graphic is

graphic file with name pnas.1208771109eq13.jpg

where the Boltzmann-averaged potential Inline graphic is a function of temperature. This temperature dependence comes from the fact that at different temperatures the blocker molecule explores different parts of the potential well with different probabilities.

As an example, consider the case when the bottom of the potential well occupying the entire channel is a step with the parameters Inline graphic and Inline graphic and the corresponding depths Inline graphic and Inline graphic, as shown in Fig. 4B. In such a case, Eq. 10 leads to

graphic file with name pnas.1208771109eq14.jpg

This shows that Inline graphic is a function of temperature unless Inline graphic = Inline graphic and/or Inline graphic or Inline graphic is zero. As a consequence, Eq. 1 is inapplicable in this case.

Nevertheless, in many important cases of temperature-independent Inline graphic, the plot of Inline graphic versus Inline graphic in the narrow temperature range corresponding to functional protein states is remarkably linear. One of them corresponds to Inline graphic or Inline graphic. Another one is the case of a localized potential well—a narrow deep binding site, illustrated in Fig. 4C. Similar close-to-linear behavior is found for a tilted potential (Fig. 4D) when the slope of the tilt is large enough. Moreover, as we demonstrate below, even in the case of a temperature-dependent potential, which leads to Inline graphic for the square-well potential shown in Fig. 4A, the van’t Hoff plot may be nearly linear.

Temperature-Dependent Strength of the Interaction: Electrostatic, van der Waals, and Hydrogen-Bonding Forces.

In the considerations presented above, we assumed that the physical interactions, described by the potential of mean force, are temperature-independent. However, for the water-filled channel pore under consideration, this is hardly the case, and potential Inline graphic is a function of temperature. Then, in the simplest case of a cylindrical channel with the square-well potential of depth Inline graphic (Fig. 4A), Eq. 10 leads to

graphic file with name pnas.1208771109eq15.jpg

As we show below, the temperature dependence of the interactions determining the depth Inline graphic of the potential well is able to significantly change the slope Inline graphic and even invert its sign.

Experimental evidence, especially the strong reduction in binding strength with the increase in bathing solution salt concentration (3), suggests that the channel–blocker interaction involves Coulomb forces. If so, the depth of the potential is expected to be temperature-dependent. This is due to the temperature dependence of the water dielectric constant Inline graphic −1.35 at room temperature (15).

Let us consider a hypothetical case wherein the depth of the potential well is defined by Coulomb interactions between the “fixed” charges on the channel lumen and the charge on the blocker molecule. The interaction energy of two charges Inline graphic and Inline graphic positioned at distance Inline graphic in 1:1 electrolyte with number concentration Inline graphic m−3, where Inline graphic is the molar electrolyte concentration and Inline graphic is the elementary charge, is given by

graphic file with name pnas.1208771109eq16.jpg

where Inline graphic is the dielectric constant of vacuum, Inline graphic is the temperature-dependent ratio of the dielectric constant of the medium to that of vacuum, and Inline graphic is the temperature-dependent Debye length,

graphic file with name pnas.1208771109eq17.jpg

 Thus, the temperature behavior of the equilibrium constant, Eq. 15, depends on both the changing dielectric constant and the distance between charges. The temperature dependence of the dielectric constant for macroscopic samples of water is well-studied. If the values of Inline graphic are taken for bulk water, which are 84.1 at 10 °C and 69.9 at 50 °C (15), then the expected temperature dependence of Inline graphic not only gets weaker but is reversed. We plotted this dependence in Fig. 5 (solid line) in comparison with the dependence expected for the temperature-independent rectangular potential well of depth Inline graphic = 13 Inline graphic at 20 °C (dashed-dotted line). The solid line is calculated by substituting Eqs. 16 and 17 into 15 and taking the temperature-dependent Inline graphic from ref. 15. Other parameters are Inline graphic = Inline graphic = 7 Inline graphic, Inline graphic = 6.02 × 1026 m−3 (1 M KCl), and Inline graphic = 0.51 nm (to match 13 Inline graphic at 20 °C for the dashed-dotted line). For both curves, we use the same reference temperature, Inline graphic = 20 °C, to facilitate their comparison. It is clear that the formal application of Eq. 1 would give a negative enthalpic contribution even in the case of very deep potential wells, corresponding to the strong attractive interactions between the channel and the blocker.

Fig. 5.

Fig. 5.

The van’t Hoff plots of the temperature dependence of the equilibrium constant are very sensitive to a particular interaction responsible for the depth of potential wells shown in Fig. 4. The dash-dotted line shows Inline graphic calculated according to Eq. 12 with the temperature-independent potential well depth Inline graphic = 13 Inline graphic using Inline graphic = 20 °C as the reference temperature, indicated by the vertical dashed line. In the case of Coulomb interactions, the temperature dependence of the dielectric constant may significantly compensate the van’t Hoff slope and even lead to its inversion, staying approximately linear in this temperature range. This is illustrated by the solid curve calculated according to Eqs. 1517, as explained in the text. Temperature dependence of van der Waals (vdW) forces may induce slope corrections in the same direction, whereas in the case of hydrogen bonding the slope of the temperature dependence is expected to increase (arrows around the dash-dotted line).

In the case of van der Waals forces at room temperature, the corresponding Inline graphic is supposed to grow nearly proportionally to the temperature (16). Then, Eq. 15 leads to Inline graphic, and the enthalpy calculated from the linear van’t Hoff plot is close to zero. On the contrary, the forces due to hydrogen bonding may decrease linearly with temperature (17). Thus, according to Eq. 15, the slope of the temperature dependence of Inline graphic will increase, as illustrated by the counterclockwise arrow in Fig. 5. It should be noted that quantitative treatment of these two cases is complicated. The point is that the depth of the potential well for the blocker in the channel is the difference between the temperature-dependent van der Waals and hydrogen-bonding interactions of the blocker molecule with water in the bulk and with water and protein in the channel. Therefore, the explicit calculation of the slopes is not as straightforward as for Coulomb interactions in bulk water, and only tentative directions of slope changes, assuming predominance of interactions with the channel, are given in Fig. 5. The crucial importance of hydration effects is commonly recognized in the wide spectrum of biologically relevant processes from simple binding reactions (18) to macromolecular association (19) and protein folding (20).

Conclusions

Evaluating the depth of the potential well describing the blocker interaction with the anthrax channel, we have shown that the linear van’t Hoff analysis of the temperature dependence of Inline graphic may be grossly misleading (21, 22) because of the temperature dependence of the actual physical forces involved in the blocker–channel interaction. The enthalpy of the reaction is equal to the potential well depth only for temperature-independent interactions and a simple shape of the well. Importantly, in the narrow temperature range corresponding to the fully functional conformations of many proteins, Inline graphic can be nearly linear in the inverse temperature even if the interactions are temperature-dependent. This is illustrated by a hypothetical case of purely Coulombic interaction shown by the solid line in Fig. 5. We conclude that in the case of van der Waals or Coulomb interactions, one should generally expect a decrease or even a change in the sign of the slope of the van’t Hoff plot. At the same time, for interactions whose strength decreases with temperature, such as hydrogen bonding, the slope of the plot characterizing the apparent enthalpy increases.

Although the analysis given above deals with a relatively narrow subclass of binding reactions, namely channel blockage by a low–molecular-weight compound, the results obtained clearly go well beyond this example. For example, in many instances of theoretical analysis of experimental data on ion channels, the temperature dependence of channel conductance is used to estimate the number and structure of barriers and wells for ion translocation. Our results clearly demonstrate that a straightforward interpretation of the temperature effects in terms of the rate theory may be misleading if the temperature dependence of barriers and wells is ignored. In many cases of practical interest, the energy landscape seen by an ion in the channel is a sum of contributions due to interactions of different natures, as exemplified by the analysis of selectivity of potassium channels (23). As many of these interactions, similar to Coulomb forces in water, include an entropic component (24), caution is needed in interpreting experimental results. Only highly detailed approaches, such as all-atom molecular dynamics simulations, where temperature effects on all of the elements of the system are explicitly taken into account, can be free of this defect. However, this is not necessarily true for methods of multiscale modeling, which inevitably include certain simplifications of interaction potentials.

For the AmPrβCD interaction with the PA63 pore studied here, the involvement of electrostatics is made clear by the strong decrease in blocker binding at increasing salt concentration in the membrane-bathing solution (3). According to the arguments presented above, the temperature dependence of Inline graphic could be expected not only to be reduced but inverted. In reality, the temperature dependence of the forces creating the potential well for the blocker molecule in the channel is much more complex. Even if long-range electrostatics were the major driving force of the binding reaction, calculating the total electrostatic energy is not an easy task (25). Several effects should be taken into account. First, the Debye screening in the channel must be different from that in the bulk due to limitation of space and, therefore, of the number of available counterions in the channel cavity. Second, the dielectric behavior of water in the strong fields around a 3-nm–diameter particle carrying seven positive charges and in the predominantly negatively charged channel pore is significantly different from that of bulk water (26), including its dependence on temperature. Third, as the binding reaction takes place within the strong confinement of the channel-forming protein, protein dielectric properties and its conformational dynamics are of importance. Except for field-induced dielectric response saturation, these effects are able to factor in both directions, either increasing temperature sensitivity or decreasing it. Nevertheless, we hope that our thermodynamic analysis of this particular case—blockage of the anthrax PA63 channel by a cationic blocker—within the framework of a simple model will provide valuable insights for the broad community of researchers who work on similar problems, including those of structure-inspired design of new efficient drugs.

Materials and Methods

Reagents, lipid bilayer technique (after ref. 27), and methods of channel reconstitution were identical to those described earlier (3). Different temperatures were maintained with a thermostatic bath connected to a water jacket with the bilayer chamber (3) and controlled by a Cole-Parmer (model 92000-00) Digi-Sense scanning thermometer. AmPrbCD was custom synthesized at CycloLab (Budapest, Hungary) with the details of the synthesis given earlier (2).

Acknowledgments

We thank Donald Rau for fruitful discussions and reading the manuscript. This study was supported by the Intramural Research Program of the National Institutes of Health, Eunice Kennedy Shriver National Institute of Child Health and Human Development, and Center for Information Technology.

Footnotes

The authors declare no conflict of interest.

*This Direct Submission article had a prearranged editor.

References

  • 1.Karginov VA, Nestorovich EM, Moayeri M, Leppla SH, Bezrukov SM. Blocking anthrax lethal toxin at the protective antigen channel by using structure-inspired drug design. Proc Natl Acad Sci USA. 2005;102(42):15075–15080. doi: 10.1073/pnas.0507488102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Karginov VA, et al. Search for cyclodextrin-based inhibitors of anthrax toxins: Synthesis, structural features, and relative activities. Antimicrob Agents Chemother. 2006;50(11):3740–3753. doi: 10.1128/AAC.00693-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Nestorovich EM, Karginov VA, Berezhkovskii AM, Bezrukov SM. Blockage of anthrax PA63 pore by a multicharged high-affinity toxin inhibitor. Biophys J. 2010;99(1):134–143. doi: 10.1016/j.bpj.2010.03.070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Voet D, Voet JG. Biochemistry. New York: John Wiley; 1995. [Google Scholar]
  • 5.Arias HR. Temperature and ionic strength dependence of quinacrine binding and quinacrine displacement elicited by high concentrations of agonists on the nicotinic acetylcholine receptor. Arch Biochem Biophys. 1996;333(1):1–11. doi: 10.1006/abbi.1996.0357. [DOI] [PubMed] [Google Scholar]
  • 6.Berezhkovskii AM, Bezrukov SM. Optimizing transport of metabolites through large channels: Molecular sieves with and without binding. Biophys J. 2005;88(3):L17–L19. doi: 10.1529/biophysj.104.057588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Bezrukov SM, Berezhkovskii AM, Szabo A. Diffusion model of solute dynamics in a membrane channel: Mapping onto the two-site model and optimizing the flux. J Chem Phys. 2007;127(11):115101. doi: 10.1063/1.2766720. [DOI] [PubMed] [Google Scholar]
  • 8.Berezhkovskii AM, Pustovoit MA, Bezrukov SM. Fluxes of non-interacting and strongly repelling particles through a single conical channel: Analytical results and their numerical tests. Chem Phys. 2010;375(2-3):523–528. doi: 10.1016/j.chemphys.2010.04.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Biró I, Pezeshki S, Weingart H, Winterhalter M, Kleinekathöfer U. Comparing the temperature-dependent conductance of the two structurally similar E. coli porins OmpC and OmpF. Biophys J. 2010;98(9):1830–1839. doi: 10.1016/j.bpj.2010.01.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Boysen RI, Wang Y, Keah HH, Hearn MTW. Observations on the origin of the non-linear van’t Hoff behaviour of polypeptides in hydrophobic environments. Biophys Chem. 1999;77(2-3):79–97. doi: 10.1016/s0301-4622(99)00002-2. [DOI] [PubMed] [Google Scholar]
  • 11.Perozzo R, Folkers G, Scapozza L. Thermodynamics of protein-ligand interactions: History, presence, and future aspects. J Recept Signal Transduct Res. 2004;24(1-2):1–52. doi: 10.1081/rrs-120037896. [DOI] [PubMed] [Google Scholar]
  • 12.Katayama H, et al. GroEL as a molecular scaffold for structural analysis of the anthrax toxin pore. Nat Struct Mol Biol. 2008;15(7):754–760. doi: 10.1038/nsmb.1442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Gilson MK, Given JA, Bush BL, McCammon JA. The statistical-thermodynamic basis for computation of binding affinities: A critical review. Biophys J. 1997;72(3):1047–1069. doi: 10.1016/S0006-3495(97)78756-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Ben-Naim A. Water and Aqueous Solutions. New York: Plenum; 1974. [Google Scholar]
  • 15. Army Cold Regions Research and Engineering Laboratory Report. (2007) Available at www.tpub.com/content/ArmyCRREL/SR98_02/SR98_020009.htm.
  • 16.Parsegian VA. van der Waals Forces: A Handbook for Biologists, Chemists, Engineers, and Physicists. Cambridge, UK: Cambridge Univ Press; 2005. [Google Scholar]
  • 17.Dougherty RC. Temperature and pressure dependence of hydrogen bond strength: A perturbation molecular orbital approach. J Chem Phys. 1998;109:7372–7378. [Google Scholar]
  • 18.Gurnev PA, Harries D, Parsegian VA, Bezrukov SM. Osmotic stress regulates the strength and kinetics of sugar binding to the maltoporin channel. J Phys Condens Matter. 2010;22(45):454110. doi: 10.1088/0953-8984/22/45/454110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Stanley C, Rau DC. Evidence for water structuring forces between surfaces. Curr Opin Colloid Interface Sci. 2011;16(6):551–556. doi: 10.1016/j.cocis.2011.04.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Privalov PL. Thermodynamic problems in structural molecular biology. Pure Appl Chem. 2007;79:1447–1462. [Google Scholar]
  • 21.Naghibi H, Tamura A, Sturtevant JM. Significant discrepancies between van’t Hoff and calorimetric enthalpies. Proc Natl Acad Sci USA. 1995;92(12):5597–5599. doi: 10.1073/pnas.92.12.5597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Chaires JB. Calorimetry and thermodynamics in drug design. Annu Rev Biophys. 2008;37:135–151. doi: 10.1146/annurev.biophys.36.040306.132812. [DOI] [PubMed] [Google Scholar]
  • 23.Noskov SY, Roux B. Importance of hydration and dynamics on the selectivity of the KcsA and NaK channels. J Gen Physiol. 2007;129(2):135–143. doi: 10.1085/jgp.200609633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Jackson MB. Molecular and Cellular Biophysics. Cambridge, UK: Cambridge Univ Press; 2006. [Google Scholar]
  • 25.Gilson MK, Honig B. Calculation of the total electrostatic energy of a macromolecular system: Solvation energies, binding energies, and conformational analysis. Proteins. 1988;4(1):7–18. doi: 10.1002/prot.340040104. [DOI] [PubMed] [Google Scholar]
  • 26.Aguilella-Arzo M, Andrio A, Aguilella VM, Alcaraz A. Dielectric saturation of water in a membrane protein channel. Phys Chem Chem Phys. 2009;11(2):358–365. doi: 10.1039/b812775a. [DOI] [PubMed] [Google Scholar]
  • 27.Montal M, Mueller P. Formation of bimolecular membranes from lipid monolayers and a study of their electrical properties. Proc Natl Acad Sci USA. 1972;69(12):3561–3566. doi: 10.1073/pnas.69.12.3561. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES