Abstract
The importance of charge-charge interactions in the thermal stability of proteins is widely known. pH and ionic strength play a crucial role in these electrostatic interactions, as well as in the arrangement of ionizable residues in each protein-folding stage. In this study, two coarse-grained models were used to evaluate the effect of pH and salt concentration on the thermal stability of a protein G variant (1PGB-QDD), which was chosen due to the quantity of experimental data exploring these effects on its stability. One of these coarse-grained models, the TKSA, calculates the electrostatic free energy of the protein in the native state via the Tanford-Kirkwood approach for each residue. The other one, CpHMD-SBM, uses a Coulomb screening potential in addition to the structure-based model Cα. Both models simulate the system in constant pH. The comparison between the experimental stability analysis and the computational results obtained by these simple models showed a good agreement. Through the TKSA method, the role of each charged residue in the protein’s thermal stability was inferred. Using CpHMD-SBM, it was possible to evaluate salt and pH effects throughout the folding process. Finally, the computational pKa values were calculated by both methods and presented a good level of agreement with the experiments. This study provides, to our knowledge, new information and a comprehensive description of the electrostatic contribution to protein G stability.
Introduction
A knowledge of the factors that affect the thermal stability of proteins is of fundamental importance in understanding the basic principles that govern the behavior of these macromolecules as well as in the rational development of biotechnological applications (1, 2, 3). It is widely known that a variation in pH and salt concentration can affect the thermal stability of proteins (4, 5, 6), indicating the importance of electrostatic interactions. Indeed, studies have shown the thermal stability of proteins can be enhanced by optimizing charge–charge interactions (7, 8, 9). Establishing a relationship between the distribution of ionizable residues and the features of folding and thermal stability can therefore provide valuable information for protein engineering.
Several levels of approximation have been proposed for protein models and for evaluating the electrostatic contributions of unfolding free energy (10, 11, 12, 13). In several of these studies, versions of structure-based models that include a fixed charge have been adopted to study the contribution of the electrostatic interaction in the unfolded state and the interaction between proteins and charged macromolecules (7, 14, 15, 16, 17, 18). Recently, to study pH effects on the folding dynamics of the N-terminal domain of ribosomal protein L9 (19), the constant-pH molecular dynamics (MD) method (CpHMD) (19, 20, 21) was implemented and the results showed a good level of agreement with the experiments. Another alternative to evaluate the charge-charge contribution to protein stability is the Tanford-Kirkwood model (TK) (22), which enables the individual electrostatic interaction of each ionizable residue to be calculated (23). Makhatadze and colleagues (24, 25, 26, 27) have managed to optimize proteins by the mutation of specific residues, which are predicted by the TK approach with a solvent accessibility modification, the Tanford-Kirkwood Solvent Accessibility (TKSA) method. In this study, both methods, CpHMD and TKSA, are used to study the importance of pH and salt concentration in 1PGB-QDD stability.
1PGB-QDD is a variant of the B1 domain of protein G with mutations T2Q, N8D, and N37D (Fig. 1 A). Protein G is a multidomain protein present in the cell wall of G streptococcus containing the immunoglobulin G binding domains denoted as B1 and B2 (28, 29, 30, 31, 32). These proteins differ by six amino acids and do not have disulfide bonds (33, 34). Also, experimental data showed that the protein folds/unfolds as a two-state model without kinetic intermediates (34). However, Sikosek et al. (35) have shown that a set of mutations may lead to new folds and as consequence, the protein folds/unfolds passing through an intermediate state. PGB1-QDD has 56 amino acids (including the NH2-terminal Met) and a regular α/β structure. The fold consists of a four-β-sheet and one α-helix bundled against the sheet. The charged residues are almost entirely surface exposed (36) and display distinct pH-dependence protein stability (33). Another important factor is that PGB1-QDD displays the highest stability at low salt concentrations (37). Due to its dependence on salt concentration and pH, this protein becomes an excellent target for studying electrostatic interactions using coarse-grained models. There are more rigorous and sophisticated ways to conduct constant pH simulations than using coarse-grained models (38, 39, 40). Nevertheless, these simplified models have a low computational cost, have been able to reproduce some experimental results, and allow an insightful overall view of the dominant folding mechanisms and effects (19).
Figure 1.

(A) Structure of PGB1-QDD is built from wild-type protein with PDB:1PGB. The side chains are shown for residues that have major contributions to protein stability. (B) Shown here is the primary sequence of PGB1-QDD with highlighted acidic residues, which have the highest charge variations in the studied pH range. To see this figure in color, go online.
Salt concentration and pH dependence in 1PGB-QDD stability were tested using the TKSA and CpHMD-SBM methods, and the results are in qualitative agreement with the experiments. The contribution of each ionizable residue to this protein stability was also calculated using the TKSA. From CpHMD simulations, it was possible to evaluate the dependence of the free energy unfolding barrier and thermal denaturation on pH. Electrostatic interaction influence at each folding stage was also analyzed. Finally, the pKa values were calculated for the folded protein (using CpHMD and TKSA) and the unfolded protein (using only CpHMD). The computational pKa values are in good agreement with the experimental values, considering the simplicity of the models. Thus, these coarse-grained models were found to be a computationally fast and cheap alternative to explore and provide useful insight into biological systems.
Methods
The TK model with solvent accessibility: TKSA
In the model proposed by Tanford and Kirkwood (TK), the protein is modeled by a spherical cavity with dielectric constant, εp, and radius, b, surrounded by an electrolyte solution treated according to the Debye-Hückel theory (22). A solvent static accessibility parameter for each charged residue was incorporated by Shire et al. (41) into the TK model. This modification was introduced to overcome the uncertainty of an adjustable charge-burial parameter beneath the dielectric interface and to allow for the irregular protein-solvent interface. With this change, the model proposed by Tanford and Kirkwood is now referred to as the “Tanford-Kirkwood model with a solvent accessibility” (TKSA) (41, 42, 43). The interaction energy, Uij, between two charged residues is given by
| (1) |
where e is the elementary charge; b is the radius of the protein; a is the closest possible approach distance of an ion; Aij, Bij, and Cij are parameters obtained from the analytical solution of the TK model, which are functions of the distance between charged residues, the dielectric constants, and the salt concentration (22, 44, 45); and SAij is the average of the solvent accessible surface area of residues i and j (41, 42, 46).
Once the electrostatic energy of the ionizable residues of the protein is determined, it is necessary to sample all the possible protonation states (23, 47). For a protein with n ionizable residues, there are χ = 2n different protonation states. The free energy, ΔGN(χ), for a protein in its native state and for given state of protonation χ is given by
| (2) |
where R is the ideal gas constant; T is the temperature in Kelvin scale; qi is the charge of the ionizable residue i in its deprotonated state; xi is 0 or 1, depending on the protonation state of the residue i; and pKa,ref,i is the reference compound pKa value of the ith ionizable residue (pKa,ref values are 3.6 for C-terminal, 4.0 for Aspartic acid, 4.5 for Glutamic acid, 6.3 for Histidine, 7.5 for N-terminal, 10.6 for Lysine, and 12.0 for Arginine).
The probability that the protein is in its native state with a particular state of protonation, ρN(χ), is
| (3) |
where ν(χ) is the number of ionizable residues that are protonated in the state of protonation χ, and ZN is the partition function for the protein in its native state.
The average total electrostatic energy, 〈Wqq〉, is given by the average over all protonation states, taking into account Eq. 3:
| (4) |
Ibarra-Molero et al. (23) and Makhatadze have shown that the contribution of electrostatic interaction energy to free energy, ΔGqq, can be described by the negative of the average total electrostatic energy of the native state, that is, ΔGqq ≈ −〈Wqq〉.
Constant-pH molecular dynamics simulations
In the structure-based model (SBM), the residues of a protein are represented by individual beads centered in Cα position (18, 48, 49, 50, 51). The energy of the protein is given by a Hamiltonian equation in which only the native interactions based on its native topology are taken into account (48). The first simplification of this model does not take into account the charged residues in the protein; the electrostatic interactions are therefore inserted by adding charged points at beads, which represent acidic/basic residues (7, 14, 19, 52). The interaction between charged residues is given by a Coulomb screening potential. The energy in a given configuration Γ with regard to configuration Γo, the native structure, is given by
| (5) |
where ro represents the distance between two subsequent residues, and θo and ϕo are the angles formed by three and four subsequent residues, respectively. All of them are taken from the native structure. Factor dij represents the distance between pairs of residues i and j (i < j – 3), εC = 2.5 kJ/mol, εr = 100 εC, εθ = 20 εC, εϕ = εNC = εC, σNC = 4 Å (48), and Kelectrostatic = 1390 kJ Å/(mol⋅e2) (7, 15, 19). The charges of residues i and j are qi and qj, respectively; κ is the inverse of Debye length (53); and εK = 80 is the dielectric constant.
Protonation/deprotonation criterion
The CpHMD method was utilized in this study. It associates a standard MD simulation with the Metropolis Monte Carlo method (MC) for sampling protein protonation states (19, 21, 54, 55, 56, 57, 58). During the simulation, the MD is regularly halted, a MC step is taken, and a random titratable residue is considered to change its protonation. The transition energy, ΔE, for this step is appraised according to Eq. 6, in which the positive and negative signs mean protonation and deprotonation, respectively. The acidic or basic residues are represented by ξ, where acidic residues present ξ = −1 and basic residues ξ = +1:
| (6) |
In Eq. 6, kB is the Boltzmann constant and T is the temperature. The first term is the contribution due to the presence of a proton bath at a given concentration determined by pH; the second one is the contribution of the compound model; and the last one, ΔEelec, is the electrostatic energy variation for the protonation state change of the titratable residue. The protonation/deprotonation is determined by the Metropolis criterion (59). Thus, if the MC step is accepted, the protonation/deprotonation of the residue will change to the new state, and the MD will continue. However, if the new state is not accepted, there is no need for updating, and the residue continues with its unchanged protonation state.
Titration curves and pKa calculation
The pKa values were calculated by fitting of titration curves using the Hill equation (60):
| (7) |
where α is the ionization degree of each acidic residue calculated by the average of the absolute valency (|Z|) generated by the simulations, and n is the Hill coefficient. This coefficient is also computed using the fit of the Eq. 7.
Simulation details
The CpHMD-SBM algorithm was implemented using Espresso Package version 3.2 in NVT ensemble and Langevin thermostat (61, 62). In the folding simulations, the protein was initialized in its native state with all titratable residues protonated. An equilibration process was executed with 107 steps and 109 steps to generate the trajectory with a time step of 0.5 fs. The data was stored every 1000 MD steps, and the Monte Carlo titration procedure was carried out every 2000 steps (19). The free energy and specific heat profiles were obtained using the weighted histogram method (63, 64, 65), calculated from different simulations at 25 temperatures for pH ranging from 1.0 to 12.0 for high and low salt concentrations. The reaction coordinate used to describe the folding was defined as the fraction of formed native contacts (Q). A native contact was considered formed when the distance between the residue i and j (i < j − 3) was shorter than 1.2 dij from the native structure. The native contact map was determined by contact-of-structural-units software (66). The mutant protein PGB1-QDD was built from PGB1 protein with PDB: 1PGB using Modeler software version 9.17 (67). The protein 3D structure graphic in Fig. 1 A was created with the UCSF Chimera package (68) and the protein sequence in Fig. 1 B was generated with Aline software (69). The TKSA method was implemented in PYTHON 2.7, and the solvent accessible surface area (SASA) was calculated with Surface Racer software (70). The TKSA method was performed for each of the 19 charged residues in PGB1-QDD and was simulated for pH from 1.0 to 12.0 at intervals of 0.5.
Results and Discussion
Electrostatic energy calculations: TKSA
In this section the contribution of electrostatic energy to protein free energy was calculated with the TKSA method. The sum of electrostatic energy of each charged residue is here referred to as ΔGelec. The ΔGelec of PGB1-QDD was calculated in a range of pH from 1.0 to 12.0, and the stability pH-dependent profile is presented in Fig. 2.
Figure 2.

Electrostatic energy contribution to free energy native state stability ΔGelec as a function of pH in kJ/mol. The values were calculated from TKSA simulation analysis for 23 different pH values from 1.0 to 12.0. The solid circles represent values for 0.15 M of a monovalent salt concentration in simulation and the open squares represent values of 2.0 M of salt concentration. (Inset) Plot shows the experimental results of the free energy difference between the unfolded and the native state ΔG° from 1PGB-QDD as a function of pH in kJ/mol. The solid circles represent values for 0.15 M of NaCl, and the open squares represent values for 2.00 M of NaCl (adapted from (37)).
The data presented in Fig. 2 of ΔGelec as a function of pH have similar behavior with slight differences for high and low salt concentrations in simulations via the TKSA method. The TKSA simulation results indicate that high levels of salt concentration do not eliminate the electrostatic interaction of surface charges and the pH dependence of protein native state stability is still strong (37). The peak of TKSA curves indicates that the protein PGB1-QDD native state is more stable around its isoelectric point, which is close to pH 4.5 for both high and low salt concentrations. The TKSA result has a level of good agreement with experimental data (37) for low salt concentrations shown as an inset in Fig. 2, and the correlation value between TKSA and experimental data is 0.83. On the other hand, for high salt concentrations, there is not a strong correlation between the TKSA and experimental results and the correlation value is 0.44. The weak correlation is due to the differences between the results for lower pH values. The behavior of experimental data for lower pH values does not present a significant stability difference from 2 M of NaCl (37), shown in Fig. 2 (inset). The correlation between TKSA and experimental data became stronger if only values of pH > 3.0 are taken into account and the new correlation value is then 0.95.
Another result obtained from TKSA simulation is the contribution of each charged residue to protein stability (23, 27, 71). Fig. 3 shows the TKSA analysis for pH values 2.5, 4.5, 7.5, and 10.0 for low salt concentrations. For pH 2.5 in Fig. 3 A, there are eight residues with ΔGqq < 0.0, which contribute favorably to protein stability. The other 11 residues have energy values close to 0 with ΔGqq ≈ 0.0. The total electrostatic free energy ΔGelec for pH 2.5 is −38.22 kJ⋅mol−1, and is calculated as the sum of ΔGqq of all charged residues (23, 24, 71). The major contribution to stability comes from six residues: K04, E15, E27, K31, D47, and K50. These six residues make an important energetic contribution to all pH values and they are responsible for 86 ± 6.2% of all favorable interaction for pH values 2.5, 4.5, 7.5, and 10.0, shown in Fig. 3. The side chains of these residues are not fully exposed to solvent, with SASA. However, the short distance between their ionizable groups of opposite charge (<2.95 ± 0.56 Å) promotes the high energetic contribution of these groups to protein stability (the side chains of K04, E15, E27, K31, D47, and K50 are highlighted in Fig. 1 A). At pH 4.5 in Fig. 3 B, similar to pH 2.5 analysis, the major electrostatic energy contribution comes from the six residues mentioned above. At pH 4.5, there are 12 residues with ΔGqq < 0.0 and the other seven residues have energy values close to 0, with ΔGqq ≈ 0.0. The total electrostatic free energy ΔGelec for pH 4.5 is −44.09 kJ⋅mol−1, which is the pH that maximizes protein stability in experiments (37) and also in the TKSA simulation shown in Fig. 2. For pH 7.5, shown in Fig. 3 C, there are also 12 residues with ΔGqq < 0.0 similar to pH 4.5. There are seven residues with ΔGqq > 0.0 contributing with unfavorable electrostatic interactions to native state stability. These seven residues—D08, D36, D37, D40, E42, D46, and E56—are all acidic residues in their deprotonated state with charge −1. The protonation/deprotonation events are detailed in the pKa comparison section. The positive ΔGqq values of these seven residues make the greatest contribution to decreasing the native state stability of PGB1-QDD, and the total electrostatic free energy ΔGelec for pH 7.5 is −35.63 kJ⋅mol−1. In Fig. 3 C, the red bars highlight the residues whose side chains are exposed to solvent >50% with SASA = 73 ± 14% for residues D36, D37, D40, and E42. The positive ΔGqq energy contribution and the side chain exposed >50% to solvent indicate that mutations in these residues should lead the protein to a more stable native state based on optimization of the electrostatic interactions on the protein surface (23, 24, 25, 26, 27, 71, 72, 73). The TKSA result for pH 10.0 is shown in Fig. 3 D with nine residues ΔGqq < 0.0, three residues with ΔGqq close to 0, and seven residues with ΔGqq > 0.0. The discussion of the favorable and unfavorable electrostatic energy contribution of each residue in pH 10.0 is similar to what has been discussed for pH 7.5. There is an increase of ΔGqq for acidic residues. The positive energy contribution passes from 12.2 kJ⋅mol−1 in pH 7.5 to 14.5 kJ⋅mol−1 in pH 10.0. The residue E19 had a significant variation from negative value ΔGqq = −1.07 kJ⋅mol−1 to a slightly positive contribution ΔGqq = 0.15 kJ⋅mol−1. The N-terminal also had a significant variation due to the fact that it becomes deprotonated with charge 0 in pH 10.0 and its energy contribution is close to 0. The total electrostatic free energy ΔGelec for pH 10.0 is −31.12 kJ⋅mol−1.
Figure 3.
Charge–charge interaction energy ΔGqq calculated by the TKSA model for each ionizable residue. (A–D) Given here are the energy profiles of pH values 2.5, 4.5, 7.5 and 10.0, respectively. The red bars indicate the residues with the side chain exposed to solvent with SASA ≥ 50% and positive energy contribution to native state stability, most of these residues are located between D36 and E42. To see this figure in color, go online.
pH and ionic strength effects on 1PGB-QDD folding: CpHMD-SBM-Cα
The previous section indicates that the drive interaction responsible for PGB1-QDD stability is electrostatic. However, the TKSA model takes into account only the static native structure of the protein in its calculations, whereas experimental data suggests that the unfolded state of PGB1-QDD is also important to its stability (37). Thus, in this section, CpHMD-SBM simulations were utilized, which can explore the effects of the unfolded and transition states on protein stability. 1PGB-QDD was therefore simulated at four different pH values (2.5, 4.5, 7.5, and 10.0) and three ionic strengths (low salt ≈ 0.01 M, intermediate salt ≈ 0.1 M, and high salt ≈ 0.4 M).
Heat capacity (Cv) and free energy profiles of PGB1-QDD were calculated to explore the role of pH and ionic strength on protein folding thermodynamics (Fig. 4). In Fig. 4 A, Cv is presented as a function of temperature in Kelvin and in reduced units T∗. T∗ is calculated by T∗ = kBT/εC, where εC is a simulation parameter, which can be chosen arbitrarily because all other parameters of the SBM model are written in function of εC. Thus, the direct comparison between absolute temperatures obtained by experiments and simulations must be done carefully. Therefore, in this study, the comparison between experimental and computational heat capacity is made observing the relative variation of the temperatures and the qualitative behavior of the curves. Heat capacity curves obtained by Lindman et al. (37) using differential scanning calorimetry present low ΔCp values in a large range of solution conditions, which indicates that 1PGB-QDD is stable in a wide range of temperatures. However, heat capacity calculated computationally does not present a shift between the baseline for the protein in its folded and unfolded states. This difference between simulation and experiment is caused by a simplification in the force field of the CpHMD-SBM-Cα model, which does not take into account the temperature dependence in its interactions (see Eq. 5).
Figure 4.
Thermodynamic properties of 1PGB-QDD folding. (A) Shown here are heat capacities at constant volume (Cv) in low, intermediate, and high salt concentrations. Solid black lines are for pH 2.5, short dashed red lines are for pH 4.5, long dashed blue lines are for pH 7.5, and dashed/dotted green lines are for pH 10.0. (B) Given here are free energy curves as a function of the reaction coordinate Q for the same three salt concentrations. All systems are at T close to TM (melting temperature) of pH 7.5 in each ionic strength. It is considered that the melting temperature corresponds to the peak of Cv. To see this figure in color, go online.
In these CpHMD-SBM-Cα computational results, electrostatic screening at the high salt concentration is too strong, making the pH effect on protein stability almost insignificant. Thus, it is not possible to see a significant variation in 1PGB-QDD thermodynamic properties at this ionic strength. As expected, the pH effect becomes more evident in the protein at low and intermediate salt concentrations, but the higher stability remains at pH 4.5. At the low salt concentration, the difference between the Cv peaks of each pH curve was maximized. The transition barrier between the folded and unfolded state in protein near TM was not affected by salting, with the barrier close to 2.0 εC. In general, the thermodynamic properties calculated by CpHMD-SBM are in qualitative agreement with the experimental results.
The comparison between experimental and computational TM is presented in Fig. 5. The general behavior TM for computational results in each salt concentration shows a good level of agreement with the experimental data, with the TM peak at pH 4.5 for all systems. At low pH, there is a decrease of TM mostly in low and intermediate salt concentrations. In this case, both experiments and simulations agree that a high ionic strength contributes to 1PGB-QDD stability. However, when the protein has a pH higher than 4.5, the highest stability in the simulations occurs at low salt concentration, followed by intermediate and high salt concentrations. On the other hand, experimental results at pH 7.5 and 10.0 indicate that the electrostatic screening caused by a high salt concentration helps to stabilize 1PGB-QDD. This difference between experimental and computational results is caused by the electrostatic energy calculated in the simulation, and this interaction will be explored below.
Figure 5.
Values of melting temperature TM∗ in reduced temperature as a function of pH. Black circles are for low salt, red squares are for an intermediate ionic strength, and blue diamonds are for high salt concentration. Inset graphic presents the experimental results of TM∗ as a function of pH in similar salt conditions to simulation (adapted from (37)). Dashed lines connecting symbols help guide the eye. To see this figure in color, go online.
The effect of electrostatic interaction and charge regulation on each folding stage of 1PGB-QDD can be evaluated by an analysis of Fig. 6. This figure presents a 2D map of the distribution of electrostatic energy as a function of Q. The net charge q in each highly populated region of distribution of electrostatic energy was also calculated and included in Fig. 6.
Figure 6.
2D map of the distribution of electrostatic energy as a function of the reaction coordinate Q for pH 2.5, 4.5, 7.5, and 10.0. The color map represents the probability distribution of the electrostatic energy and Q, normalized by its highest value. All the systems are at T ≈ TM and intermediate salt concentration. The value q represents the net charge of 1PGB-QDD in each minimum. To see this figure in color, go online.
Previous results showed that the highest stability of 1PGB-QDD occurs at pH 4.5. This is because the protein in this condition has the lowest mean electrostatic energy 〈Eelec〉 compared to the protein at pH 2.5, 7.5, and 10.0. For both folded and unfolded states, the 〈Eelec〉 is near −4 εC and net charge q = −1, not favoring any state in comparison to each other. A similar discussion applies to the protein at pH 7.5: the folded and unfolded states have 〈Eelec〉 close to −2 εC and q = −5 (experimental net charge (37) is −6). An increase in mean electrostatic energy promotes a loss of stability and is caused by this negative q, although, even with a high negative net charge, 〈Eelec〉 is kept negative. This favorable energy explains the fact that low salt concentration promotes more stability in 1PGB-QDD at pH 7.5 than intermediate and high salt concentrations (Fig. 5), which does not agree with the experimental results. Although the net charge is similar (experimental q = −6 and computational q = −5), the arrangement of charges in the Cα model can be different from the more realistic case, and creates more regions of favorable interaction. At pH 10.0, the same thing occurs: the protein has a high negative q = −7 or q = −6 and a negative value of 〈Eelec〉 close to −1 εC.
Fig. 6 also shows that the pH effect at pH 4.5 and 7.5 is strong enough to keep the protein net charge, irrespective of its conformational state. However, at the extreme pH values 2.5 and 10.0, some differences in electrostatic energy and net charge arise, depending on the folding stages. In these cases, the charge regulation is more evident, and the interaction between charged groups becomes fundamental for their protonation states. With the protein at pH 2.5, the 2D map presents two populated regions in the unfolded state, one with 〈Eelec〉 close to 1 εC (q = 5) and other near 2.0 εC (q = 6), the second being the most populated. In the folded state, in this same pH condition, three populated regions arise—the most populated with the highest 〈Eelec〉 = 4 εC and q = 7; an intermediate region in 〈Eelec〉 = 2 εC and q = 6; and the lowest populated region with 〈Eelec〉 = 1 εC and q = 5. The loss of stability at this low pH condition may be associated with the fact that 〈Eelec〉 of the unfolded protein is lower than 〈Eelec〉 in folded state. Thus, the electrostatic interaction favors the unfolded state.
Comparison between computational and experimental pKa
In this final section, the experimental pKa and the pKa calculated by TKSA (for folded state), and CpHMD (for folded and unfolded states), were compared to evaluate the sampling of charges using these models for different pH conditions. The computational pKa and the Hill coefficient n were calculated fitting individual titration curves with the Hill equation (60) (examples of these curves are presented in Fig. 7). The Hill coefficient values are associated with the correlation between titratable groups. For the cases where n = 1, the electrostatic energy is constant throughout the titration; for n > 1, the attraction between the charged groups tends to change linearly with pH; and when the Hill coefficient is <1, the repulsion changes linearly with pH (74). For 1PGB-QDD, the experimental and computational n values tend to be near or <1. The highest discrepancy between these values occurs for the results of TKSA model, which presents low Hill coefficients. This discrepancy is caused by the residues that are electrostatically coupled, because this model does not take into account the MD and the protein is held in its compact crystal structure.
Figure 7.
Titration curves of residue D22: ionization degree α as a function of pH. (A) The value α is calculated for the folded 1PGB-QDD using the CpHMD model. (B) The value α is calculated for the unfolded 1PGB-QDD via the CpHMD model. (C) The value α is calculated for the folded protein using the TKSA model. Red curves are the fit of the Hill equation (Eq. 7). To see this figure in color, go online.
The results are presented in Table 1. The errors in pKa values were calculated using a set of three different simulations, and an error <0.1 pKa unit was obtained for each ionizable residue. Thus, these error values were omitted from Table 1.
Table 1.
Comparison between Experimental and Computational pKa Values for the Acidic Residues in Folded and Unfolded 1PGB-QDD Protein
| Residue | Experimentala |
CpHMDb |
TKSAc |
|||||||
|---|---|---|---|---|---|---|---|---|---|---|
| pKa Folded | n Folded | pKa Unfolded | n Unfolded | pKa Folded | n Folded | pKa Unfolded | n Unfolded | pKa Folded | n Folded | |
| D08 | 4.9 | 0.7 | 3.7 | 0.9 | 3.5 | 0.8 | 3.5 | 0.9 | 4.2 | 0.5 |
| E15 | 4.6 | 0.9 | 4.4 | 0.9 | 4.0 | 0.8 | 4.1 | 0.9 | 1.8 | 0.8 |
| E19 | 3.9 | 1.0 | 4.5 | 0.8 | 4.0 | 0.8 | 4.4 | 0.9 | 4.0 | 0.9 |
| D22 | 3.0 | 1.0 | 3.9 | 0.9 | 3.7 | 0.8 | 3.9 | 0.9 | 3.4 | 0.8 |
| E27 | 4.8 | 0.8 | 4.4 | 0.9 | 3.5 | 0.8 | 3.7 | 0.9 | 0.6 | 0.6 |
| D36 | 4.2 | 0.8 | 4.1 | 0.8 | 4.1 | 0.7 | 4.2 | 0.7 | 4.1 | 0.8 |
| D37 | 6.5 | 1.1 | 4.2 | 0.8 | 4.1 | 0.7 | 4.3 | 0.7 | 4.0 | 0.7 |
| D40 | 4.1 | 0.7 | 4.2 | 0.7 | 4.0 | 0.7 | 4.1 | 0.8 | 4.3 | 0.7 |
| E42 | 4.8 | 0.7 | 4.4 | 0.9 | 4.6 | 0.8 | 4.7 | 0.8 | 5.0 | 0.7 |
| D46 | 3.8 | 0.9 | 4.0 | 0.9 | 4.0 | 0.7 | 4.1 | 0.8 | 4.0 | 0.7 |
| D47 | 3.1 | 0.9 | 3.9 | 0.9 | 3.6 | 0.7 | 3.8 | 0.8 | 2.9 | 0.8 |
| E56 | 3.8 | 0.6 | 4.7 | 0.8 | 4.6 | 0.7 | 4.4 | 0.9 | 6.5 | 0.4 |
| C-ter | 3.1 | 0.7 | 3.6 | 0.7 | — | — | — | — | 3.5 | 0.6 |
Error values for all cases are <0.1 units and have been omitted. Experimental values are from (74).
The experimental pKa and n values for the acidic residues measured by Lindman et al. (74).
The computational pKa and n values fitting the individual titration curves with the Hill equation; simulation using CpHMD.
The computational pKa and n values fitting the individual titration curves with the Hill equation; simulation using TKSA.
Only differences higher than 1.0 pKa unit between experimental and computational values were considered significant (19). Comparing the experimental pKa of folded 1PGB-QDD protein and the values calculated using TKSA, only four residues presented a difference higher than 1.0 pKa unit. This is a good level of agreement considering the simplicity of the model, which can output the results in <1 h. E15 and E27 have a downshift in computational pKa that may be caused by the overrated interaction calculated in this model that these residues make with K04 and K21, respectively. D37 has a large upshift in its experimental pKa, which may be caused by the desolvation effects associated with side-chain burial (74) that TKSA is not taking into account sufficiently. Finally, the residue E56 has two charged groups that cause a large electrostatic repulsion in the TKSA model; this repulsion may explain the high upshift in its computational pKa value.
Using CpHMD, the comparison with the experimental results is even better, with only three residues of folded 1PGB-QDD presenting >1.0 pKa unit of difference between the simulation and the experimental pKa values. E27 has a similar downshift in pKa values to that obtained using TKSA; the reason for this difference is the high interaction that this residue makes with K04. In the case of D08, the downshift of 1.4 pKa units obtained in computational results could be caused by a model limitation. Considering the native structure of the protein, the distance between the Cα of D08 and Cα K13 is near 5.1 Å, whereas the distance between the carboxyl group of D08 and ammonium group of K13 is close to 8.3 Å. This difference in the distances promotes a more favorable interaction between these residues in CpHMD, resulting in this disagreement between experimental and simulation results. In residue D37, as previously stated, desolvation effects have a strong impact on its experimental pKa value and the CpHMD model does not treat this effect adequately. When the pKa of the unfolded protein was considered, the agreement between experimental values and CpHMD values is remarkable: in no case does the difference between these values exceed 1.0 pKa unit.
Conclusions
In this study, the protein 1PGB-QDD was studied by two coarse-grained models with the objective of evaluating the pH and salting effect on protein stability. The results of this study were compared with experimental results. The first model tested here was TKSA, which analyzes the effect of each ionizable residue in the electrostatic free energy of the protein in its native state. Using this simple model, it was possible to quantify the importance of each charged residue in 1PGB-QDD stability. The comparison between TKSA calculations and experimental results also indicates that the driving force of this stability is the electrostatic interaction. The second model used in this study was CpHMD-SBM, which was able to explore the protein folding process in constant pH (19). With CpHMD-SBM simulations, the folded and unfolded states and the transition between these states could be studied. The pH effect in 1PGB-QDD stability explored by simulation was found to be in good agreement with the experimental data. Similarly, the salt effect was evaluated, and the simulations obtained the same general behavior in protein stability as the experiments. Finally, pKa was calculated using TKSA in the folded protein and CpHMD-SBM for folded and unfolded 1PGB-QDD. The pKa values calculated by the two models are near experimental values. Therefore, most of the effects associated with pH and ionic strength variations in 1PGB-QDD stability were qualitatively described by both coarse-grained models.
Author Contributions
V.M.d.O., V.d.G.C., and V.B.P.L. designed research. V.M.d.O. and V.d.G.C. performed research. Data were analyzed by V.M.d.O., V.d.G.C., F.B.d.S., D.L.Z.C., and S.J.d.C. All authors contributed to the manuscript writing.
Acknowledgments
We thank the Center for Scientific Computing (NCC/GridUNESP) of São Paulo State University (UNESP) for computational resources.
V.M.d.O. was supported by the National Council for Scientific and Technological Development (CNPq) under Grant Process No. 141985/2013-5. V.d.G.C. was funded by grants 2016/13998-8 and 2017/09662-7, São Paulo Research Foundation (FAPESP) and Higher Education Personnel Improvement Coordination (CAPES). F.B.d.S. was supported by the Higher Education Personnel Improvement Coordination (CAPES). D.L.Z.C. acknowledges the São Paulo Research Foundation (FAPESP), grant No. 2013/13151-7, for financial support. V.B.P.L. was supported by the National Council for Scientific and Technological Development (CNPq) and the São Paulo Research Foundation (FAPESP) under grant Nos. 2014/06862-7 and 2016/19766-1.
Editor: Nathan Baker.
Footnotes
Vinicius Martins de Oliveira and Vinicius de Godoi Contessoto contributed equally to this work.
References
- 1.Roca M., Liu H., Warshel A. On the relationship between thermal stability and catalytic power of enzymes. Biochemistry. 2007;46:15076–15088. doi: 10.1021/bi701732a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Bye J.W., Platts L., Falconer R.J. Biopharmaceutical liquid formulation: a review of the science of protein stability and solubility in aqueous environments. Biotechnol. Lett. 2014;36:869–875. doi: 10.1007/s10529-013-1445-6. [DOI] [PubMed] [Google Scholar]
- 3.Huber K.V., Olek K.M., Superti-Furga G. Proteome-wide drug and metabolite interaction mapping by thermal-stability profiling. Nat. Methods. 2015;12:1055–1057. doi: 10.1038/nmeth.3590. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Yamasaki M., Yano H., Aoki K. Differential scanning calorimetric studies on bovine serum albumin: I. Effects of pH and ionic strength. Int. J. Biol. Macromol. 1990;12:263–268. doi: 10.1016/0141-8130(90)90007-w. [DOI] [PubMed] [Google Scholar]
- 5.Kim K., Kim S., Kwon D. Changes of glycinin conformation due to pH, heat and salt determined by differential scanning calorimetry and circular dichroism. Int. J. Food Sci. Technol. 2004;39:385–393. [Google Scholar]
- 6.Bye J.W., Falconer R.J. Thermal stability of lysozyme as a function of ion concentration: a reappraisal of the relationship between the Hofmeister series and protein stability. Protein Sci. 2013;22:1563–1570. doi: 10.1002/pro.2355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Tripathi S., Garcìa A.E., Makhatadze G.I. Alterations of nonconserved residues affect protein stability and folding dynamics through charge-charge interactions. J. Phys. Chem. B. 2015;119:13103–13112. doi: 10.1021/acs.jpcb.5b08527. [DOI] [PubMed] [Google Scholar]
- 8.Wunderlich M., Martin A., Schmid F.X. Stabilization of the cold shock protein CspB from Bacillus subtilis by evolutionary optimization of Coulombic interactions. J. Mol. Biol. 2005;347:1063–1076. doi: 10.1016/j.jmb.2005.02.014. [DOI] [PubMed] [Google Scholar]
- 9.Papaleo E., Olufsen M., Brandsdal B.O. Optimization of electrostatics as a strategy for cold-adaptation: a case study of cold- and warm-active elastases. J. Mol. Graph. Model. 2007;26:93–103. doi: 10.1016/j.jmgm.2006.09.012. [DOI] [PubMed] [Google Scholar]
- 10.Date M.S., Dominy B.N. Modeling the influence of salt on the hydrophobic effect and protein fold stability. Commun. Comput. Phys. 2013;13:90–106. [Google Scholar]
- 11.Luitz M.P., Bomblies R., Zacharias M. Adenylylation of Tyr77 stabilizes Rab1b GTPase in an active state: a molecular dynamics simulation analysis. Sci. Rep. 2016;6:19896. doi: 10.1038/srep19896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Su J.-G., Han X.-M., Wang J.-H. Impacts of the charged residues mutation S48E/N62H on the thermostability and unfolding behavior of cold shock protein: insights from molecular dynamics simulation with Gō model. J. Mol. Model. 2016;22:91. doi: 10.1007/s00894-016-2958-4. [DOI] [PubMed] [Google Scholar]
- 13.Coronado M.A., Caruso I.P., Eberle R.J. Cold shock protein A from Corynebacterium pseudotuberculosis: role of electrostatic forces in the stability of the secondary structure. Protein Pept. Lett. 2017;24:358–367. doi: 10.2174/0929866524666170207153808. [DOI] [PubMed] [Google Scholar]
- 14.Azia A., Levy Y. Nonnative electrostatic interactions can modulate protein folding: molecular dynamics with a grain of salt. J. Mol. Biol. 2009;393:527–542. doi: 10.1016/j.jmb.2009.08.010. [DOI] [PubMed] [Google Scholar]
- 15.Givaty O., Levy Y. Protein sliding along DNA: dynamics and structural characterization. J. Mol. Biol. 2009;385:1087–1097. doi: 10.1016/j.jmb.2008.11.016. [DOI] [PubMed] [Google Scholar]
- 16.Zarrine-Afsar A., Zhang Z., Chan H.S. Kinetic consequences of native state optimization of surface-exposed electrostatic interactions in the Fyn SH3 domain. Proteins. 2012;80:858–870. doi: 10.1002/prot.23243. [DOI] [PubMed] [Google Scholar]
- 17.Chen T., Song J., Chan H.S. Theoretical perspectives on nonnative interactions and intrinsic disorder in protein folding and binding. Curr. Opin. Struct. Biol. 2015;30:32–42. doi: 10.1016/j.sbi.2014.12.002. [DOI] [PubMed] [Google Scholar]
- 18.Mouro P.R., de Godoi Contessoto V., Pereira Leite V.B. Quantifying nonnative interactions in the protein-folding free-energy landscape. Biophys. J. 2016;111:287–293. doi: 10.1016/j.bpj.2016.05.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Contessoto V.G., de Oliveira V.M., Leite V.B.P. NTL9 folding at constant pH: the importance of electrostatic interaction and pH dependence. J. Chem. Theory Comput. 2016;12:3270–3277. doi: 10.1021/acs.jctc.6b00399. [DOI] [PubMed] [Google Scholar]
- 20.Khandogin J., Brooks C.L., 3rd Constant pH molecular dynamics with proton tautomerism. Biophys. J. 2005;89:141–157. doi: 10.1529/biophysj.105.061341. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Mongan J., Case D.A. Biomolecular simulations at constant pH. Curr. Opin. Struct. Biol. 2005;15:157–163. doi: 10.1016/j.sbi.2005.02.002. [DOI] [PubMed] [Google Scholar]
- 22.Tanford C., Kirkwood J.G. Theory of protein titration curves. I. General equations for impenetrable spheres. J. Am. Chem. Soc. 1957;79:5333–5339. [Google Scholar]
- 23.Ibarra-Molero B., Loladze V.V., Sanchez-Ruiz J.M. Thermal versus guanidine-induced unfolding of ubiquitin. An analysis in terms of the contributions from charge-charge interactions to protein stability. Biochemistry. 1999;38:8138–8149. doi: 10.1021/bi9905819. [DOI] [PubMed] [Google Scholar]
- 24.Makhatadze G.I., Loladze V.V., Thomas S.T. Contribution of surface salt bridges to protein stability: guidelines for protein engineering. J. Mol. Biol. 2003;327:1135–1148. doi: 10.1016/s0022-2836(03)00233-x. [DOI] [PubMed] [Google Scholar]
- 25.Permyakov S.E., Makhatadze G.I., Berliner L.J. How to improve nature: study of the electrostatic properties of the surface of α-lactalbumin. Protein Eng. Des. Sel. 2005;18:425–433. doi: 10.1093/protein/gzi051. [DOI] [PubMed] [Google Scholar]
- 26.Strickler S.S., Gribenko A.V., Makhatadze G.I. Protein stability and surface electrostatics: a charged relationship. Biochemistry. 2006;45:2761–2766. doi: 10.1021/bi0600143. [DOI] [PubMed] [Google Scholar]
- 27.Gribenko A.V., Patel M.M., Makhatadze G.I. Rational stabilization of enzymes by computational redesign of surface charge-charge interactions. Proc. Natl. Acad. Sci. USA. 2009;106:2601–2606. doi: 10.1073/pnas.0808220106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Björck L., Kastern W., Widebäck K. Streptococcal protein G, expressed by streptococci or by Escherichia coli, has separate binding sites for human albumin and IgG. Mol. Immunol. 1987;24:1113–1122. doi: 10.1016/0161-5890(87)90080-0. [DOI] [PubMed] [Google Scholar]
- 29.Sjöbring U., Björck L., Kastern W. Protein G genes: structure and distribution of IgG-binding and albumin-binding domains. Mol. Microbiol. 1989;3:319–327. doi: 10.1111/j.1365-2958.1989.tb00177.x. [DOI] [PubMed] [Google Scholar]
- 30.Björck L., Kronvall G. Purification and some properties of streptococcal protein G, a novel IgG-binding reagent. J. Immunol. 1984;133:969–974. [PubMed] [Google Scholar]
- 31.Reis K.J., Ayoub E.M., Boyle M.D. Streptococcal Fc receptors. I. Isolation and partial characterization of the receptor from a group C streptococcus. J. Immunol. 1984;132:3091–3097. [PubMed] [Google Scholar]
- 32.Kmiecik S., Kolinski A. Folding pathway of the b1 domain of protein G explored by multiscale modeling. Biophys. J. 2008;94:726–736. doi: 10.1529/biophysj.107.116095. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Alexander P., Fahnestock S., Bryan P. Thermodynamic analysis of the folding of the streptococcal protein G IgG-binding domains B1 and B2: why small proteins tend to have high denaturation temperatures. Biochemistry. 1992;31:3597–3603. doi: 10.1021/bi00129a007. [DOI] [PubMed] [Google Scholar]
- 34.Alexander P., Orban J., Bryan P. Kinetic analysis of folding and unfolding the 56 amino acid IgG-binding domain of streptococcal protein G. Biochemistry. 1992;31:7243–7248. doi: 10.1021/bi00147a006. [DOI] [PubMed] [Google Scholar]
- 35.Sikosek T., Krobath H., Chan H.S. Theoretical insights into the biophysics of protein Bi-stability and evolutionary switches. PLOS Comput. Biol. 2016;12:e1004960. doi: 10.1371/journal.pcbi.1004960. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Gronenborn A.M., Filpula D.R., Clore G.M. A novel, highly stable fold of the immunoglobulin binding domain of streptococcal protein G. Science. 1991;253:657–661. doi: 10.1126/science.1871600. [DOI] [PubMed] [Google Scholar]
- 37.Lindman S., Xue W.-F., Linse S. Salting the charged surface: pH and salt dependence of protein G B1 stability. Biophys. J. 2006;90:2911–2921. doi: 10.1529/biophysj.105.071050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Wallace J.A., Shen J.K. Continuous constant pH molecular dynamics in explicit solvent with pH-based replica exchange. J. Chem. Theory Comput. 2011;7:2617–2629. doi: 10.1021/ct200146j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Huang Y., Chen W., Shen J. All-atom continuous constant pH molecular dynamics with particle mesh Ewald and titratable water. J. Chem. Theory Comput. 2016;12:5411–5421. doi: 10.1021/acs.jctc.6b00552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Shen J.K. Uncovering specific electrostatic interactions in the denatured states of proteins. Biophys. J. 2010;99:924–932. doi: 10.1016/j.bpj.2010.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Shire S.J., Hanania G.I., Gurd F.R. Electrostatic effects in myoglobin. Hydrogen ion equilibria in sperm whale ferrimyoglobin. Biochemistry. 1974;13:2967–2974. doi: 10.1021/bi00711a028. [DOI] [PubMed] [Google Scholar]
- 42.Tanford C., Roxby R. Interpretation of protein titration curves. Application to lysozyme. Biochemistry. 1972;11:2192–2198. doi: 10.1021/bi00761a029. [DOI] [PubMed] [Google Scholar]
- 43.Orttung W.H. Proton binding and dipole moment of hemoglobin. Refined calculations. Biochemistry. 1970;9:2394–2402. doi: 10.1021/bi00814a002. [DOI] [PubMed] [Google Scholar]
- 44.Kirkwood J.G. Theory of solutions of molecules containing widely separated charges with special application to zwitterions. J. Chem. Phys. 1934;2:351–361. [Google Scholar]
- 45.Da Silva F.L.B., Jönsson B., Penfold R. A critical investigation of the Tanford-Kirkwood scheme by means of Monte Carlo simulations. Protein Sci. 2001;10:1415–1425. doi: 10.1110/ps.42601. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Havranek J.J., Harbury P.B. Tanford-Kirkwood electrostatics for protein modeling. Proc. Natl. Acad. Sci. USA. 1999;96:11145–11150. doi: 10.1073/pnas.96.20.11145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Bashford D., Karplus M. Multiple-site titration curves of proteins: an analysis of exact and approximate methods for their calculation. J. Phys. Chem. 1991;95:9556–9561. [Google Scholar]
- 48.Clementi C., Nymeyer H., Onuchic J.N. Topological and energetic factors: what determines the structural details of the transition state ensemble and “en-route” intermediates for protein folding? An investigation for small globular proteins. J. Mol. Biol. 2000;298:937–953. doi: 10.1006/jmbi.2000.3693. [DOI] [PubMed] [Google Scholar]
- 49.Noel J.K., Whitford P.C., Onuchic J.N. SMOG@ctbp: simplified deployment of structure-based models in GROMACS. Nucleic Acids Res. 2010;38:W657–W661. doi: 10.1093/nar/gkq498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Contessoto V.G., Lima D.T., Leite V.B. Analyzing the effect of homogeneous frustration in protein folding. Proteins. 2013;81:1727–1737. doi: 10.1002/prot.24309. [DOI] [PubMed] [Google Scholar]
- 51.de Mendonça M.R., Rizzi L.G., Alves N.A. Inferring a weighted elastic network from partial unfolding with coarse-grained simulations. Proteins. 2014;82:119–129. doi: 10.1002/prot.24381. [DOI] [PubMed] [Google Scholar]
- 52.Tzul F.O., Schweiker K.L., Makhatadze G.I. Modulation of folding energy landscape by charge-charge interactions: linking experiments with computational modeling. Proc. Natl. Acad. Sci. USA. 2015;112:E259–E266. doi: 10.1073/pnas.1410424112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Ullner M., Woodward C.E., Jnsson B. A Debye-Hückel theory for electrostatic interactions in proteins. J. Chem. Phys. 1996;105:2056–2065. [Google Scholar]
- 54.Baptista A.M., Teixeira V.H., Soares C.M. Constant-pH molecular dynamics using stochastic titration. J. Chem. Phys. 2002;117:4184–4200. [Google Scholar]
- 55.Mongan J., Case D.A., McCammon J.A. Constant pH molecular dynamics in generalized Born implicit solvent. J. Comput. Chem. 2004;25:2038–2048. doi: 10.1002/jcc.20139. [DOI] [PubMed] [Google Scholar]
- 56.Williams S.L., Blachly P.G., McCammon J.A. Measuring the successes and deficiencies of constant pH molecular dynamics: a blind prediction study. Proteins. 2011;79:3381–3388. doi: 10.1002/prot.23136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Pathak A.K. Effect of a buried ion pair in the hydrophobic core of a protein: an insight from constant pH molecular dynamics study. Biopolymers. 2015;103:148–157. doi: 10.1002/bip.22577. [DOI] [PubMed] [Google Scholar]
- 58.Pathak A.K. Constant pH molecular dynamics study on the doubly mutated staphylococcal nuclease: capturing the microenvironment. RSC Advances. 2015;5:94926–94932. [Google Scholar]
- 59.Metropolis N., Rosenbluth A.W., Teller E. Equation of state calculations by fast computing machines. J. Chem. Phys. 1953;21:1087–1092. [Google Scholar]
- 60.Swails J.M., York D.M., Roitberg A.E. Constant pH replica exchange molecular dynamics in explicit solvent using discrete protonation states: implementation, testing, and validation. J. Chem. Theory Comput. 2014;10:1341–1352. doi: 10.1021/ct401042b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Arnold A., Lenz O., Holm C. ESPResSo 3.1: molecular dynamics software for coarse-grained models. In: Griebel M., Schweitzer M.A., editors. Number 89 in Lecture Notes in Computational Science and Engineering. Springer; Heidelberg, Germany: 2013. pp. 1–23. (Meshfree Methods for Partial Differential Equations VI). [Google Scholar]
- 62.Limbach H.J., Arnold A., Holm C. ESPResSo: an extensible simulation package for research on soft matter systems. Comput. Phys. Commun. 2006;174:704–727. [Google Scholar]
- 63.Ferrenberg A.M., Swendsen R.H. New Monte Carlo technique for studying phase transitions. Phys. Rev. Lett. 1988;61:2635–2638. doi: 10.1103/PhysRevLett.61.2635. [DOI] [PubMed] [Google Scholar]
- 64.Kumar S., Rosenberg J.M., Kollman P.A. The weighted histogram analysis method for free energy calculations on biomolecules. I. The method. J. Comput. Chem. 1992;13:1011–1021. [Google Scholar]
- 65.Kumar S., Rosenberg J.M., Kollman P.A. Multidimensional free energy calculations using the weighted histogram analysis method. J. Comput. Chem. 1995;16:1339–1350. [Google Scholar]
- 66.Sobolev V., Sorokine A., Edelman M. Automated analysis of interatomic contacts in proteins. Bioinformatics. 1999;15:327–332. doi: 10.1093/bioinformatics/15.4.327. [DOI] [PubMed] [Google Scholar]
- 67.Webb B., Sali A. Current Protocols in Bioinformatics. John Wiley; Hoboken, NJ: 2002. Comparative protein structure modeling using MODELLER. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Pettersen E.F., Goddard T.D., Ferrin T.E. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 2004;25:1605–1612. doi: 10.1002/jcc.20084. [DOI] [PubMed] [Google Scholar]
- 69.Bond C.S., Schüttelkopf A.W. ALINE: a WYSIWYG protein-sequence alignment editor for publication-quality alignments. Acta Crystallogr. D Biol. Crystallogr. 2009;65:510–512. doi: 10.1107/S0907444909007835. [DOI] [PubMed] [Google Scholar]
- 70.Tsodikov O.V., Record M.T., Jr., Sergeev Y.V. Novel computer program for fast exact calculation of accessible and molecular surface areas and average surface curvature. J. Comput. Chem. 2002;23:600–609. doi: 10.1002/jcc.10061. [DOI] [PubMed] [Google Scholar]
- 71.Ibarra-Molero B., Sanchez-Ruiz J.M. Genetic algorithm to design stabilizing surface-charge distributions in proteins. J. Phys. Chem. B. 2002;106:6609–6613. [Google Scholar]
- 72.Loladze V.V., Makhatadze G.I. Energetics of charge-charge interactions between residues adjacent in sequence. Proteins. 2011;79:3494–3499. doi: 10.1002/prot.23132. [DOI] [PubMed] [Google Scholar]
- 73.Makhatadze G.I., Loladze V.V., Lopez M.M. Mechanism of thermostabilization in a designed cold shock protein with optimized surface electrostatic interactions. J. Mol. Biol. 2004;336:929–942. doi: 10.1016/j.jmb.2003.12.058. [DOI] [PubMed] [Google Scholar]
- 74.Lindman S., Bauer M.C., Linse S. pKa values for the unfolded state under native conditions explain the pH-dependent stability of PGB1. Biophys. J. 2010;99:3365–3373. doi: 10.1016/j.bpj.2010.08.078. [DOI] [PMC free article] [PubMed] [Google Scholar]





