Hostname: page-component-77c89778f8-rkxrd Total loading time: 0 Render date: 2024-07-21T07:26:11.592Z Has data issue: false hasContentIssue false

Protonation and pK changes in protein–ligand binding

Published online by Cambridge University Press:  29 July 2013

Alexey V. Onufriev*
Affiliation:
Departments of Computer Science and Physics, 2050 Torgersen Hall, Virginia Tech, Blacksburg, VA 24061, USA
Emil Alexov*
Affiliation:
Department of Physics and Astronomy, Clemson University, Clemson, SC 29634, USA
*
*Author for correspondence: Alexey V. Onufriev, Departments of Computer Science and Physics, 2050 Torgersen Hall, Virginia Tech, Blacksburg, VA 24061, USA. Email: alexey@cs.vt.edu
*Author for correspondence: Emil Alexov, Department of Physics and Astronomy, Clemson University, Clemson, SC 29634, USA. Email: ealexov@clemson.edu

Abstract

Formation of protein–ligand complexes causes various changes in both the receptor and the ligand. This review focuses on changes in pK and protonation states of ionizable groups that accompany protein–ligand binding. Physical origins of these effects are outlined, followed by a brief overview of the computational methods to predict them and the associated corrections to receptor–ligand binding affinities. Statistical prevalence, magnitude and spatial distribution of the pK and protonation state changes in protein–ligand binding are discussed in detail, based on both experimental and theoretical studies. While there is no doubt that these changes occur, they do not occur all the time; the estimated prevalence varies, both between individual complexes and by method. The changes occur not only in the immediate vicinity of the interface but also sometimes far away. When receptor–ligand binding is associated with protonation state change at particular pH, the binding becomes pH dependent: we review the interplay between sub-cellular characteristic pH and optimum pH of receptor–ligand binding. It is pointed out that there is a tendency for protonation state changes upon binding to be minimal at physiologically relevant pH for each complex (no net proton uptake/release), suggesting that native receptor–ligand interactions have evolved to reduce the energy cost associated with ionization changes. As a result, previously reported statistical prevalence of these changes – typically computed at the same pH for all complexes – may be higher than what may be expected at optimum pH specific to each complex. We also discuss whether proper account of protonation state changes appears to improve practical docking and scoring outcomes relevant to structure-based drug design. An overview of some of the existing challenges in the field is provided in conclusion.

Type
Review Article
Copyright
Copyright © Cambridge University Press 2013 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aguilar, B., Anandakrishnan, R., Ruscio, J. Z. & Onufriev, A. V. (2010). Statistics and physical origins of pK and ionization state changes upon protein–ligand binding. Biophysics Journal 98, 872880.CrossRefGoogle ScholarPubMed
Alexov, E. (2004). Calculating proton uptake/release and binding free energy taking into account ionization and conformation changes induced by protein–inhibitor association: application to plasmepsin, cathepsin D and endothiapepsin–pepstatin complexes. Proteins 56, 572584.CrossRefGoogle ScholarPubMed
Alexov, E., Mehler, E. L., Baker, N., Baptista, A. M., Huang, Y., Milletti, F., Nielsen, J. E., Farrell, D., Carstensen, T., Olsson, M. H., Shen, J. K., Warwicker, J., Williams, S. & Word, J. M. (2011). Progress in the prediction of pka values in proteins. Proteins 79, 3260–75.CrossRefGoogle ScholarPubMed
Anandakrishnan, R., Aguilar, B. & Onufriev, A. V. (2012). H++ 3.0: automating pK prediction and the preparation of biomolecular structures for atomistic molecular modeling and simulations. Nucleic Acids Resarch 40 (Web Server issue), 537541.CrossRefGoogle ScholarPubMed
Antosiewicz, J., McCammon, J. A. & Gilson, M. K. (1994). Prediction of pH-dependent properties of proteins. Journal of Molecnlor Biology 238, 415436.CrossRefGoogle ScholarPubMed
Baker, B. M. & Murphy, K. P. (1996). Evaluation of linked protonation effects in protein binding reactions using isothermal titration calorimetry. Biophysical Journal 71, 20492055.CrossRefGoogle ScholarPubMed
Bas, D. C., Rogers, D. M. & Jensen, J. H. (2008). Very fast prediction and rationalization of pKa values for protein–ligand complexes. Proteins 73, 765783.CrossRefGoogle ScholarPubMed
Bashford, D. & Karplus, M. (1990). pKa's of ionizable groups in proteins: atomic detail from a continuum electrostatic model. Biochemistry 29, 1021910225.CrossRefGoogle ScholarPubMed
Beroza, P., Fredkin, D. R., Okamura, M. Y. & Feher, G. (1991). Protonation of interacting residues in a protein by Monte Carlo method. Proceeding of the National Academy of Sciences of the United States of America 88, 58045808.CrossRefGoogle Scholar
Beroza, P., Fredkin, D. R., Okamura, M. Y. & Feher, G. (1991). Protonation of interacting residues in a protein by a Monte Carlo method: application to lysozyme and the photosynthetic reaction center of rhodobacter sphaeroides. Proceeding of the National Academy of Sciences of the United States of America 88, 58045808.CrossRefGoogle Scholar
Betts, M. J. & Sternberg, M. J. E. (1999). An analysis of conformational changes on protein–protein association: implications for predictive docking. Protein Engineering 12, 271283.CrossRefGoogle ScholarPubMed
Bidwai, A. K., Ok, E. Y. & Erman, J. E. (2008). pH dependence of cyanide binding to the ferric heme domain of the direct oxygen sensor from Escherichia coli and the effect of alkaline denaturation. Biochemistry 47, 1045810470.CrossRefGoogle Scholar
Bjarnadottir, U. & Nielsen, J. E. E. (2010). Calculating pKa values in the cAMP-dependent protein kinase: the effect of conformational change and ligand binding. Protein Science: a Publication of the Protein Society 19, 24852497.CrossRefGoogle ScholarPubMed
Blundell, C. D., Mahoney, D. J., Cordell, M. R., Almond, A., Kahmann, J. D., Perczel, A., Taylor, J. D., Campbell, I. D. & Day, A. J. (2007). Determining the molecular basis for the pH-dependent interaction between the link module of human TSG-6 and hyaluronan. Journal of Biological Chemistry 282, 1297612988.CrossRefGoogle ScholarPubMed
Boehr, D. D. & Wright, P. E. (2008). Biochemistry: how do proteins interact? Science 320, 14291430.CrossRefGoogle ScholarPubMed
Bombarda, E. & Ullmann, G. M. (2010). pH-dependent pKa values in proteins: a theoretical analysis of protonation energies with practical consequences for enzymatic reactions. The Journal of Physical Chemistry B 114, 19942003.CrossRefGoogle ScholarPubMed
Bordner, A. J. & Abagyan, R. (2005). Statistical analysis and prediction of protein–protein interfaces. Proteins 60, 353366.CrossRefGoogle Scholar
Brandsdal, B. O., Smalås, A. O. & Åqvist, J. (2006). Free energy calculations show that acidic p1 variants undergo large pKa shifts upon binding to trypsin. Proteins 64, 740748.CrossRefGoogle ScholarPubMed
Bruylants, G., Wintjens, R., Looze, Y., Redfield, C. & Bartik, K. (2007). Protonation linked equilibria and apparent affinity constants: the thermodynamic profile of the alpha–chymotrypsin–proflavin interaction. European Biophysics Journal 37, 1118.CrossRefGoogle ScholarPubMed
Buckle, A. M., Schreiber, G. & Fersht, A. R. (1994). Protein–protein recognition: crystal structural analysis of a barnase–barstar complex at 2.0-a resolution. Biochemistry 33, 8878–89.CrossRefGoogle Scholar
Buczek, O., Koscielska-Kasprzak, K., Krowarsch, D., Dadlez, M. & Otlewski, J. (2002). Analysis of serine proteinase–inhibitor interaction by alanine shaving. Protein Sciences 11, 806819.CrossRefGoogle ScholarPubMed
Carter, P., Lesk, V., Islam, S. & Sternberg, M. (2005). Protein–protein docking using 3d-dock in rounds 3,4 of capri. Proteins 60, 281288.CrossRefGoogle Scholar
Cassidy, C. S., Lin, J. & Frey, P. A. (1997). A new concept for the mechanism of action of chymotrypsin: the role of the low-barrier hydrogen bond. Biochemistry 36, 45764584.CrossRefGoogle ScholarPubMed
Chan, P., Lovric, J. & Warwicker, J. (2006). Subcellular ph and predicted pH-dependent features of proteins. Proteomics, 6, 34943501.CrossRefGoogle ScholarPubMed
Chang, A., Scheer, M., Grote, A., Schomburg, I. & Schomburg, D. (2009). Brenda, amenda and frenda the enzyme information system: new content and tools in 2009. Nucleic Acids Research 35, D588D592.CrossRefGoogle Scholar
Cho, A. E., Guallar, V., Berne, B. J. & Friesner, R. (2005). Importance of accurate charges in molecular docking: quantum mechanical/molecular mechanical (QM/MM) approach. Journal of Computational Chemistry 26, 915931.CrossRefGoogle ScholarPubMed
Clackson, T. & Wells, J. A. (1995). A hot spot of binding energy in a hormone–receptor interface. Science 267, 383386.CrossRefGoogle Scholar
Commet, A., Boswell, N., Yocum, C. F. & Popelka, H. (2012). pH optimum of the photosystem ii h(2)o oxidation reaction: effects of psbo, the manganese-stabilizing protein, cl(-) retention, and deprotonation of a component required for o(2) evolution activity. Biochemistry 51, 38083818.CrossRefGoogle Scholar
Czodrowski, P., Sotriffer, C. A. & Klebe, G. (2007). Protonation changes upon ligand binding to trypsin and thrombin: structural interpretation based on pKa calculations and ITC experiments. Journal of Molecular Biology 367, 13471356.CrossRefGoogle ScholarPubMed
Dan, A., Ofran, Y. & Kliger, Y. (2009). Large-scale analysis of secondary structure changes in proteins suggests a role for disorder-to-order transitions in nucleotide binding proteins. Proteins 78, 236248, 1 February 2010.CrossRefGoogle Scholar
Davenport, H. (1996). Physiology of the Digestive Tract. Chicago: Year Book Medical Publishers, Inc.Google Scholar
DelBuono, G. S., Figueirido, F. E. & Levy, R. M. (1994). Intrinsic pKa s of ionizable residues in proteins: an explicit solvent calculation for lysozyme. Proteins 20, 8597.CrossRefGoogle Scholar
Demchuk, E. & Wade, R. C. (1996). Improving the continuum dielectric approach to calculating pKas of ionizable groups in proteins. The Journal of Physical Chemistry 100, 1737317387.CrossRefGoogle Scholar
Djurdjevic-Pahl, A., Hewage, C. & Malthouse, J. P. (2005). Ionisations within a subtilisinglyoxal inhibitor complex. Biochimica et Biophysica Acta (BBA) – Proteins & Proteomics 1749, 3341.CrossRefGoogle ScholarPubMed
Dolinsky, T. J., Czodrowski, P., Li, H., Nielsen, J. E., Jensen, J. H., Klebe, G. & Baker, N. A. (2007). PDB2PQR: expanding and upgrading automated preparation of biomolecular structures for molecular simulations. Nucleic Acids Research 35 2007 July; (Web Server issue), W522W525.CrossRefGoogle ScholarPubMed
Donnini, S., Villa, A., Groenhof, G., Mark, A. E., Wierenga, R. K. & Juffer, A. H. (2009). Inclusion of ionization states of ligands in affinity calculations. Proteins 76, 138150.CrossRefGoogle ScholarPubMed
Elcock, A. H., Sept, D. & Mccammon, J. A. (2001). Computer simulation of protein–protein interactions. Journal of Physical Chemistry B 105, 15041518.CrossRefGoogle Scholar
Garcia-Moreno, B. (2009). Adaptations of proteins to cellular and subcellular pH. Journal of Biology 8, 96.CrossRefGoogle ScholarPubMed
Ge, X. & Roux, B. (2010). Absolute binding free energy calculations of sparsomycin analogs to the bacterial ribosome. Journal of. Physical Chemistry B 114, 95259539.CrossRefGoogle Scholar
Georgescu, R., Alexov, E. & Gunner, M. (2002). Combining conformational flexibility and continuum electrostatics for calculating pKas in proteins. Biophysical. Journal 83, 17311748.CrossRefGoogle ScholarPubMed
Gerstein, M., Lesk, A. M. & Chothia, C. (1994). Structural mechanisms for domain movements in proteins. Biochemistry 33, 67396749.CrossRefGoogle ScholarPubMed
Gilson, M. K. & Zhou, H. X. (2007). Calculation of protein–ligand binding affinities. Annual Review of Biophyises and Biomolecular Structure 36, 2142.CrossRefGoogle ScholarPubMed
Gilson, M. K. (1993). Multiple-site titration and molecular modeling: two rapid methods for computing energies and forces for ionizable groups in proteins. Proteins 15, 266282.CrossRefGoogle ScholarPubMed
Goh, C-S., Milburn, D. & Gerstein, M. (2004). Conformational changes associated with protein–protein interactions. Current Opinion Structure Biology 14, 104109.CrossRefGoogle ScholarPubMed
Gohlke, H., Hendlich, M. & Klebe, G. (2000). Knowledge-based scoring function to predict protein–ligand interactions. Journal of Molecular Biology 295, 337356.CrossRefGoogle ScholarPubMed
Gohlke, H., Kiel, C. & Case, D. A. (2003). Insights into protein–protein binding by binding free energy calculation and free energy decomposition for the ras–raf and ras–ralgds complexes. Journal of Molecular Biology 330, 891913.CrossRefGoogle ScholarPubMed
Gohlke, H. & Klebe, G. (2002). Approaches to the description and prediction of the binding affinity of Small-molecule ligands to macromolecular receptors. Angewandte Chemie International Edition 41, 26442676.3.0.CO;2-O>CrossRefGoogle Scholar
Gómez, J. & Freire, E. (1995). Thermodynamic mapping of the inhibitor site of the aspartic protease endothiapepsin. Journal of Molecular Biology 252, 337350.CrossRefGoogle ScholarPubMed
Gordon, J. C., Myers, J. B., Folta, T., Shoja, V., Heath, L. S. & Onufriev, A. (2005). H++: a server for estimating pKas and adding missing hydrogens to macromolecules. Nucleic Acids Research, 33 (Web Server issue) 368371.CrossRefGoogle Scholar
Green, D. F. & Tidor, B. (2005). Design of improved protein inhibitors of HIV-1 cell entry: optimization of electrostatic interactions at the binding interface. Proteins 60, 644657.CrossRefGoogle ScholarPubMed
Grimsley, G. R., Scholtz, J. M. & Pace, C. N. (2009). A summary of the measured pK values of the ionizable groups in folded proteins. Protein Sciences 18, 247251.CrossRefGoogle ScholarPubMed
Guyton, A. (1976). Textbook of Medical Physiology. Philadelphia: W.B. Saunders Co.Google Scholar
Hansen, M. J., Olsen, J. G., Bernichtein, S., O'Shea, C., Sigurskjold, B. W., Goffin, V. & Kragelund, B. B. (2011). Development of prolactin receptor antagonists with reduced pH-dependence of receptor binding. Journal of Molecnlar Recognition 24, 533547.CrossRefGoogle ScholarPubMed
Holmes, K C., Jon, I. A., Jahn, W. & Schroder, R. R. (2003). Electron cryo-microscopy shows how strong binding of myosin to actin releases nucleotide. Nature 425, 423427.CrossRefGoogle ScholarPubMed
Horn, J. R., Ramaswamy, S. & Murphy, K. P. (2003). Structure and energetics of protein–protein interactions: the role of conformational heterogeneity in OMTKY3 binding to serine proteases. Journal of Molecular Biology 331, 497508.CrossRefGoogle ScholarPubMed
Jensen, J. H. (2008). Calculating pH and salt dependence of protein-protein binding. Current Pharmaceutical Biotechnology 9, 96102.CrossRefGoogle ScholarPubMed
Jones, S. & Thornton, J. M. (1997). Principles of protein–protein interactions. 93, 1320.Google Scholar
Jorgensen, W. L. (2004). The many roles of computation in drug discovery. Science 303, 18131818.CrossRefGoogle ScholarPubMed
Joshi, M. D., Sidhu, G., Nielsen, J. E., Brayer, G. D., Withers, S. G. & Mcintosh, L. P. (2001). Dissecting the electrostatic interactions and pH-dependent activity of a family 11 glycosidase. Biochemistry 40, 1011510139.CrossRefGoogle ScholarPubMed
Kangas, E. & Tidor, B. (1999). Charge optimization leads to favorable electrostatic binding free energy. Physical Review E 59, 59585961.CrossRefGoogle ScholarPubMed
Khandogin, J. & Brooks, C. L. (2006). Toward the accurate first-principles prediction of ionization equilibria in proteins. Biochemistry 45, 93639373.CrossRefGoogle ScholarPubMed
Kieseritzky, G. & Knapp, E. W. (2008a). Improved pKa prediction: combining empirical and semimicroscopic methods. Journal of Computational Chemistry 29, 25752581.CrossRefGoogle ScholarPubMed
Kieseritzky, G. & Knapp, E-W. (2008b). Optimizing pKA computation in proteins with pH adapted conformations. Proteins 71, 13351348.CrossRefGoogle Scholar
Kozlov, A. G. & Lohman, T. M. (2000). Large contributions of coupled protonation equilibria to the observed enthalpy and heat capacity changes for ssDNA binding to escherichia coli SSB protein. Proteins Suppl 4, 822.3.0.CO;2-H>CrossRefGoogle Scholar
Kresheck, G. C., Vitello, L. B. & Erman, J. E. (1995). Calorimetric studies on the interaction of horse ferricytochrome c and yeast cytochrome c peroxidase. Biochemistry 34, 83988405.CrossRefGoogle ScholarPubMed
Krieger, E., Dunbrack, R. L. Jr., Hooft, R. W. & Kriege, B. (2012). Assignment of protonation states in proteins and ligands: combining pKa prediction with hydrogen bonding network optimization. Methods in Molecular Biology 819, 405421.CrossRefGoogle ScholarPubMed
Krowarsch, D., Dadlez, M., Buczek, O., Krokoszynska, I., Smalas, A. O. & Otlewski, J. (1999). Interscaffolding additivity: binding of p1 variants of bovine pancreatic trypsin inhibitor to four serine proteases. Journal of Molecular Biology 289, 175186.CrossRefGoogle ScholarPubMed
Kulichikhin, K. Y., Greenway, H., Byrne, L. & Colmer, T. D. (2009). Regulation of intracellular pH during anoxia in rice coleoptiles in acidic and near neutral conditions. Journal of Experimental Botany 60, 21192128.CrossRefGoogle ScholarPubMed
Kundrotas, P. J. & Alexov, E. (2006). Electrostatic properties of protein-protein complexes. Biophysies Journal 91, 17241736.CrossRefGoogle ScholarPubMed
Labute, P. (2009). Protonate3d: assignment of ionization states and hydrogen coordinates to macromolecular structures. Proteins 75, 187205.CrossRefGoogle ScholarPubMed
Lee, M. S., Salsbury, F. R. & Brooks, C. L. (2004). Constant-pH molecular dynamics using continuous titration coordinates. Proteins 56, 738752.CrossRefGoogle ScholarPubMed
Lee, M. S. & Olson, M. A. (2008). Calculation of absolute ligand binding free energy to a ribosome-targeting protein as a function of solvent model. Journal of Physical Chemistry 112, 1341113417.CrossRefGoogle ScholarPubMed
Lensink, M. F., Mendez, R. & Wodak, S. J. (2007). Docking and scoring protein complexes: Capri 3rd edition. Proteins 74, 484495.Google Scholar
Li, X., Jacobson, M. P., Zhu, K., Zhao, S. & Friesner, R. A. (2007). Assignment of polar states for protein amino acid residues using an interaction cluster decomposition algorithm and its application to high resolution protein structure modeling. Proteins 66, 824837.CrossRefGoogle ScholarPubMed
Li, X-Z., Walker, B. & Michaelides, A. (2011). Quantum nature of the hydrogen bond. Proceedings of the National Academy of Sciences 108, 63696373.CrossRefGoogle Scholar
Linderstrom-Lang, K. (1924). On the ionization of proteins. Competes Rendus des Travaux du Laboratoire Carlsberg 15 (7), 129Google Scholar
Lowman, H. B., Cunningham, B. C. & Wells, J. A. (1991). Mutational analysis and protein engineering of receptor-binding determinants in human placental lactogen. Journal of Biological Chemistry 266, 1098210988.CrossRefGoogle ScholarPubMed
Maiti, A. & Drohat, A. C. (2011). Dependence of substrate binding and catalysis on pH, ionic strength, and temperature for thymine dna glycosylase: insights into recognition and processing of g.t mispairs. DNA Repair 10, 545–53.CrossRefGoogle ScholarPubMed
Maler, L., Blankenship, J., Rance, M. & Chazin, W. J. (2000). Site–site communication in the EF-hand Ca2+ – binding protein calbindin d9k. Nature Structure Molecular Biology 7, 245250.Google ScholarPubMed
Mason, A. C. & Jensen, J. H. (2008). Protein–protein binding is often associated with changes in protonation state. Proteins 71, 8191.CrossRefGoogle ScholarPubMed
Massova, I. & Kollman, P. A. (1999). Computational alanine scanning to probe protein–protein interactions: a novel approach to evaluate binding free energies. Journal of the American Chemical Society 121, 81338143.CrossRefGoogle Scholar
McCammon, J. A. (1998). Theory of biomolecular recognition. Current Opinion in Structural Biology 8, 245249.CrossRefGoogle ScholarPubMed
McCoy, A. J., Epa, V. C. & Colman, P. M. (1997). Electrostatic complementarity at protein/protein interfaces. Journal of Molecular Biology 268, 570584.CrossRefGoogle ScholarPubMed
McDonald, S. M., Willson, R. C. & McCammon, J. A. (1995). Determination of the pKa values of titratable groups of an antigen–antibody complex, HyHEL-5-hen egg lysozyme. Protein Engineering 8, 915924.CrossRefGoogle ScholarPubMed
McGuffee, S. R. & Elcock, A. H. (2006). Atomically detailed simulations of concentrated protein solutions: the effects of salt, pH, point mutations, and protein concentration in simulations of 1000-molecule systems. Journal of the American Chemical Society 128, 1209812110.CrossRefGoogle ScholarPubMed
McPhalen, C. A., Svendsen, I., Jonassen, I. & James, M. N. (1985). Crystal and molecular structure of chymotrypsin inhibitor 2 from barley seeds in complex with subtilisin novo. 82, 72427246.CrossRefGoogle Scholar
Mendez, R., Leplae, R., Lensink, M. & Wodak, S. (2005). Assessment of capri predictions in rounds 3–5 shows progress in docking procedures. Proteins 60, 150169.CrossRefGoogle ScholarPubMed
Milletti, F., Storchi, L. & Cruciani, G. (2009). Predicting protein pK(a) by environment similarity. Proteins 76, 484495.CrossRefGoogle ScholarPubMed
Milletti, F. & Vulpetti, A. (2010). Tautomer preference in PDB complexes and its impact on structure-based drug discovery. Journal of Chemical Information Modeling 50, 10621074.CrossRefGoogle ScholarPubMed
Mishra, S. H., Spring, A. M. & Germann, M. W. (2009). Thermodynamic profiling of HIV RREIIB RNAzinc finger interactions. Journal of Molecular Biology 393, 369382.CrossRefGoogle ScholarPubMed
Misra, V. K., Hecht, J. L., Yang, A-S. & Honig, B. (1998). Electrostatic contributions to the binding free energy of the »cI repressor to DNA. Biophysical Journal 75, 22622273.CrossRefGoogle Scholar
Mitra, R. C., Zhang, Z. & Alexov, E. (2011). In silico modeling of pH-optimum of protein–protein binding. Proteins 79, 925936.CrossRefGoogle ScholarPubMed
Mobley, D. L. & Dill, K. A. (2009). Binding of small-molecule ligands to proteins: what you see is not always what you get. Structure 17, 489498.CrossRefGoogle ScholarPubMed
Mongan, J., Case, D. A. & McCammon, J. A. (2004). Constant pH molecular dynamics in generalized born implicit solvent. Journal of. Computational Chemistry 25, 20382048.CrossRefGoogle ScholarPubMed
Murphy, K. P., Xie, D., Garcia, K. C., Amzel, L. M. & Freire, E. (1993). Structural energetics of peptide recognition: angiotensin II/antibody binding. Proteins 15, 113120.CrossRefGoogle ScholarPubMed
Myers, J., Grothaus, G., Narayanan, S. & Onufriev, A. (2006). A simple clustering algorithm can be accurate enough for use in calculations of pKs in macromolecules. Proteins 63, 928938.CrossRefGoogle ScholarPubMed
Nielsen, J. E., Gunner, M. R. & Garcia-Moreno, B. E. (2011). The pKa cooperative: a collaborative effort to advance structure-based calculations of pKa values and electrostatic effects in proteins. Proteins 79, 32493259.CrossRefGoogle ScholarPubMed
Nielsen, J. E. & McCammon, J. A. (2003). On the evaluation and optimization of protein X-ray structures for pKa calculations. Protein Science: a Publication of the Protein Society 12, 313326.CrossRefGoogle ScholarPubMed
Nielson, J. E. & Vriend, G. (2001). Optimizing the hydrogen-bond network in Poisson–Boltzmann equation-based pK a calculations. Proteins 43, 403412.CrossRefGoogle Scholar
Ofran, Y. & Rost, B. (2003). Predicted protein–protein interaction sites from local sequence information. FEBS Letters 544, 236239.CrossRefGoogle ScholarPubMed
Onufriev, A., Case, D. A. & Ullmann, G. M. (2001). A novel view of pH titration in biomolecules. Biochemistry 40, 34133419.CrossRefGoogle ScholarPubMed
Onufriev, A., Smondyrev, A. & Bashford, D. (2003). Proton affinity changes during unidirectional proton transport in the bacteriorhodopsin photocycle. Journal of Molecular Biology 332, 11831193.CrossRefGoogle ScholarPubMed
Onufriev, A. & Ullmann, G. M. (2004). Decomposing complex cooperative ligand binding into simple components: connections between microscopic and macroscopic models. Journal of Physical Chemistry B 108, 1115711169.CrossRefGoogle Scholar
Palpant, N. J., Houang, E. M., Delport, W., Hastings, K. E., Onufriev, A. V., Sham, Y. Y. & Metzger, J. M. (2010). Pathogenic peptide deviations support a model of adaptive evolution of chordate cardiac performance by troponin mutations. Physiological Genomics 42, 287299.CrossRefGoogle Scholar
Park, M-S., Gao, C. & Stern, H. A. (2011). Estimating binding affinities by docking/scoring methods using variable protonation states. Proteins 79, 304314.CrossRefGoogle ScholarPubMed
Perozzo, R., Folkers, G. & Scapozza, L. (2004). Thermodynamics of protein–ligand interactions: history, presence, and future aspects. Journal of Receptor and Signal Transduction Research 24, 152.CrossRefGoogle ScholarPubMed
Pierce, B. G., Hourai, Y. & Weng, Z. (2007). Accelerating protein docking in zdock using an advanced 3d convolution library. PLoS One 6, e24657.CrossRefGoogle Scholar
Qasim, M. A., Ranjbar, M. R., Wynn, R., Anderson, S. & Laskowski, M. (1995). Ionizable p1 residues in serine proteinase inhibitors undergo large pK shifts on complex formation. Journal of Biological Chemistry 270, 2741927422.CrossRefGoogle Scholar
Rapp, C. S., Schonbrun, C., Jacobson, M. P., Kalyanaraman, C. & Huang, N. (2009). Automated site preparation in physics-based rescoring of receptor ligand complexes. Proteins 77, 5261.CrossRefGoogle ScholarPubMed
Reeves Gibbs, M. D. R. A., Choudhary, P. R. & Bergquist, P. L. (2010). Alteration of the pH optimum of a family 11 xylanase, xynb6 of dictyoglomus thermophilum. New Biotechnology 27, 803–9.CrossRefGoogle Scholar
Rodinger, T., Howell, P. L. & Pomes, R. (2008). Calculation of absolute protein–ligand binding free energy using distributed replica sampling. Journal of Chemical Physics 129, 155102.CrossRefGoogle ScholarPubMed
Sackett, K., TerBush, A. & Weliky, D. P. (2011). Hiv gp41 six-helix bundle constructs induce rapid vesicle fusion at Ph 3.5 and little fusion at pH 7.0: understanding pH dependence of protein aggregation, membrane binding, and electrostatics, and implications for hiv-host cell fusion. European Biophysics Journal 40, 489502.CrossRefGoogle ScholarPubMed
Sakurai, K., Oobatake, M. & Goto, Y. (2001). Salt-dependent monomerdimer equilibrium of bovine b-lactoglobulin at pH 3. Protein Sciences 10, 23252335.CrossRefGoogle Scholar
Schneidman-Duhovny, D., Inbar, Y., Nussinov, R. & Wolfson, R. (2005). Geometry-based flexible and symmetric protein docking. Proteins 60, 224231.CrossRefGoogle ScholarPubMed
Schomburg, I., Chang, A., Ebeling, C., Gremse, M., Heldt, C., Huhn, G., & Schomburg, D. (2004). Brenda, the enzyme database: updates and major new developments. Nucleic Acids Research 32, D431–D43.CrossRefGoogle ScholarPubMed
Schreiber, G. & Fersht, A. R. (1993). Interaction of barnase with its popypeptide inhibitor barstar studied by protein engineering. Biochemistry 32, 51455150.CrossRefGoogle Scholar
Schreiber, G. & Fersht, A. R. (1995). Energetics of protein–protein interactions: analysis of the barnase–barstar interface by single mutations and double mutant cycles. Journal of Molecular Biology 248, 478486.CrossRefGoogle ScholarPubMed
Schueler-Furman, O., Wang, C. & Bake, D. (2005). Progress in protein–protein docking: atomic resolution in the predictions in the Capri experiment using rosattadock with an improved treatment of the side-chain flexibility. Proteins 60, 187194.CrossRefGoogle ScholarPubMed
Sham, Y. Y., Chu, Z. T. & Warshel, A. (1997). Consistent calculations of pKa pK a's of ionizable residues in proteins: semi-microscopic and microscopic approaches. Journal of Physical Chemistry 101, 44584472.CrossRefGoogle Scholar
Sheinerman, F. B. & Honig, B. (2002). On the role of electrostatic interactions in the design of protein–protein interfaces. Journal of Molecular Biology 318, 161177.CrossRefGoogle ScholarPubMed
Shin, C. J., Wong, S., Davis, M. J. & Ragan, M. A. (2009). Protein–protein interaction as a predictor of subcellular location. BMC Systems Biology 3, 28.CrossRefGoogle ScholarPubMed
Simonson, T., Carlsson, J. & Case, D. A. (2004). Proton binding to proteins: pKa calculations with explicit and implicit solvent models. Jorunal of American Chemistry Society 126, 41674180.CrossRefGoogle ScholarPubMed
Sims, P. A., Wong, C. F., Vuga, D., Mccammon, A. J. & Sefton, B. M. (2005). Relative contributions of desolvation, inter- and intramolecular interactions to binding affinity in protein kinase systems. Journal of Computational Chemistry 26, 668681.CrossRefGoogle ScholarPubMed
Singh, N. & Warshel, A. (2010). Absolute binding free energy calculations: on the accuracy of computational scoring of protein–ligand interactions. Proteins 78, 17051723.CrossRefGoogle ScholarPubMed
Smith, R., Brereton, I. M., Chai, R. Y. & Kent, S. B. (1996). Ionization states of the catalytic residues in HIV-1 protease. Nature Structural and Molecular Biology 3, 946950.CrossRefGoogle ScholarPubMed
Song, Y., Mao, J. & Gunner, M. R. (2009). MCCE2: Improving protein pKa calculations with extensive side chain rotamer sampling. Journal of Computational Chemistry 30, 22312247.CrossRefGoogle ScholarPubMed
Spassov, V. & Bashford, D. (1998). Electrostatic coupling to pH-titrating sites as a source of cooperativity in protein–ligand binding. Protein Science : a Publication of the Protein Society 7, 20122025.CrossRefGoogle ScholarPubMed
Spassov, V. Z. & Yan, L. (2008). A fast and accurate computational approach to protein ionization. Protein. Sciences 17, 19551970.CrossRefGoogle ScholarPubMed
Spitzer, R. & Jain, A. N. (2012). Surflex-dock: docking benchmarks and real-world application. Journal of Computer Aided Molcenlar Design 26 (6), 687699.CrossRefGoogle ScholarPubMed
Steuber, H., Czodrowski, P., Sotriffer, C. A. & Klebe, G. (2007). Tracing changes in protonation: a prerequisite to factorize thermodynamic data of inhibitor binding to aldose reductase. Journal of Molecular Biology 373, 13051320.CrossRefGoogle ScholarPubMed
Stranzl, G. R.Gruber, K., Steinkellner, G., Zangger, K., Schwab, H. & Kratky, C. (2004). Observation of a short, strong hydrogen bond in the active site of hydroxynitrile lyase from hevea brasiliensis explains a large pKa shift of the catalytic base induced by the reaction intermediate. Journal of Biological Chemistry 279, 36993707.CrossRefGoogle Scholar
Swails, J. M., Meng, Y., Walker, A. F., Marti, M. A., Estrin, D. A. & Roitberg, A. E. (2009). pH-dependent mechanism of nitric oxide release in nitrophorins 2 and 4. Journal of Physical Chemistry B 113, 11921201.CrossRefGoogle ScholarPubMed
Takahashi, T., Nakamura, H. & Walda, A. (1992). Electrostatic forces in two lysozymes: calculations and measurements of histidine pKa values. Biopolymers 32, 897909.CrossRefGoogle ScholarPubMed
Talley, K. & Alexov, E. (2010). On the pH-optimum of activity and stability of proteins. Proteins 78, 26992706.CrossRefGoogle ScholarPubMed
Tanford, C. (1970). Protein denaturation. c. theoretical models for the mechanism of denaturation. Advances in Protein Chemistry 24, 195.CrossRefGoogle ScholarPubMed
Tanford, C. & Kirkwood, J. (1957). Theory of protein titration curves. Journal of the American Chemical Society 79, 53335339.CrossRefGoogle Scholar
Tanford, C. & Roxby, R. (1972). Interpretation of protein titration curves. Biochemistry 11, 21922198.CrossRefGoogle ScholarPubMed
Taylor, R. D., Jewsbury, P. J., & Essex, J. W. (2002). A review of protein–small molecule docking methods. Journal of Computer-Aided Molecular Design 16, 151166.CrossRefGoogle ScholarPubMed
Ten Brink, T. & Exner, T. E. (2009). Influence of protonation, tautomeric, and stereoisomeric states on protein–ligand docking results. Journal of Chemical Information and Modelling 49, 15351546.CrossRefGoogle ScholarPubMed
Todorov, N. P., Monthoux, P. H. & Alberts, I. L. (2006). The influence of variations of ligand protonation and tautomerism on protein–ligand recognition and binding energy landscape. Journal of Chemical Information and Modelling 46, 11341142.CrossRefGoogle ScholarPubMed
Trott, O. & Olson, A. J. (2010). Autodock vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. Journal of Computational Chemistry 31, 455–61.CrossRefGoogle ScholarPubMed
Trylska, J., Antosiewicz, J., Geller, M., Hodge, C. N., Klabe, R. M., Head, M. S., & Gilson, M. K. (1999). Thermodynamic linkage between the binding of protons and inhibitors to HIV-1 protease. Protein Science 8, 180195.CrossRefGoogle ScholarPubMed
Tsai, C., Lin, S., Wolfson, H. & Nussinov, R. (1997). Studies of protein–protein interfaces: a statistical analysis of the hydrophobic effect. Protein Science 6, 5364.CrossRefGoogle Scholar
Ullmann, G. M. & Knapp, E-W. (1999). Electrostatic models for computing protonation and redox equilibria in proteins. European Biophysical Journal 28, 533551.CrossRefGoogle ScholarPubMed
Ullmann, R. T. & Ullmann, G. M. (2012). GMCT: a Monte Carlo simulation package for macromolecular receptors. Journal of Computational Chemistry, 33, 887900.CrossRefGoogle ScholarPubMed
Voet, D. & Voet, J. G. (1995). Biochemistry. John Wiley & Sons, 2 ed. New York.Google Scholar
Wang, J., Deng, Y. & Roux, B. (2006). Absolute binding free energy calculations using molecular dynamics simulations with restraining potentials. Biophysical Journal 91, 27982814.CrossRefGoogle ScholarPubMed
Wang, L., Witham, S., Zhang, Z., Li, L., Hodsdon, M. & Alexov, E. (2013). In silico investigation of pH-dependence of prolactin and human growth hormone binding to human prolactin receptor. Communications in Computation and Physics 13, 207222.CrossRefGoogle ScholarPubMed
Wang, R., Lu, Y. & Wang, S. (2003). Comparative evaluation of 11 scoring functions for molecular docking. Journal of Medicinal Chemistry 46, 22872303.CrossRefGoogle ScholarPubMed
Wang, Y. X., Freedberg, D. I., Yamazaki, T., Wingfield, P. T., Stahl, S. J., Kaufman, J. D., Kiso, Y. & Torchia, D. A. (1996). Solution NMR evidence that the HIV-1 protease catalytic aspartyl groups have different ionization states in the complex formed with the asymmetric drug KNI-272. Biochemistry 35, 99459950.CrossRefGoogle ScholarPubMed
Warren, G. L., Andrews, C. W., Capelli, A.-M., Clarke, B., LaLonde, J., Lambert, M. H., Lindvall, M., Nevins, N., Semus, S. F., Senger, S., Tedesco, G., Wall, I. D., Woolven, J. M., Peishoff, C. E., & Head, M. S. (2005). A critical assessment of docking programs and scoring functions. Journal of Medicinal Chemistry 49, 59125931.CrossRefGoogle Scholar
Wittayanarakul, K., Hannongbua, S. & Feig, M. (2008). Accurate prediction of protonation state as a prerequisite for reliable MM-PB(GB)SA binding free energy calculations of HIV-1 protease inhibitors. Journal of Computational Chemistry 29, 673685.CrossRefGoogle ScholarPubMed
Woo, H. J. (2008). Calculation of absolute protein–ligand binding constants with the molecular dynamics free energy perturbation method. Methods in Molecular Biology 443, 109120.CrossRefGoogle ScholarPubMed
Woo, H. J. & Roux, B. (2005). Calculation of absolute protein–ligand binding free energy from computer simulations. Proc Natl Acad Sci USA 102, 68256830.CrossRefGoogle ScholarPubMed
Xie, D., Gulnik, S., Collins, L., Gustchina, E., Suvorov, L., & Erickson, J. W. (1997). Dissection of the pH dependence of inhibitor binding energetics for an aspartic protease: direct measurement of the protonation states of the catalytic aspartic acid residues. Biochemistry, 36, 1616616172.CrossRefGoogle Scholar
Yamazaki, T., Nicholson, L. K., Wingfield, P., Stahl, S. J., Kaufman, J. D., Eyermann, C. J., Hodge, N. C., Lam, P. Y. & Torchia, D. A. (1994). NMR and X-ray evidence that the HIV protease catalytic aspartyl groups are protonated in the complex formed by the protease and a non-peptide cyclic urea-based inhibitor. Journal of the American Chemical Society 116, 1079110792.CrossRefGoogle Scholar
Yan, S. & Wu, G. (2011). Searching of predictors to predict pH optimum of cellulases. Applied Biochemistry and Biotechnology 165, 856–69.CrossRefGoogle ScholarPubMed
Yang, A. S. & Honig, B. (1993). On the pH dependence of protein stability. Journal of Molecular Biology 231, 459474.CrossRefGoogle Scholar
Yang, A-S., Gunner, M. R., Sampogna, R., Sharp, K. & Honig, B. (1993). On the calculation of pKas in proteins. Proteins: Structure, Function, and Genetics 15, 252265.CrossRefGoogle ScholarPubMed
Yang, A. S. & Honig, B. (1992). Electrostatic effects on protein stability. Current Opinion in Structural Biology 2, 4045.CrossRefGoogle Scholar
Yang, J. H., Park, J. Y., Kim, S. H. & Yoo, Y. J. (2008). Shifting pH optimum of bacillus circulans xylanase based on molecular modeling. Journal of Biotechnology 133, 294300.CrossRefGoogle ScholarPubMed
Zapata-Torres, G., Fierro, A., Miranda-Rojas, S., Guajardo, C., Salgado, J. C., Saez-Briones, P. & Celis-Barros, C. (2012). Influence of protonation on substrate and inhibitor interactions at the active site of human monoamine oxidase-a. Journal of Chemical Information and Modelling 52, 12131221.CrossRefGoogle ScholarPubMed
Zhou, H. X. & Gilson, M. K. (2009). Theory of free energy and entropy in noncovalent binding. Chemical Reviews, 109, 4092–107.CrossRefGoogle ScholarPubMed