Abstract
The DelPhiPKa is a widely used and unique approach to compute pKa’s of ionizable groups that does not require molecular surface to be defined. Instead, it uses smooth Gaussian-based dielectric function to treat computational space via Poisson-Boltzmann equation (PBE). Here we report an expansion of DelPhiPKa functionality to enable inclusion of salt in the modeling protocol. The method considers the salt mobile ions in solvent phase without defining solute-solvent boundary. Instead, the ions are penalized to enter solute interior via a desolvation penalty term in the Boltzmann factor in the framework of PBE. Hence, the concentration of ions near to protein is balanced by the desolvation penalty and electrostatic interactions. The study reveals that correlation between experimental and calculated pKa’s is improved significantly by taking into consideration the presence of salt. Furthermore, it is demonstrated that DelphiPKa reproduces the salt sensitivity of experimentally measured pKa’s. Another new development of DelPhiPKa allows for computing the pKa’s of polar residues such as cysteine, serine, threonine and tyrosine. With this regard, DelPhiPKa is benchmarked against experimentally measured cysteine and tyrosine pKa’s and for cysteine it is shown to outperform other existing methods (DelPhiPKa RMSD of 1.73 versus RMSD between 2.40 and 4.72 obtained by other existing pKa prediction methods).
Keywords: pKa’s, electrostatics, salt concentration, Poisson-Boltzmann equation, proteins, Gaussian-based dielectric function
Introduction
The acidic and basic amino acid residues of a protein may be ionized at a particular pH. The ionization state of these residues contribute to conformation, stability, solubility and function of the protein as a function of pH.1,2,3 Hence, ionizable residues of biological molecules play a significant role in protein-protein interaction, protein-ligand binding, enzymatic reaction etc.4,5 As pKa determines the ionization state of the residues, it is important to know the pKa values of titrable groups in proteins and how these values depend on the surrounding environment. However, accurate prediction of pKa’s of ionizable groups remain a challenge since it requires to determine the equilibrium distribution of ionization states of side chain titrable residues along with the conformational changes.
Several computational techniques have been explored for pKa’s calculations.6,7,8,9,10 These methods can be broadly classified into three classes: i) microscopic ii) empirical and iii) macroscopic methods. All these techniques have their weakness and strength in capturing the biophysical insights which in turn establish their ability to match the experimentally determined pKa values. All these three classes of approaches are described in detail by Alexov et al.6 Microscopic methods such as molecular dynamics (MD), quantum mechanics molecular mechanics (QM/MM) consider the atomic level of detail of the system while calculating the thermodynamics properties. Generally, these methods are more accurate as they incorporate the conformational change of protein while a residue changes from protonated to unprotonated form. However, convergence problem in performing the configurational sampling, large computational cost makes these methods unrealistic for most of the protein applications. On the other hand, empirical methods use knowledge-based terms and scoring functions to describe the effect of the environment on pKa’s of protein residues.11,12,13,14 Since they do not consider conformational sampling explicitly, they are very fast and in some cases quite accurate. Indeed, PROPKA has recently been found to be a reliable protein pKa predictor,14,15,16,17 with reported RMSD is as low as 0.89. Another classes of methods are macroscopic methods based on continuum electrostatics (CE) using either Poisson Boltzmann (PB) or General Born (GB) equations. Traditional implementation of these methods requires that the computational space be divided into solute and solvent regions and no conformational changes to be considered. Several groups tried to improve the CE approaches by coupling ionization and conformation using Monte Carlo (MC) sampling technique, introducing side chain flexibility,18,19,20 modifying hydrogen bond orientation,21,22 and adding extensive side chain rotamer sampling.23,22 All these additional implementations made the PB based CE method more accurate. Recently we reported a Gaussian-based method, where a smooth Gaussian-based dielectric function describes the dielectric properties of the solute and solvent on the same footage. This pKa’s prediction method, which has been already implemented in DelPhi,24 uses Gaussian-based smooth dielectric function which in turn mimic the plausible conformational changes associated with changes of ionization states. In this approach, there is no need to define molecular surface while calculating the electrostatic energies and the corresponding pKa’s.
The electrostatic interactions within solute depend on the properties of the solvent. Typically charge residues are located on the surface of the solute, these are accessible to salts (NaCl, KCl, K2SO4, etc.) which are present in the solvent. Thus, the presence of salt affects the screening of electrostatic interactions. Furthermore, ions can bind to oppositely charged protein residues25,26 or also can cause chaotropic effect in solvent structure. Therefore, protein stability can be changed in the presence of salt. It mostly depends on the charge distribution on protein surface and in turn on pH as indicated by the following experimental observations. It has been found that the stability of chymotrypsinogen increases with salt concentration below pH 1.5 but it decreases at higher pH. Surprisingly, no effect on stability of barnase has been revealed in presence of 600 mM KCl at pH 3.5 but it stabilizes the protein at higher and lower pH. Hence, the effect of salt on protein stability is not linear and upfront. Therefore, it is difficult to understand clearly the role of salt from investigation at a single pH. It is essential to study the titration curve at a wide range of pHs and salt concentration to reveal the interplay between ionization processes and salt contribution.
Typically, titratable groups are considered to be Asp, Glu, His, Lys and Arg residues, since they are frequently ionized at physiological pH. However, there are other residues, typically refereed as polar residues, which may also titrate in physiological pH. The list includes Ser, Thr, Cys and Tyr residues. These groups play an important role when present in catalytic site and the state of protonation is the key to their function. Therefore, it is essential to determine the pKa for these groups to understand catalytic reaction and it’s pH dependence. For example, serine protease is an enzyme that catalyzes the hydrolysis of a peptide bond with an active site serine residue.27 The serine residue in this case acts as a nucleophile during the catalysis. Furthermore, the catalytic cycles of cysteine protease and tyrosine phosphatase involve the deprotonation of cysteine or tyrosine.28,29 Lastly, in cysteine protease family, thiolate is an essential intermediate which undergoes nucleophilic attack in the active site of protein.30,31 Therefore, it is essential to know the ionization state of serine, cysteine, tyrosine or threonine to understand the pH dependence of catalysis.
In this work, we report two new functionalities of DelPhiPKa. The first one is the addition of salt in the modeling protocol. This is done without determining solute-solvent interface, i.e. molecular surface free protocol. The upgraded DelphiPKa is benchmarked against experimentally measured pKa’s and is shown to deliver better results compared with the previous version. The second development allows for polar residues such as cysteine, serine, threonine and tyrosine to be treated as titratable residues and their pKa’s to be predicted. The DelPhiPKa predictions for pKa’s of cysteine residues in catalytic site are tested by comparing with experimental data and it is shown that the results are better than previously reported by other existing methods. DelPhiPKa is also shown to reproduce the experimental pKa for tyrosine and to predict the pKa shift based on the surrounding environment for serine.
Methods
We briefly describe the DelPhiPKa methods for predicting pKa’s of titrable residues here. Details of implementation of the method can be found elsewhere.24 DelPhiPKa method calculates the probability of protonation of each titrable residue as a function of pH and determines the pKa as pH at which probability of protonation is 50%. For doing so, the electrostatic free energy of the titrable residues in their protonated as well as deprotonated states are calculated using DelPhi built-in module. It is important to mention, that the calculations are done with smooth Gaussian-based dielectric function and do not require determining solute-solvent interface32,33. Our previous works32,33 have demonstrated that such an approach results in dielectric function inside the macromolecule that varies from low reference value of about 2 up to 20 and more depending on the atomic packing. At the van der Walls (vdW) surface it increases smoothly and reaches the value of 80 in bulk water. Thus, Gaussian-based smooth dielectric function is designed to capture several effects: (a) dielectric inhomogeneity of macromolecules; (b) existence of water cavities inside, if any; (c) the fuzziness of the macromolecule-water region; and (d) the ability of ions to visit space close to vdW surface.
Electrostatic free energy
Smooth Gaussian based dielectric function is used for all the electrostatic energy calculations, the implementation of the model in DelPhi has been described in previous work.34 The details of the calculation of electrostatic energy and it’s components (charge-charge pairwise interaction, ; polar energy term, ; desolvation energy, ) were explained in earlier paper24. These three energy components are determined for ith residue in both charged (, , ) as well as neutral state (, , ). We obtain the change in pairwise interaction, polar energy term and delsovation energy due to change in protonation state as follows:
| (1) |
| (2) |
| (3) |
Therefore, the total electrostatic energy of the protein at given protonation state is expressed as
| (4) |
In this way, total electrostatic energy of each microstate is calculated and using the distribution of this energy, probability of ionization of ith residue is calculated at a given pH. The probability of ionization of ith residue at a given pH is determined using Boltzmann distribution of corresponding electrostatic energy
| (5) |
Where system has total M microstates and at mth microstate, energy is at a given pH. is 1 when ith residue is ionized and 0 if it is neutral. The probability of ionization of each ionizable residue is calculated as a function of pH. The pH at which probability of ionization is 50%, defined as pKa. To reduce computational cost, DelPhiPKa applies clustering approach to enumerate Boltzmann distribution as described in the original paper32.
The treatment of mobile ions in the framework of PBE by adding a desolvation penalty term is introduced by Jia et al.35 The same method is used here for treating a given concentration of salt which is comprised of mobile ions. The desolvation penalty is calculated via Born equation
| (6) |
where is the Avogadro constant, z is the valence, e is the charge, is the effective radius of the ion, is the permittivity of the vacuum. and are the dielectric constant at a given location and in bulk water respectively. Therefore, incorporating the desolvation penalty in PBE, the corresponding PBE is written as
| (7) |
The term is nonzero only in the close proximity of the protein molecules, it is zero in bulk water.
Database and parameters of DelPhiPKa
In order to predict the accuracy of the method, we benchmarked the calculated pKa’s against an experimental protein pKa’s database (http://compbio.clemson.edu/databases/database_pKa.xlsx) compiled by us, which is a large dataset containing 82 proteins and 773 residues. Most of the residues of this database are surface exposed and accessible to the salt. The database includes protein pKa database (http://pka.engr.ccny.cuny.edu/) and cases compiled from literature. Note that pKa-cooperative database is not suitable for this work, since here we emphasize on the role of salt, while pKa-cooperative set is comprised of buried residues away from water phase (and thus not sensitive to salt). The structure files were taken from PDB databank. The missing atoms and residues were re-generated using PROFIX (a software module within the JACKAL package, https://honiglab.c2b2.columbia.edu/software/Jackal/Jackalmanual.htm#profix). We removed all the compounds, other than the amino acids, such as heme, SO4, PO4, all ions and water molecules. For the pKa’s calculations, we used sigma=0.70 and internal dielectric=8 in DelphiPKa. These parameters are optimized through benchmarking against experimental data.24 It is shown in an earlier paper24 that the Gaussian dielectric provides a significant improvement in pKa’s prediction compared to homogeneous model. Therefore, in this work, Gaussian smooth dielectric model is used. The atomic charges and radii are taken from Charmm forcefield parameters for all the calculations.36
Results and Discussions
With regards with the new features of DelPhiPKa reported in this work, several investigations were carried out. First, we investigated the change of the performance of DelPhiPKa upon adding the salt option; this was done by benchmarking the predicted pKa’s against experimentally determined pKa’s (see Method section) in presence and absence of salt. Secondly, we probed the performance of DelPhiPKa to predict experimentally measured pKa shifts of various residues due to variation in salt concentrations. Finally, we checked DelPhiPKa performance against experimental data of pKa’s of Cys and Tyr residues. We also tested the ability of DelPhiPKa in predicting pKa shift of Ser residue depending on the structural environment.
Does the salt option improve pKa’s predictions?
The total RMSD and correlation coefficient for the predicted pKa’s of total 752 residues (Asp, Glu, His and Lys) using DelphiPKa is 0.74 and 0.93, respectively, in presence of salt (I=0.15M). The plot of experimental versus calculated pKa’s is shown in Figure 1a. The slope and intercept of the linear fit is 0.94 and 0.41, respectively. The deviation of calculated from experimental pKa, both for all the residues as well as for each residue type, are presented in Table I. Among the pKa values of 752 residues, for 487 residues (65%), the deviation of calculated values from experimental one is less than 0.5 pK unit. For 660 residues (88%), the deviation is less than 1 pK unit and only for 3% of the residues, the shift is greater than 2 pK unit. Most of the residues, which falls under these 3%, are buried inside the protein (Table II). Thus, Table II indicates that buried residues (RSA< 10) have higher RMSDs irrespective of residue types. Better predictions can be made if larger variance of the Gaussian function is used as pointed in earlier work.32 Focusing on different residue types separately, it has been found that the RMSD for Glu and Asp is similar (0.70 and 0.72 respectively) [Figure 1a]. However, for Lys, the RMSD is significantly lower (0.39) and for HIS it’s the largest (1.08). Hence, the overall correlation of the calculated pKa’s with experiment seems exceptionally good for such a large dataset.
Figure 1.
Benchmark of calculated pKa against protein pKa database for each individual residue type at (a) I=0.15M and (b) I=0
Table I:
Statistics of the deviation of calculated pKa from the experimental dataset
| |pKaexp-pKacal| | Total (752) | Glu (258) | Asp (275) | His (111) | Lys (109) |
|---|---|---|---|---|---|
| >0.5 | 35% | 37% | 32% | 54% | 20% |
| >1.0 | 12% | 13% | 11% | 24% | 3% |
| >1.5 | 6% | 6% | 6% | 11% | --- |
| >2.0 | 3% | 2% | 4% | 6% | --- |
| >2.5 | --- | --- | 2% | 5% | --- |
| >3.0 | --- | --- | 1% | 3% | --- |
Table II:
Statistics of RMSD based on the residue positions
| Relative Surface accessible area | Total | Glu | Asp | His | Lys | |||||
|---|---|---|---|---|---|---|---|---|---|---|
| % | RMSD | % | RMSD | % | RMSD | % | RMSD | % | RMSD | |
| RSA < 10 | 0.07 | 1.75 | 0.06 | 1.41 | 0.08 | 1.66 | 0.16 | 2.07 | --- | --- |
| 10 < RSA < 20 | 0.06 | 0.96 | 0.04 | 1.27 | 0.08 | 0.49 | 0.11 | 1.31 | 0.01 | 0.17 |
| 20 < RSA < 50 | 34 | 0.64 | 34 | 0.59 | 27 | 0.72 | 45 | 0.72 | 37 | 0.44 |
| 50 < RSA | 53 | 0.53 | 57 | 0.59 | 57 | 0.51 | 30 | 0.65 | 63 | 0.36 |
The agreement between calculated pKa’s and the experimental one drops significantly when salt concentration is not taken into account. The total RMSD in this case becomes 0.94. Therefore, presence of salt plays a crucial role in determining the pKa of ionizable residues. The plot of experimental pKa vs calculated pKa for all the 752 residues in absence of salt is shown in Figure 1b. The slope and intercept for the linear fit is 1.075 and 0.63 respectively. Figure 1b also displays deviation of calculated pKa with respect to experimental one for each individual residue types.
The RMSDs for Glu, Asp and Lys in absence of salt are much higher (0.91, 0.92 and 0.94, respectively) compared to RMSDs in presence of salt (0.70, 0.72 and 0.39, respectively). For His, it’s practically the same (1.06 without salt and 1.08 in presence of salt). These results indicate that salt effect is playing a vital role in determining ionization state of titratable groups.
It should be pointed out that the improvement of the accuracy of pKa’s predictions is not only due to inclusion of the salt, but the treatment of the system via a smooth Gaussian-based dielectric function. If one uses the traditional protocol that considers solute as a low dielectric cavity with a sharp border between solute and solvent, with the same parameters and salt concentration reported above, the corresponding RMSD is 1.24 (Figure S1), compared with 0.74 above.
For accessing the quality of the results, we compared the calculated data with the null model predictions, that is setting all protein pKa shifts to 0 and so, all the pKa values are equal to the intrinsic pKa of the corresponding residues. In Table III, the RMSD of all the residue types are shown in presence of salt (I=0.15M), absence of salt (I=0) and null model. Table III reflects that DelPhiPKa outperforms the null model by an average of 0.19 pK unit. While this is a relatively small improvement, it should be pointed out that previous works were unable to improve the null model.37,16 One of the reasons for such relatively small improvement is that the experimental shifts of pKa values are only between 0.5 to 1.5 pK units.
Table III:
RMSD for each residue type
| GLU | ASP | HIS | LYS | |
|---|---|---|---|---|
| Total number | 258 | 275 | 111 | 109 |
| Calc. pKa at I=0.15 | 0.70 | 0.75 | 1.08 | 0.39 |
| Calc. pKa at I=0 | 0.91 | 0.92 | 1.06 | 0.94 |
| Null-model | 1.04 | 0.93 | 1.21 | 0.49 |
Benchmarking DelPhiPKa against experimental data of salt dependence of pKa’s
Here we investigate whether the predicted pKa shifts due to change in salt concentration correlate with the experimental data. We collected 4 datasets from several studies38,39,40 for four different proteins where pKa’s were determined experimentally as a function of salt concentration (salt concentration was varied in the range of 0.01 M to 1.5 M). Lee at al.38 performed acid/base titrations of SNase by using H NMR spectroscopy at different concentrations of KCl. It is reported that with increase in the salt concentration from 0.01 M to 1.5 M, pKa values of His-8, His-46, His-121 and His-124 increase by 0.92, 0.44, 1.05 and 0.93 pK units, respectively. We calculated the pKa values of these four His residues using DelPhiPKa. The salt dependencies of experimental and calculated pKa values for SNase are compared in supporting information (Table S1). The correlation between the calculated and experimental pKa values and the variations of both experimental and calculated pKa’s as a function of salt concentration for these four His residues are shown in Figure S2 and Figure S3, respectively. The correlation coefficient (R) of 0.95, 0.96, 0.98 and 0.99 for His-8, His-46, His-121 and His-124, respectively, indicates that DelPhiPKa methods can well reproduce the salt dependence of measured pKa values.
Kao et al.39 investigated the salt dependence of His pKa values of sperm whale myglobin (Pc-Mb). The experimentally measured pKa values indicate that despite of having difference in their solvent accessibility and nature of the surrounding residues, most of the His residues exhibit similar rise of their pKa values (~0.3 pK unit) on changing the salt concentration from 0.02 M to 1.5 M (Table S2). The correlation of experimental versus calculated pKa’s and the change in experimental as well as predicted pKa’s with variation of salt concentration for all the His residues are presented in Figures S4 and S5, respectively. In Table S2, for His-36, measured experimental value indicates that pKa’s are insensitive to salt concentration between 0.02 M and 1.5 M. DelPhiPKa also predicts that there is a negligible increase (0.02 pK unit) in pKa’s by increasing the salt concentration from 0.02 to 0.2M and a slight change of 0.05 pK unit going from 0.5 M to 1.5 M of salt. This observation reveals that the DelPhiPKa model is reasonably accurate in predicting the effect of salt. We estimated the correlation coefficient of the calculated pKa’s of all the His residues except His-36 are plotted in Figure S4. We obtained the correlation coefficient of 0.98, 0.91, 0.85, 0.98, 0.63 and 0.99 for His-12, His-48, His-81, His-113, His-116 and His-119, respectively for Pc-Mb. In the same paper,39 authors also described the salt sensitivity of His pKa for horse heart Mb (Eq-Mb). The calculated pKa’s are compared with the experimentally measured pKa’s in Table S3. Similarly, to Pc-Mb, for His-36, the pKa values are found to be not influenced by salt concentration both in experiment and model calculation. Correlation between experimental and calculated pKa’s and variation of these pKa’s with salt concentration are plotted in Figure S6 and S7 respectively for the 6 His residues in Eq-Mb excluding His-36. We can see in Figure S6 that DelPhiPKa predicted pKa’s behave the same way as experimental pKa’s, as we change the salt concentration.
Abe et al.40 evaluated the influence of salt concentration on the pKa values of acidic residues in hen egg white lysozyme. The predicted pKa values along with experimental pKa’s are shown in Table S4 which reveals that except for Glu35 and Asp87, DelPhiPKa can capture the salt sensitivity of the pKa for other residues similar way that in experiment. Glu35 is located in the active site of lysozyme and it’s not surface exposed. Therefore, the influence of the presence of salt is negligible in this case. We can observe in Table S4 that the pKa of Asp87 decreases with increase of salt concentration from 0.005 M to 0.1M but increases between 0.1 M and 0.4 M. Such an unusual behavior may be caused by the fact that Asp87 is located at the N-terminal of α-helix (residues 88–98) and the conformation may be affected by the salt concentration. Since in our calculations the structure is kept rigid, this may explain why we were unable to reproduce the abnormal change of experimentally measured pKa. The correlation between predicted and experimental pKa’s are estimated and plotted for six acidic residues (except Glu35 and Asp87) in Figure S8. The correlation coefficient for Glu7, Asp18, Asp48, Asp52, Asp101 and Asp119 is determined to be 0.95, 0.99, 0.88, 0.94, 0.93 and 0.99, respectively. In Figure S9, the variation of both experimental and calculated pKa with varying salt concentration is plotted.
Lastly, we compared all the calculated pKa’s with the corresponding experimental data coming from 78 residues and four proteins and plotted them together in Figure 2 (excluding His-36 in both Pc-Mb and Eq-Mb, Glu35 and Asp87 in lysozyme). The correlation coefficient of 0.92 and RMSD of 0.54 indicate that DelPhiPKa method can accurately predict the effects of salt concentration.
Figure 2:
Overall correlation between experimental and calculated pKa’s in various salt concentrations.
Benchmarking Cys pKa’s.
The pKa of a Cys residue, not involved in sulphate bridge, can vary significantly compared to their intrinsic pKa. It was shown experimentally that pKa’s of Cys residues in non-catalytic sites are usually in the range of 7 to 9 whereas in the catalytic site they can be as low as 2.88.41,42,43 Thus, the pKa value of Cys residue depends on the environment that protein provides. Rowley et al.44 generated a Cys pKa’s experimental data set comprised of 18 Cys residues in 12 proteins including both elevated and lowered pKa’s. The same work calculated the pKa of these 18 residues with three methods using implicit solvent model (MCCE, PROPKA and H++) as well as explicit solvent models with CHARMM36 and AMBER force field. The article concluded that the methods using implicit solvent model is not reliable in predicting Cys pKa’s with RMSD between 3.41 and 4.72 whereas the RMSD obtained using explicit solvent models varied between 2.40 and 3.20. It was suggested by the authors that the methods still need to be improved for predicting Cys pKa since the null-model predicts the RMSD of 2.74 which is close to the best RMSD achieved by the explicit model. Here, we benchmark the new DelPhiPKa against the same dataset. The calculated pKa’s are compared with the experimentally measured data and plotted in Figure 3. The RMSD obtained is 1.73 which is the lowest among all the five methods including implicit as well as explicit solvent models. Thus, the new version of DelPhiPKa outperforms all other existing methods (implicit and explicit water models) and also the null-model in predicting the Cys pKa’s.
Figure 3:
Correlation between experimental and calculated cysteine pKa’s using the new DelPhiPKa.
Benchmarking Tyr pKa’s
There are only few experimental pKa values for Tyr residues reported in the literature45 among which structure files are available for three of them.46,47 Because of that, we calculated and compared the pKa’s only for these three residues belonging to two different proteins. The predicted pKa for Tyr30 and Tyr49 in ribonuclease Sa (RNase Sa) is 10.42 and 10.60 respectively whereas the experimentally measured pKa’s are 11.3 and 10.63 respectively.45 Khare et al.47 measured the pKa of Tyr33 in the B1 immunoglobulin G- (IgG-) binding domains of protein G, the value of which is 11 whereas DelPhiPKa predicts the pKa as 10.72. These results indicate that DelPhiPKa is successful in predicting pKa’s for Tyr residues as well (an example of Ser pKa is provided in supplementary material, Figure S10).
Conclusions
A new upgraded version of DelPhiPka is reported in this work that enables salt concentration to be included in the surface-free protocol. In this way, DelPhiPKa remains a unique continuum electrostatic approach to calculate pKa’s without determining the boundary between solute and solvent. The inclusion of the salt in the calculations is shown to deliver better results comparing with the case without salt. This confirms previous observation made in case of salt dependent of protein-protein binding.35 It should be pointed out that the improvement in pKa’s predictions is significant (about 0.2pK units) and it is achieved on larger dataset (larger than the original pKa dataset, see Method section). Furthermore, the protocol was tested against experimentally determined changes of pKa’s upon change of the salt concentration and very good correlations were obtained. It should be mentioned that both experimental and computed pKa changes are very small (less than half of pKa unit), which makes the comparison quite difficult. Lastly, DelPhiPKa was enabled to calculate pKa’s of groups typically referred as polar groups. It was shown that the method outperforms all existing method and lower the RMSD even compared to explicit water models. Thus, here we report important enrichments of DelPhiPKa capabilities and significant improvement of its prediction accuracy. The improvement of prediction accuracy is attributed to both, the Gaussian-based smooth dielectric function and the novel treatment of ions.
Supplementary Material
Acknowledgments
Acknowledgement: We thank Neocles Basil for pointing to us the necessity of allowing pKa’s predictions for polar groups. The work was supported by a grant from NIH, grant number R01GM093937.
References:
- 1.Honig B, Nicholls A. Classical electrostatics in biology and chemistry. Science. 1995;268(5214):1144. [DOI] [PubMed] [Google Scholar]
- 2.Warshel A, Russell ST. Calculations of electrostatic interactions in biological systems and in solutions. Quarterly Reviews of Biophysics. 1984;17(3):283–422. [DOI] [PubMed] [Google Scholar]
- 3.Onufriev AV, Alexov E. Protonation and pK changes in protein–ligand binding. Quarterly Reviews of Biophysics. 2013;46(2):181–209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Cherny VV, Murphy R, Sokolov V, Levis RA, DeCoursey TE. Properties of Single Voltage-gated Proton Channels in Human Eosinophils Estimated by Noise Analysis and by Direct Measurement. The Journal of General Physiology. 2003;121(6):615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Bashford D Macroscopic electrostatic models for protonation states in proteins. Vol 92004. [DOI] [PubMed] [Google Scholar]
- 6.Alexov E, Mehler Ernest L, Baker N, et al. Progress in the prediction of pKa values in proteins. Proteins: Structure, Function, and Bioinformatics. 2011;79(12):3260–3275. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Nicholls A, Honig B. A rapid finite difference algorithm, utilizing successive over-relaxation to solve the Poisson–Boltzmann equation. Journal of Computational Chemistry. 1991;12(4):435–445. [Google Scholar]
- 8.Bashford D An object-oriented programming suite for electrostatic effects in biological molecules An experience report on the MEAD project. Paper presented at: Scientific Computing in Object-Oriented Parallel Environments; 1997//, 1997; Berlin, Heidelberg. [Google Scholar]
- 9.Mehler EL. The Lorentz–Debye–Sack Theory and Dielectric Screening of Electrostatic Effects in Proteins and Nucleic Acids. Vol 31996. [Google Scholar]
- 10.Schutz Claudia N, Warshel A. What are the dielectric “constants” of proteins and how to validate electrostatic models? Proteins: Structure, Function, and Bioinformatics. 2001;44(4):400–417. [DOI] [PubMed] [Google Scholar]
- 11.Godoy-Ruiz R, Perez-Jimenez R, Garcia-Mira MM, Plaza del Pino IM, Sanchez-Ruiz JM. Empirical parametrization of pK values for carboxylic acids in proteins using a genetic algorithm. Biophysical Chemistry. 2005;115(2):263–266. [DOI] [PubMed] [Google Scholar]
- 12.Spassov VZ, Karshikov AD, Atanasov BP. Electrostatic interactions in proteins. A theoretical analysis of lysozyme ionization. Biochimica et Biophysica Acta (BBA) - Protein Structure and Molecular Enzymology. 1989;999(1):1–6. [Google Scholar]
- 13.Krieger E, Nielsen JE, Spronk CAEM, Vriend G. Fast empirical pKa prediction by Ewald summation. Journal of Molecular Graphics and Modelling. 2006;25(4):481–486. [DOI] [PubMed] [Google Scholar]
- 14.Li H, Robertson Andrew D, Jensen Jan H. Very fast empirical prediction and rationalization of protein pKa values. Proteins: Structure, Function, and Bioinformatics. 2005;61(4):704–721. [DOI] [PubMed] [Google Scholar]
- 15.Bas Delphine C, Rogers David M, Jensen Jan H. Very fast prediction and rationalization of pKa values for protein–ligand complexes. Proteins: Structure, Function, and Bioinformatics. 2008;73(3):765–783. [DOI] [PubMed] [Google Scholar]
- 16.Olsson MHM, Søndergaard CR, Rostkowski M, Jensen JH. PROPKA3: Consistent Treatment of Internal and Surface Residues in Empirical pKa Predictions. Journal of Chemical Theory and Computation. 2011;7(2):525–537. [DOI] [PubMed] [Google Scholar]
- 17.Søndergaard CR, Olsson MHM, Rostkowski M, Jensen JH. Improved Treatment of Ligands and Coupling Effects in Empirical Calculation and Rationalization of pKa Values. Journal of Chemical Theory and Computation. 2011;7(7):2284–2295. [DOI] [PubMed] [Google Scholar]
- 18.You TJ, Bashford D. Conformation and hydrogen ion titration of proteins: a continuum electrostatic model with conformational flexibility. Biophysical Journal. 1995;69(5):1721–1733. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Beroza P, Case DA. Including Side Chain Flexibility in Continuum Electrostatic Calculations of Protein Titration. The Journal of Physical Chemistry. 1996;100(51):20156–20163. [Google Scholar]
- 20.Warwicker J Improved pK(a) calculations through flexibility based sampling of a water-dominated interaction scheme. Protein Science : A Publication of the Protein Society. 2004;13(10):2793–2805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Nielsen JE, Andersen KV, Honig B, et al. Improving macromolecular electrostatics calculations. Protein Engineering, Design and Selection. 1999;12(8):657–662. [DOI] [PubMed] [Google Scholar]
- 22.Nielsen Jens E, Vriend G. Optimizing the hydrogen-bond network in Poisson–Boltzmann equation-based pKa calculations. Proteins: Structure, Function, and Bioinformatics. 2001;43(4):403–412. [DOI] [PubMed] [Google Scholar]
- 23.Georgescu RE, Alexov EG, Gunner MR. Combining conformational flexibility and continuum electrostatics for calculating pK(a)s in proteins. Biophysical Journal. 2002;83(4):1731–1748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Wang L, Li L, Alexov E. pKa predictions for proteins, RNAs, and DNAs with the Gaussian dielectric function using DelPhi pKa. Proteins: Structure, Function, and Bioinformatics. 2015;83(12):2186–2197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Petukh M, Kimmet T, Alexov E. BION web server: predicting non-specifically bound surface ions. Bioinformatics. 2013;29(6):805–806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Petukh M, Zhenirovskyy M, Li C, Li L, Wang L, Alexov E. Predicting nonspecific ion binding using DelPhi. Biophys J. 2012;102(12):2885–2893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Sprang S, Standing T, J Fletterick R, et al. The Three-Dimensional Structure of Asn102 Mutant of Trypsin: Role of Asp102 in Serine Protease Catalysis. Vol 2371987. [DOI] [PubMed] [Google Scholar]
- 28.Denu JM, Dixon JE. Protein tyrosine phosphatases: mechanisms of catalysis and regulation. Current Opinion in Chemical Biology. 1998;2(5):633–641. [DOI] [PubMed] [Google Scholar]
- 29.Kolmodin K, Åqvist J. The catalytic mechanism of protein tyrosine phosphatases revisited. FEBS Letters. 2001;498(2):208–213. [DOI] [PubMed] [Google Scholar]
- 30.Otto H- H, Schirmeister T. Cysteine Proteases and Their Inhibitors. Chemical Reviews. 1997;97(1):133–172. [DOI] [PubMed] [Google Scholar]
- 31.Chapman HA, Riese RJ, Shi G- P. EMERGING ROLES FOR CYSTEINE PROTEASES IN HUMAN BIOLOGY. Annual Review of Physiology. 1997;59(1):63–88. [DOI] [PubMed] [Google Scholar]
- 32.Wang L, Li L, Alexov E. pKa predictions for proteins, RNAs, and DNAs with the Gaussian dielectric function using DelPhi pKa. Proteins. 2015;83(12):2186–2197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Li L, Li C, Zhang Z, Alexov E. On the Dielectric “Constant” of Proteins: Smooth Dielectric Function for Macromolecular Modeling and Its Implementation in DelPhi. J Chem Theory Comput. 2013;9(4):2126–2136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Li L, Li C, Zhang Z, Alexov E. On the Dielectric “Constant” of Proteins: Smooth Dielectric Function for Macromolecular Modeling and Its Implementation in DelPhi. Journal of Chemical Theory and Computation. 2013;9(4):2126–2136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Jia Z, Li L, Chakravorty A, Alexov E. Treating ion distribution with Gaussian-based smooth dielectric function in DelPhi. Journal of Computational Chemistry. 2017;38(22):1974–1979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Best RB, Zhu X, Shim J, et al. Optimization of the Additive CHARMM All-Atom Protein Force Field Targeting Improved Sampling of the Backbone ϕ, ψ and Side-Chain χ1 and χ2 Dihedral Angles. Journal of Chemical Theory and Computation. 2012;8(9):3257–3273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Antosiewicz J, McCammon JA, Gilson MK. The Determinants of pKas in Proteins. Biochemistry. 1996;35(24):7819–7833. [DOI] [PubMed] [Google Scholar]
- 38.Lee KK, Fitch CA, Lecomte JTJ, García-Moreno EB. Electrostatic Effects in Highly Charged Proteins: Salt Sensitivity of pKa Values of Histidines in Staphylococcal Nuclease. Biochemistry. 2002;41(17):5656–5667. [DOI] [PubMed] [Google Scholar]
- 39.Kao Y- H, Fitch CA, Bhattacharya S, Sarkisian CJ, Lecomte JTJ, García-Moreno EB. Salt Effects on Ionization Equilibria of Histidines in Myoglobin. Biophysical Journal. 2000;79(3):1637–1654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Abe Y, Ueda T, Iwashita H, et al. Effect of Salt Concentration on the pKa of Acidic Residues in Lysozyme. The Journal of Biochemistry. 1995;118(5):946–952. [DOI] [PubMed] [Google Scholar]
- 41.Pinitglang S, Watts AB, Patel M, et al. A Classical Enzyme Active Center Motif Lacks Catalytic Competence until Modulated Electrostatically. Biochemistry. 1997;36(33):9968–9982. [DOI] [PubMed] [Google Scholar]
- 42.Zhang ZY, Dixon JE. Active site labeling of the Yersinia protein tyrosine phosphatase: The determination of the pKa of the active site cysteine and the function of the conserved histidine 402. Biochemistry. 1993;32(36):9340–9345. [DOI] [PubMed] [Google Scholar]
- 43.Lohse DL, Denu JM, Santoro N, Dixon JE. Roles of Aspartic Acid-181 and Serine-222 in Intermediate Formation and Hydrolysis of the Mammalian Protein-Tyrosine-Phosphatase PTP1. Biochemistry. 1997;36(15):4568–4575. [DOI] [PubMed] [Google Scholar]
- 44.Awoonor-Williams E, Rowley CN. Evaluation of Methods for the Calculation of the pKa of Cysteine Residues in Proteins. Journal of Chemical Theory and Computation. 2016;12(9):4662–4673. [DOI] [PubMed] [Google Scholar]
- 45.Grimsley Gerald R, Scholtz JM, Pace CN. A summary of the measured pK values of the ionizable groups in folded proteins. Protein Science. 2008;18(1):247–251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Laurents DV, Huyghues-Despointes BMP, Bruix M, et al. Charge–Charge Interactions are Key Determinants of the pK Values of Ionizable Groups in Ribonuclease Sa (pI=3.5) and a Basic Variant (pI=10.2). Journal of Molecular Biology. 2003;325(5):1077–1092. [DOI] [PubMed] [Google Scholar]
- 47.Khare D, Alexander P, Antosiewicz J, Bryan P, Gilson M, Orban J. pKa Measurements from Nuclear Magnetic Resonance for the B1 and B2 Immunoglobulin G-Binding Domains of Protein G: Comparison with Calculated Values for Nuclear Magnetic Resonance and X-ray Structures. Biochemistry. 1997;36(12):3580–3589. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.



