Comparison of Entropic Contributions to Binding in a “Hydrophilic” versus “Hydrophobic” Ligand−Protein Interaction
Abstract

In the present study we characterize the thermodynamics of binding of histamine to recombinant histamine-binding protein (rRaHBP2), a member of the lipocalin family isolated from the brown-ear tick Rhipicephalus appendiculatus. The binding pocket of this protein contains a number of charged residues, consistent with histamine binding, and is thus a typical example of a “hydrophilic” binder. In contrast, a second member of the lipocalin family, the recombinant major urinary protein (rMUP), binds small hydrophobic ligands, with a similar overall entropy of binding in comparison with rRaHBP2. Having extensively studied ligand binding thermodynamics for rMUP previously, the data we obtained in the present study for HBP enables a comparison of the driving forces for binding between these classically distinct binding processes in terms of entropic contributions from ligand, protein, and solvent. In the case of rRaHBP2, we find favorable entropic contributions to binding from desolvation of the ligand; however, the overall entropy of binding is unfavorable due to a dominant unfavorable contribution arising from the loss of ligand degrees of freedom, together with the sequestration of solvent water molecules into the binding pocket in the complex. This contrasts with binding in rMUP where desolvation of the protein binding pocket makes a minor contribution to the overall entropy of binding given that the pocket is substantially desolvated prior to binding.
Introduction


Materials and Methods
Overexpression and Purification of rRaHBP2 and 13C,15N (>97%)-Enriched rRaHBP2
NMR Measurements
NMR Resonance Assignments
NMR Relaxation Measurements
Analysis of NMR Relaxation Data
X-ray Crystallography
Crystallization and Data Collection
rRaHBP2(D24R) | rRaHBP2(D24R)-histamine | |
---|---|---|
wavelength (Å) | 1.54 | 0.95 |
resolution range (Å) | 12−2.25 (2.37−2.25) | 55−1.55 (1.63−1.55) |
unique reflections | 18800 (2735) | 65264 (9456) |
completeness (%) | 99.3 (100.0) | 97.2 (96.9) |
multiplicity | 4.0 (4.0) | 2.9 (3) |
Rsymi | 0.09 (0.30) | 0.059 (0.54) |
space group | P21 | P212121 |
no. of molecules per asymmetric unit | 2 | 2 |
unit cell dimensions (nm) | a = 57.96; b = 56.61; c = 63.80; α = γ = 90; β = 107.07 | a = 75.22; b = 78.73; c = 77.65; α = β = γ = 90° |
Rwork (Rfree) (%) | 22.8 (27.6) | 17.2 (21.4) |
bond length (Å) | 0.012 | 0.011 |
angles (deg) | 1.279 | 1.305 |
molprobity clash score | 11.9 | 5.07 |
Ramachandran favored (%) | 97.0 | 99.1 |
Values in parentheses are for highest resolution shell. Rsym = ∑hkl∑i(Ii(hkl) − Imean(hkl))/∑hkl∑i(Ii(hkl)).
Structure Determination and Refinement
ITC Experiments
Molecular Dynamics Simulations

Solvation Thermodynamics Calculations

Results and Discussion
Crystal Structure of rRaHBP2(D24R)-Histamine Complex
Figure 1

Figure 1. Stereo view of the binding site of rRaHBP2(D24R) with bound histamine. This site corresponds to the H site of the wild-type protein. Binding-site residues are colored blue, the ligand is colored red, and ordered water molecules are shown as green spheres. Figure prepared using MOLMOL (36)
ITC Measurements
Figure 2

Figure 2. Typical ITC isotherms for the binding of histamine to (above) rRaHBP2(D24R) and (below) wild-type rRaHBP2 at 298 K.
temp (K) | ΔH° (kJ/mol) | TΔS° (kJ/mol) | ΔG° (kJ/mol) | Kd (nM) | |
---|---|---|---|---|---|
rRaHBP-2(D24R) | |||||
278 | −42.7a ± 0.9b | 6.5 ± 1 | −49.2 ± 0.6 | 2.4 ± 0.6 | |
288 | −48.3 ± 3.1 | 0.4 ± 4.5 | −48.8 ± 1.1 | 2.9 ± 1.3 | |
298 | −58.3 ± 1.2 | −9.3 ± 2.5 | −49.1 ± 1.4 | 2.5 ± 1.4 | |
rRaHBP-2 (wt) | |||||
H site | 298 | −70.9 ± 7.1 | −24.4 ± 7.3 | −46.5 ± 1.3 | 7.0 ± 3.6 |
L site | 298 | −53.2 ± 3.3 | −12.0 ± 3.6 | −41.2 ± 1.5 | 59.6 ± 34.8 |
Values are expressed as the mean of three measurements.
Standard errors were determined from duplicate experiments by error propagation.
Entropic Contribution from the Protein Backbone
Figure 3

Figure 3. Stereo images of the rRaHBP2(D24R)−histamine complex, showing positive (green) and negative (blue) contributions or no contribution (within the standard error, red) to the overall binding entropy from amide bond vectors on binding histamine (blue and yellow spheres).
Entropic Contribution from Protein Side Chains
ΤΔΔS (complex-apo) kJ/mol | |
---|---|
backbone (N−H) | +16.4 ± 1.0 |
side chain (C−C) | +17.4 ± 1.8 |
side chain (C−C)a | +12.7 ± 0.16 |
Contribution from binding-pocket residue side chains only (namely, D110, F108, Y100, E82, W42, D39, E135, V41, Y36, N130).
Entropic Contributions to Binding from Other Sources

ligand | ΔH° (kJ/mol) | TΔS° (kJ/mol) | ΔG° (kJ/mol) |
---|---|---|---|
histamine | −137.2 | −67.1 | −70.1 |
2-methyl imidazole | −96.3 | −53.3 | −43.0 |
(experimental)a | −42.9 | ||
n-propylamine | −65.5 | −49.0 | −16.5 |
(experimental)a | −55.8 | −37.4 | −18.4 |
Values taken from ref 51.
Decomposition of the Overall Entropy of Binding
Description | MUP-IPMPa | rRaHBP2(D24R)−histamine |
---|---|---|
TΔS°i | ||
protein DOF | −0.8 ± 3.8 | +29.8 ± 9.9 |
ligand DOF | ca. −37 | ca. −59 |
−TΔS°solvL | ||
ligand desolvation | +26.7 ± 8.4 | +67.1 |
TΔS°solvPL − TΔS°solvP | ||
desolvation of protein/complex | +0.4 ± 9.2 | −47 ± 10.2 |
TΔS°b | ||
observed entropy | −10.7 ± 0.5 | −9.3 ± 2.5 |
Data taken from ref 4.
Driving Forces for Ligand Binding in rRaHBP2(D24R) Compared with rMUP
Figure 4

Figure 4. (Left) Typical mean square displacements of solvent water molecules in the binding pocket of uncomplexed rRaHBP2(D24R) (red, green, and blue traces) in comparison with bulk water (black trace). (Right) Typical mean square displacements of solvent water molecules in the binding pocket of rRaHBP2(D24R) in complex with histamine (red, green, blue, and gray traces). In both panels the inset shows rotational autocorrelation functions for these waters using the same color scheme.
Supporting Information
Complete refs 20 and 24, backbone resonance assignments, R1, R2 and ssNOE values, 15N amide model-free parameters and TΔSconf values for backbone amide groups for rRaHBP2(D24R) and rRaHBP2(D24R) in complex with histamine. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgment
This work was supported by BBSRC, grant number BB/E000991/1, and by The Wellcome Trust, grant number 075520. We thank Dr. Gary Thompson for assistance with RELAX calculations.
References
This article references 59 other publications.
- 1Bingham, R., Bodenhausen, G., Findlay, J. H. B. C., Hsieh, S.-Y., Kalverda, A. P., Kjellberg, A., Perazzolo, C., Phillips, S. E. V., Seshadri, K., Turnbull, W. B., and Homans, S. W. J. Am. Chem. Soc. 2004, 126, 1675– 1681
- 2Barratt, E., Bingham, R., Warner, D. J., Laughton, C. A., Phillips, S. E. V., and Homans, S. W. J. Am. Chem. Soc. 2005, 127, 11827– 11834
- 3Malham, R., Johnstone, S., Bingham, R. J., Barratt, E., Phillips, S. E., Laughton, C. A., and Homans, S. W. J. Am. Chem. Soc. 2005, 127, 17061– 7
- 4Shimokhina, N., Bronowska, A., and Homans, S. W. Angew. Chem., Int. Ed. 2006, 45, 6374– 6376
- 5Delaglio, F., Grzesiek, S., Vuister, G. W., Zhu, G., Pfeifer, J., and Bax, A. J. Biomol. NMR 1995, 6, 277– 293[Crossref], [PubMed], [CAS], Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXhtVSmurfK&md5=a670fca5b164083e2178fafd2fb951ffNMRPipe: a multidimensional spectral processing system based on UNIX pipesDelaglio, Frank; Grzesiek, Stephan; Vuister, Geerten W.; Zhu, Guang; Pfeifer, John; Bax, AdJournal of Biomolecular NMR (1995), 6 (3), 277-93CODEN: JBNME9; ISSN:0925-2738. (ESCOM)The NMRPipe system is a UNIX software environment of processing, graphics, and anal. tools designed to meet current routine and research-oriented multidimensional processing requirements, and to anticipate and accommodate future demands and developments. The system is based on UNIX pipes, which allow programs running simultaneously to exchange streams of data under user control. In an NMRPipe processing scheme, a stream of spectral data flows through a pipeline of processing programs, each of which performs one component of the overall scheme, such as Fourier transformation or linear prediction. Complete multidimensional processing schemes are constructed as simple UNIX shell scripts. The processing modules themselves maintain and exploit accurate records of data sizes, detection modes, and calibration information in all dimensions, so that schemes can be constructed without the need to explicitly define or anticipate data sizes or storage details of real and imaginary channels during processing. The asynchronous pipeline scheme provides other substantial advantages, including high flexibility, favorable processing speeds, choice of both all-in-memory and disk-bound processing, easy adaptation to different data formats, simpler software development and maintenance, and the ability to distribute processing tasks on multi-CPU computers and computer networks.
- 6Vranken, W. F., Boucher, W., Stevens, T. J., Fogh, R. H., Pajon, A., Llinas, M., Ulrich, E. L., Markley, J. L., Ionides, J., and Laue, E. D. Proteins 2005, 59, 687– 696
- 7Farrow, N. A., Muhandiram, R., Singer, A. U., Pascal, S. M., Kay, C. M., Gish, G., Shoelson, S. E., Pawson, T., Forman-Kay, J. D., and Kay, L. E. Biochemistry (Moscow) 1994, 33, 5984– 6003[ACS Full Text
], [CAS], Google Scholar
7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXktVersbk%253D&md5=57fb1299b4acd2eb8788a83f1c322257Backbone Dynamics of a Free and a Phosphopeptide-Complexed Src Homology 2 Domain Studied by 15N NMR RelaxationFarrow, Neil A.; Muhandiram, Ranjith; Singer, Alex U.; Pascal, Steven M.; Kay, Cyril M.; Gish, Gerry; Shoelson, Steven E.; Pawson, Tony; Forman-Kay, Julie D.; Kay, Lewis E.Biochemistry (1994), 33 (19), 5984-6003CODEN: BICHAW; ISSN:0006-2960.The backbone dynamics of the C-terminal Src Homol. 2 (SH2) domain of phosphatidylinositol phospholipase Cγ1 were investigated. Two forms of the domain were studied, one in complex with a high-affinity binding peptide derived from the platelet-derived growth factor receptor and the other in the absence of this peptide. Two-dimensional 1H-15N NMR methods, employing pulsed field gradients, were used to det. steady-state 1H-15N NOE values and T1 and T2 15N relaxation times. Backbone dynamics were characterized by the overall correlation time (τm), order parameters (S2), effective correlation times for internal motions (τe), and, if required, terms to account for motions on a microsecond-to-millisecond-time scale. An extended 2-time-scale formalism was used for residues having relaxation data that could not be fit adequately using a single-time-scale formalism. The overall correlation times of the uncomplexed and complexed forms of SH2 were found to be 9.2 and 6.5 ns, resp., suggesting that the uncomplexed form was in a monomer-dimer equil. This was subsequently confirmed by hydrodynamic measurements. Anal. of order parameters revealed that residues in the so-called phosphotyrosine-binding loop exhibited higher than av. disorder in both forms of SH2. Although localized differences in order parameters were obsd. between the uncomplexed and complexed forms of SH2, overall, higher order parameters were not found in the peptide-bound form, indicating that on av., picosecond-time-scale disorder is not reduced upon binding peptide. The relaxation data of the SH2-phosphopeptide complex were fit with fewer exchange terms than the uncomplexed form. This may reflect the monomer-dimer equil. that exists in the uncomplexed form or may indicate that the complexed form has lower conformational flexibility on a microsecond-to-millisecond-time scale. - 8Johnson, B. A. and Blevins, R. A. J. Biomol. NMR 1994, 4, 603– 614[Crossref], [PubMed], [CAS], Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmt1Gkurw%253D&md5=6a93a0d9704fbbb654abc192203c68aeNMRView: a computer program for the visualization and analysis of NMR dataJohnson, Bruce A.; Blevins, Richard A.Journal of Biomolecular NMR (1994), 4 (5), 603-14CODEN: JBNME9; ISSN:0925-2738.NMRView is a computer program designed for the visualization and anal. of NMR data. It allows the user to interact with a practically unlimited no. of 2D, 3D and 4D NMR data files. Any no. of spectral windows can be displayed on the screen in any size and location. Automatic peak picking and facilitated peak anal. features are included to aid in the assignment of complex NMR spectra. NMR View provides structure anal. features and data transfer to and from structural generation programs, allowing for a tight coupling between the spectral anal. and structural generation.,. Visual correlation between structures and spectra can be done with the Mol. Data Viewer, a mol. graphics program with bidirectional communication to NMR View. The used interface can be customized and a command language is provided to allow for the automation of various tasks.
- 9García de la Torre, J., Huertas, M. L., and Carrasco, B. J. Magn. Reson. 2000, 147, 138– 146
- 10d’Auvergne, E. J. and Gooley, P. R. J. Biomol. NMR 2008, 40, 107– 119
- 11d’Auvergne, E. J. and Gooley, P. R. J. Biomol. NMR 2008, 40, 121– 133
- 12Yang, D. and Kay, L. E. J. Mol. Biol. 1996, 263, 369– 82
- 13Leslie, A. G. W. In CCP4 & ESFEACMB Newsletter On Protein Crystallography; Daresbury Laboratory: Warington, 1992.Google ScholarThere is no corresponding record for this reference.
- 14CCP4. Acta Crystallogr. 1994, D50, 760− 763Google ScholarThere is no corresponding record for this reference.
- 15Murshudov, G. N., Vagin, A. A., and Dodson, E. J. Acta Crystallogr. 1997, D53, 240– 255Google ScholarThere is no corresponding record for this reference.
- 16Emsley, P. and Cowtan, K. Acta Crystallogr. 2004, D60, 2126– 2132Google ScholarThere is no corresponding record for this reference.
- 17Laskowski, R. A., MacArthur, M. W., Moss, D. S., and Thornton, J. M. J. Appl. Crystallogr. 1993, 26, 283– 291[Crossref], [CAS], Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXit12lurY%253D&md5=fb69fc4410cd716aaaa7cc0db06b3ed2PROCHECK: a program to check the stereochemical quality of protein structuresLaskowski, Roman A.; MacArthur, Malcolm W.; Moss, David S.; Thornton, Janet M.Journal of Applied Crystallography (1993), 26 (2), 283-91CODEN: JACGAR; ISSN:0021-8898.The PROCHECK suite of programs provides a detailed check on the stereochem. of a protein structure. Its outputs comprise a no. of plots in PostScript format and a comprehensive residue-by-residue listing. These give an assessment of the overall quality of the structure as compared with well-refined structures of the same resoln. and also highlight regions that may need further investigation. The PROCHECK programs are useful for assessing the quality not only of protein structures in the process of being solved but also of existing structures and of those being modeled on known structures.
- 18Davis, I. W., Leaver-Fay, A., Chen, V. B., Block, J. N., Kapral, G. J., Wang, X., Murray, L. W., Arendall, I., W. B., Snoeyink, J., Richardson, J. S., and Richardson, D. C. Nucleic Acids Res. 2007, 35W375−W383
- 19Navaza, J. Acta Crystallogr., Sect. A 1994, 50, 157– 163
- 20Case, D. A. et al. AMBER 8; University of California: San Francisco, 2004.Google ScholarThere is no corresponding record for this reference.
- 21Cornell, W. D., Cieplak, P., Bayly, C. I., Gould, I. R., Merz, K. M., Ferguson, D. M., Spellmeyer, D. C., Fox, T., Caldwell, J. W., and Kollman, P. A. J. Am. Chem. Soc. 1995, 117, 5179– 5197[ACS Full Text
], [CAS], Google Scholar
21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXlsFertrc%253D&md5=2c8901ed614d543db51d3a28f5b6053cA Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic MoleculesCornell, Wendy D.; Cieplak, Piotr; Bayly, Christopher I.; Gould, Ian R.; Merz, Kenneth M., Jr.; Ferguson, David M.; Spellmeyer, David C.; Fox, Thomas; Caldwell, James W.; Kollman, Peter A.Journal of the American Chemical Society (1995), 117 (19), 5179-97CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The authors present the derivation of a new mol. mech. force field for simulating the structures, conformational energies, and interaction energies of proteins, nucleic acids, and many related org. mols. in condensed phases. This effective two-body force field is the successor to the Weiner et al. force field and was developed with some of the same philosophies, such as the use of a simple diagonal potential function and electrostatic potential fit atom centered charges. The need for a 10-12 function for representing hydrogen bonds is no longer necessary due to the improved performance of the new charge model and new van der Waals parameters. These new charges are detd. using a 6-31G* basis set and restrained electrostatic potential (RESP) fitting and have been shown to reproduce interaction energies, free energies of solvation, and conformational energies of simple small mols. to a good degree of accuracy. Furthermore, the new RESP charges exhibit less variability as a function of the mol. conformation used in the charge detn. The new van der Waals parameters have been derived from liq. simulations and include hydrogen parameters which take into account the effects of any geminal electroneg. atoms. The bonded parameters developed by Weiner et al. were modified as necessary to reproduce exptl. vibrational frequencies and structures. Most of the simple dihedral parameters have been retained from Weiner et al., but a complex set of .vphi. and ψ parameters which do a good job of reproducing the energies of the low-energy conformations of glycyl and alanyl dipeptides was developed for the peptide backbone. - 22Hornak, V., Abel, R., Okur, A., Strockbine, B, Roitberg, A., and Simmerling, C. Proteins: Struct., Funct., Bioinf. 2006, 65, 712– 725
- 23Wang, J., Wolf, R. M., Caldwell, J., Kollman, P., and Case, D. A. J. Comput. Chem. 2004, 25, 1157– 1174[Crossref], [PubMed], [CAS], Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksFakurc%253D&md5=2992017a8cf51f89290ae2562403b115Development and testing of a general Amber force fieldWang, Junmei; Wolf, Romain M.; Caldwell, James W.; Kollman, Peter A.; Case, David A.Journal of Computational Chemistry (2004), 25 (9), 1157-1174CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)We describe here a general Amber force field (GAFF) for org. mols. GAFF is designed to be compatible with existing Amber force fields for proteins and nucleic acids, and has parameters for most org. and pharmaceutical mols. that are composed of H, C, N, O, S, P, and halogens. It uses a simple functional form and a limited no. of atom types, but incorporates both empirical and heuristic models to est. force consts. and partial at. charges. The performance of GAFF in test cases is encouraging. In test I, 74 crystallog. structures were compared to GAFF minimized structures, with a root-mean-square displacement of 0.26 Å, which is comparable to that of the Tripos 5.2 force field (0.25 Å) and better than those of MMFF 94 and CHARMm (0.47 and 0.44 Å, resp.). In test II, gas phase minimizations were performed on 22 nucleic acid base pairs, and the minimized structures and intermol. energies were compared to MP2/6-31G* results. The RMS of displacements and relative energies were 0.25 Å and 1.2 kcal/mol, resp. These data are comparable to results from Parm99/RESP (0.16 Å and 1.18 kcal/mol, resp.), which were parameterized to these base pairs. Test III looked at the relative energies of 71 conformational pairs that were used in development of the Parm99 force field. The RMS error in relative energies (compared to expt.) is about 0.5 kcal/mol. GAFF can be applied to wide range of mols. in an automatic fashion, making it suitable for rational drug design and database searching.
- 24Frisch, M. J. et al. Gaussian 98 (Revision A.9); Gaussian, Inc.: Pittsburgh, PA, 1998.Google ScholarThere is no corresponding record for this reference.
- 25Cornell, W. D., Cieplak, P., Bayly, C. I., and Kollman, P. A. J. Am. Chem. Soc. 1993, 115, 9620– 9631[ACS Full Text
], [CAS], Google Scholar
25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXmt1Gmsbk%253D&md5=448e45857da2a2fcdbb19e127e7b48a1Application of RESP charges to calculate conformational energies, hydrogen bond energies, and free energies of solvationCornell, Wendy D.; Cieplak, Piotr; Bayly, Christopher I.; Kollmann, Peter A.Journal of the American Chemical Society (1993), 115 (21), 9620-31CODEN: JACSAT; ISSN:0002-7863.The authors apply a new restrained electrostatic potential fit charge model (two-stage RESP) to conformational anal. and the calcn. of intermol. interactions. Specifically, the authors study conformational energies in butane, Me Et thioether, three simple alcs., three simple amines, and 1,2-ethanediol as a function of charge model (two-stage RESP vs std. ESP) and 1-4 electrostatic scale factor. The authors demonstrate that the two-stage RESP model with a 1-4 electrostatic scale factor ∼1/1.2 is a very good model, as evaluated by comparison with high-level ab initio calcns. For methanol and N-methylacetamide interactions with TIP3P water, the two-stage RESP model leads to hydrogen bonds only slightly weaker than found with the std. ESP changes. In tests on DNA base pairs, the two-stage RESP model leads to hydrogen bonds which are ∼1 kcal/mol weaker than those calcd. with the std. ESP charges but closer in magnitude to the best currently available ab initio calcns. Furthermore, the two-stage RESP charges, unlike the std. ESP charges, reproduce the result that Hoogsteen hydrogen bonding is stronger than Watson-Crick hydrogen bonding for adenine-thymine base pairs. The free energies of solvation of both methanol and trans-N-methylacetamide were also calcd. for the std. ESP and two-stage RESP models and both were in good agreement with expt. The authors have combined the use of two-stage RESP charges with multiple conformational fitting in studies of conformationally dependent dipole moments and energies of propylamine. The authors find that the combination of these approaches is synergistic in leading to useful charge distributions for mol. simulations. Two-stage RESP charges thus reproduce both intermol. and intramol. energies and structure quite well, making this charge model a crit. advancement in the development of a general force field for modeling biol. macromols. and their ligands, both in the gas phase and in soln. - 26Darden, T., York, D., and Pedersen, L. J. Chem. Phys. 1993, 98, 10089– 10092[Crossref], [CAS], Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXks1Ohsr0%253D&md5=3c9f230bd01b7b714fd096d4d2e755f6Particle mesh Ewald: an N·log(N) method for Ewald sums in large systemsDarden, Tom; York, Darrin; Pedersen, LeeJournal of Chemical Physics (1993), 98 (12), 10089-92CODEN: JCPSA6; ISSN:0021-9606.An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolution using fast Fourier transforms. Timings and accuracies are presented for three large cryst. ionic systems.
- 27Lipari, G. and Szabo, A. J. Am. Chem. Soc. 1982, 104, 4546– 4559[ACS Full Text
], [CAS], Google Scholar
27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XltVejtr8%253D&md5=7188e058f88fbaf1ba066d08a77e13a7Model-free approach to the interpretation of nuclear magnetic resonance relaxation in macromolecules. 1. Theory and range of validityLipari, Giovanni; Szabo, AttilaJournal of the American Chemical Society (1982), 104 (17), 4546-59CODEN: JACSAT; ISSN:0002-7863.A new approach to the interpretation of NMR relaxation expts. on macromols. in soln. is presented. Theor. foundations are dealt with and the range of validity of the approach is established. For both isotropic and anisotropic overall motion, the unique information on fast internal motions contained in relaxation expts. can be completely specified by 2 model-independent quantities: (1) a generalized order parameter, .SCRIPTS., which is a measure of the spatial restriction of the motion, and (2) an effective correlation time, τe, which is a measure of the rate of motion. A simple expression for the spectral d. involving these 2 parameters is derived and is exact when the internal (but not overall) motions are in the extreme narrowing limit. The model-free approach consists of using the above spectral d. to det. least-squares-fit relaxation data by treating .SCRIPTS.2 and τe as adjustable parameters. The range of validity of this approach is illustrated by analyzing error-free relaxation data generated by using sophisticated dynamic models. Empirical rules are presented that allow one to est. the accuracy of .SCRIPTS.2 and τe extd. by using the model-free approach by considering their numerical values, the resonance frequencies, and the parameters for the overall motion. For fast internal motions, it is unnecessary to use approaches based on complicated spectral densities derived within the framework of a model because all models that can give the correct value of .SCRIPTS.2 work equally well. The unique dynamic information (.SCRIPTS. and τe) can be easily extd. by using the model-free approach. If a phys. picture of the motion is desired, the numerical values of .SCRIPTS.2 and τe can be readily interpreted within a phys. reasonable model. - 28Chatfield, D. C., Szabo, A., and Brooks, B. R. J. Am. Chem. Soc. 1998, 120, 5301– 5311
- 29MacRaild, C. A., Daranas, A. H., Bronowska, A., and Homans, S. W. J. Mol. Biol. 2007, 368, 822– 832
- 30Stockmann, H., Bronowska, A., Syme, N. R., Thompson, G. S., Kalverda, A. P., Warriner, S. L., and Homans, S. W. J. Am. Chem. Soc. 2008, 130, 12420– 12426
- 31Best, R. B., Clarke, J., and Karplus, M. J. Mol. Biol. 2005, 349, 185– 203
- 32Schlitter, J. Chem. Phys. Lett. 1993, 215, 617– 621
- 33Klamt, A. and Schuurmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 5, 799– 805
- 34Smith, D. E. and Haymet, A. D. J. J. Chem. Phys. 1993, 98, 6445– 6454
- 35Paesen, G. C., Adams, P. L., Harlos, K., Nuttall, P. A., and Stuart, D. I. Mol. Cell 1999, 3, 661– 671
- 36Koradi, R., Billeter, M., and Wuthrich, K. J. Mol. Graphics 1996, 14, 51[Crossref], [PubMed], [CAS], Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xis12nsrc%253D&md5=dab669ff481b19cde8fbe2e4234711faMOLMOL: a program for display and analysis of macromolecular structuresKoradi, Reto; Billeter, Martin; Wuethrich, KurtJournal of Molecular Graphics (1996), 14 (1), 51-5, plates, 29-32CODEN: JMGRDV; ISSN:0263-7855. (Elsevier)MOLMOL is a mol. graphics program for display, anal., and manipulation of three-dimensional structures of biol. macromols., with special emphasis on NMR soln. structures of proteins and nucleic acids. MOLMOL has a graphical user interface with menus, dialog boxes, and online help. The display possibilities include conventional presentation, as well as novel schematic drawings, with the option of combining different presentations in one view of a mol. Covalent mol. structures can be modified by addn. or removal of individual atoms and bonds, and three-dimensional structures can be manipulated by interactive rotation about individual bonds. Special efforts were made to allow for appropriate display and anal. of the sets of typically 20-40 conformers that are conventionally used to represent the result of an NMR structure detn., using functions for superimposing sets of conformers, calcn. of root mean square distance (RMSD) values, identification of hydrogen bonds, checking and displaying violations of NMR constraints, and identification and listing of short distances between pairs of hydrogen atoms.
- 37Goldberg, R. N., Kishore, N., and Lennen, R. M. J. Phys. Chem. Ref. Data 2002, 31, 231– 370
- 38Ikura, M., Kay, L. E., and Bax, A. Biochemistry (Moscow) 1990, 29, 4659– 4667[ACS Full Text
], [CAS], Google Scholar
38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXitFegs7g%253D&md5=cc22f1867438dfbca353275b77773b55A novel approach for sequential assignment of proton, carbon-13, and nitrogen-15 spectra of larger proteins: heteronuclear triple-resonance three-dimensional NMR spectroscopy. Application to calmodulinIkura, Mitsuhiko; Kay, Lewis E.; Bax, AdBiochemistry (1990), 29 (19), 4659-67CODEN: BICHAW; ISSN:0006-2960.The approach is demonstrated for the protein calmodulin (16.7 kilodaltons), uniformly (∼95%) labeled with 15N and 13C. Sequential assignment of the backbone residues by std. methods was not possible because of the very narrow chem. shift distribution range of both NH and CαH protons in this largely α-helical protein. The combined use of 4 new types of heteronuclear 3-dimensional (3D) NMR spectra together with the previously described homonuclear Hartmann-Hahn-heteronuclear multiple quantum correlation 3D expt. (Marion, D. et al., 1989) can provide unambiguous sequential assignment of protein backbone resonances. Sequential connectivity is derived from 1-bond J couplings and the procedure is independent of backbone conformation. All the new 3D NMR expts. use 1H detection and rely on multiple-step magnetization transfers via well-resolved 1-bond J couplings, offering high sensitivity and requiring a total of only 9 days for the recording of all 5 3D spectra. Because the combination of 3D spectra offers at least 2 and often 3 independent pathways for detg. sequential connectivity, the new assignment procedure is easily automated. Complete assignments are reported for the 1H, C, and N backbone resonances of calmodulin, complexed with Ca. - 39Sattler, M., Schleucher, J., and Griesinger, C. Prog. Nucl. Magn. Reson. Spectrosc. 1999, 34, 93– 158
- 40Muhandiram, D. R., Yamazaki, T., Sykes, B. D., and Kay, L. E. J. Am. Chem. Soc. 1995, 117, 11536– 11544
- 41Millet, O., Muhandiram, D. R., Skrynnikov, N. R., and Kay, L. E. J. Am. Chem. Soc. 2002, 124, 6439– 6448
- 42Yu, L., Zhu, C. X., Tse-Dinh, Y. C., and Fesik, S. W. Biochemistry (Moscow) 1996, 35, 9661– 6[ACS Full Text
], [CAS], Google Scholar
42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XktVCqtLc%253D&md5=df27fe6eeb2f7baccb073af6c10d4ccbBackbone dynamics of the C-terminal domain of Escherichia coli topoisomerase I in the absence and presence of single-stranded DNAYu, Liping; Zhu, Chang-Xi; Tse-Dinh, Yuk-Ching; Fesik, Stephen W.Biochemistry (1996), 35 (30), 9661-9666CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)The backbone dynamics of the C-terminal DNA-binding domain of Escherichia coli topoisomerase I was characterized in the absence and presence of single-stranded DNA by NMR spectroscopy. 15N spin-lattice relaxation times (T1), spin-spin relaxation times (T2), and heteronuclear NOEs were detd. for the uniformly 15N-labeled protein. These data were analyzed by using the model-free formalism to derive the model-free parameters (S2, τe, and Rex) for each backbone N-H bond vector and the overall mol. rotational correlation time (τm). The mol. rotational correlation time τm was detd. to be 7.49 ns for the free and 12.7 ns for the complexed protein. Several residues were found to be much more mobile than the av., including 11 residues at the N-terminus, 2 residues at the C-terminus, and residues 25 and 31-35 which are located in a region of the protein that binds to DNA. The binding of ssDNA to the free protein caused an increase in the order parameters (S2) for a small no. of residues and a decrease in the order parameters (S2) for the majority of the residues. In particular, upon binding to ssDNA, the mobility of the 1st α-helix and the 2 β-sheets was increased, and the mobility of a few specific residues in the loops/turns was restricted. These results differed from previous studies on the backbone dynamics of mol. complexes in which reduced mobilities were typically obsd. upon ligand binding. - 43Arumugam, S., Gao, G., Patton, B. L., Semenchenko, V., Brew, K., and Van Doren, S. R. J. Mol. Biol. 2003, 327, 719– 34
- 44Yun, S., Jang, D. S., Kim, D. H., Choi, K. Y., and Lee, H. C. Biochemistry (Moscow) 2001, 40, 3967– 73[ACS Full Text
], [CAS], Google Scholar
44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXhsFKgs7s%253D&md5=89add77eb20e5813a81900488e26de1615N NMR Relaxation Studies of Backbone Dynamics in Free and Steroid-Bound Δ5-3-Ketosteroid Isomerase from Pseudomonas testosteroniYun, Sunggoo; Jang, Do Soo; Kim, Do-Hyung; Choi, Kwan Yong; Lee, Hee CheonBiochemistry (2001), 40 (13), 3967-3973CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)The backbone dynamics of Δ5-3-ketosteroid isomerase (KSI) from Pseudomonas testosteroni has been studied in free enzyme and its complex with a steroid ligand, 19-nortestosterone hemisuccinate (19-NTHS), by 15N relaxation measurements. The relaxation data were analyzed using the model-free formalism to ext. the model-free parameters (S2, τe, and Rex) and the overall rotational correlation time (τm). The rotational correlation times were 19.23 ± 0.08 and 17.08 ± 0.07 ns with the diffusion anisotropies (D‖/D.perp.) of 1.26 ± 0.03 and 1.25 ± 0.03 for the free and steroid-bound KSI, resp. The binding of 19-NTHS to free KSI causes a slight increase in the order parameters (S2) for a no. of residues, which are located mainly in helix A1 and strand B4. However, the majority of the residues exhibit reduced order parameters upon ligand binding. In particular, strands B3, B5, and B6, which have most of the residues involved in the dimer interaction, have the reduced order parameters in the steroid-bound KSI, indicating the increased high-frequency (pico- to nanosecond) motions in the intersubunit region of this homodimeric enzyme. Our results differ from those of previous studies on the backbone dynamics of monomeric proteins, in which high-frequency internal motions are typically restricted upon ligand binding. - 45Shapiro, Y. E., Kahana, E., Tugarinov, V., Liang, Z., Freed, J. H., and Meirovitch, E. Biochemistry (Moscow) 2002, 41, 6271– 81[ACS Full Text
], [CAS], Google Scholar
45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XjtVKhurg%253D&md5=cc1c988a7898412a04ac350f977d4ac4Domain Flexibility in Ligand-Free and Inhibitor-Bound Escherichia coli Adenylate Kinase Based on a Mode-Coupling Analysis of 15N Spin RelaxationShapiro, Yury E.; Kahana, Edith; Tugarinov, Vitali; Liang, Zhichun; Freed, Jack H.; Meirovitch, EvaBiochemistry (2002), 41 (20), 6271-6281CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Adenylate kinase from Escherichia coli (AKeco), consisting of a 23.6-kDa polypeptide chain folded into domains CORE, AMPbd, and LID catalyzes the reaction AMP + ATP ↔ 2ADP. The domains AMPbd and LID execute large-amplitude movements during catalysis. Backbone dynamics of ligand-free and AP5A-inhibitor-bound AKeco is studied with slowly relaxing local structure (SRLS) 15N relaxation, an approach particularly suited when the global (τm) and the local (τ) motions are likely to be coupled. For AKeco τm = 15.1 ns, whereas for AKeco*AP5A τm = 11.6 ns. The CORE domain of AKeco features an av. squared order parameter, <S2>, of 0.84 and correlation times τf = 5-130 ps. Most of the AKeco*AP5A backbone features <S2> = 0.90 and τf = 33-193 ps. These data are indicative of relative rigidity. Domains AMPbd and LID of AKeco, and loops β1/α1, α2/α3, α4/β3, α5/β4, and β8/α7 of AKeco*AP5A, feature a novel type of protein flexibility consisting of nanosecond peptide plane reorientation about the Ci-1α-Ciα axis, with correlation time τ.perp. = 5.6-11.3 ns. The other microdynamic parameters underlying this dynamic model include S2 = 0.13-0.5, τ|| on the ps time scale, and a diffusion tilt βMD ranging from 12 to 21°. For the ligand-free enzyme the τ.perp. mode was shown to represent segmental domain motion, accompanied by conformational exchange contributions Rex ≤ 4.4 s-1. Loop α4/β3 and α5/β4 dynamics in AKeco*AP5A is related to the "energetic counter-balancing of substrate binding" effect apparently driving kinase catalysis. The other flexible AKeco*AP5A loops may relate to domain motion toward product release. - 46Chervenak, M. C. and Toone, E. J. J. Am. Chem. Soc. 1994, 116, 10533– 10539
- 47Daranas, A. H., Shimizu, H., and Homans, S. W. J. Am. Chem. Soc. 2004, 126, 11870– 11876[ACS Full Text
], [CAS], Google Scholar
47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXntFWiur0%253D&md5=cbb20561d54f2b3c15f70f04f55dc368Thermodynamics of Binding of D-Galactose and Deoxy Derivatives thereof to the L-Arabinose-binding ProteinDaranas, Antonio Hernandez; Shimizu, Hiroki; Homans, Steve W.Journal of the American Chemical Society (2004), 126 (38), 11870-11876CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)We report the thermodn. of binding of D-galactose and deoxy derivs. thereof to the arabinose binding protein (ABP). The "intrinsic" (solute-solute) free energy of binding ΔG°int at 308 K for the 1-, 2-, 3-, and 6-hydroxyl groups of galactose is remarkably const. (∼-30 kJ/mol), despite the fact that each hydroxyl group subtends different nos. of hydrogen bonds in the complex. The substantially unfavorable enthalpy of binding (∼30 kJ/mol) of 1-deoxygalactose, 2-deoxygalactose, and 3-deoxygalactose in comparison with galactose, cannot be readily accounted for by differences in solvation, suggesting that solute-solute hydrogen bonds are enthalpically significantly more favorable than solute-solvent hydrogen bonds. In contrast, the substantially higher affinity for 2-deoxygalactose in comparison with either 1-deoxygalactose or 3-deoxygalactose derives from differences in the solvation free energies of the free ligands. - 48Lundquist, J. J., Debenham, S. D., and Toone, E. J. J. Org. Chem. 2000, 65, 8245– 8250
- 49Turnbull, W. B., Precious, B. L., and Homans, S. W. J. Am. Chem. Soc. 2004, 126, 1047– 1054[ACS Full Text
], [CAS], Google Scholar
49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXkt1eksg%253D%253D&md5=e453e7810b86d0e2c3a6c0299e32ee4fDissecting the Cholera Toxin-Ganglioside GM1 Interaction by Isothermal Titration CalorimetryTurnbull, W. Bruce; Precious, Bernie L.; Homans, Steve W.Journal of the American Chemical Society (2004), 126 (4), 1047-1054CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The complex of cholera toxin and ganglioside GM1 is one of the highest affinity protein-carbohydrate interactions known. Herein, the GM1 pentasaccharide is dissected into smaller fragments to det. the contribution of each of the key monosaccharide residues to the overall binding affinity. Displacement isothermal titrn. calorimetry (ITC) has allowed the measurement of all of the key thermodn. parameters for even the lowest affinity fragment ligands. Anal. of the std. free energy changes using Jencks' concept of intrinsic free energies reveals that the terminal galactose and sialic acid residues contribute 54% and 44% of the intrinsic binding energy, resp., despite the latter ligand having little appreciable affinity for the toxin. This anal. also provides an est. of 25.8 kJ mol-1 for the loss of independent translational and rotational degrees of freedom on complexation and presents evidence for an alternative binding mode for ganglioside GM2. The high affinity and selectivity of the GM1-cholera toxin interaction originates principally from the conformational preorganization of the branched pentasaccharide rather than through the effect of cooperativity, which is also reinvestigated by ITC. - 50Chang, C. A., Chen, W., and Gilson, M. K. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 1534– 1539
- 51Cabani, S., Gianni, P., Mollica, V., and Lepori, L. J. Solution Chem. 1981, 10, 563– 595
- 52Young, T., Abel, R., Kim, B., Berne, B. J., and Friesner, R. A. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 808– 813
- 53Qvist, J., Davidovic, M., Hamelberg, D., and Halle, B. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 6296– 6301
- 54Henchman, R. J. Chem. Phys. 2007, 126064504
- 55Spolar, R. S., Livingstone, J. R., and Record, M. T. J. Biochemistry (Moscow) 1992, 31, 3947– 3955[ACS Full Text
], [CAS], Google Scholar
55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XhvFSmtb4%253D&md5=18a0267fc7138758d8eff451b37006e6Use of liquid hydrocarbon and amide transfer data to estimate contributions to thermodynamic functions of protein folding from the removal of nonpolar and polar surface from waterSpolar, Ruth S.; Livingstone, Jeff R.; Record, M. Thomas, Jr.Biochemistry (1992), 31 (16), 3947-55CODEN: BICHAW; ISSN:0006-2960.This extension of the liq. hydrocarbon model seeks to quantify the thermodn. contributions to protein stability from the removal of nonpolar and polar surface from water. Thermodn. data for the transfer of hydrocarbons and org. amides from water to the pure liq. phase are analyzed to obtain contributions to the thermodn. of folding from the redn. in water-accessible surface area. Although the removal of nonpolar surface makes the dominant contribution to the std. heat capacity change of folding (ΔCfold°), here the authors show that inclusion of the contribution from removal of polar surface allows a quant. prediction of ΔCfold° within the uncertainty of the calorimetrically detd. value. Moreover, anal. of the contribution of polar surface area to the enthalpy of transfer of liq. amides provides a means of estg. the contributions from changes in nonpolar and polar surface area as well as other factors to the enthalpy of folding (ΔHfold°). In addn. to ests. of ΔHfold°, this extension of the liq. hydrocarbon model provides a thermodn. explanation for the observation that the specific enthalpy of folding (cal g-1) of a no. of globular proteins converges to a common value at approx. 383 K. Because amts. of nonpolar and polar surface area buried by these proteins upon folding are found to be linear functions of mol. mass, ests. of both ΔCfold° and ΔHfold° may be obtained given only the mol. mass of the protein of interest. Use of ΔCfold° and ΔHfold° in conjunction with a melting temp. (Tm) detd. under specified soln. conditions allows one to est. the stability (ΔGfold°) under these conditions at any temp. Calcd. values of ΔGfold° in the range 293-320 K appear generally to agree with those detd. by calorimetric and noncalorimetric methods. This anal. should provide a basis for predicting stability of proteins whose structural and thermodn. properties are similar to those in the data set used for this anal. Deviations from these predictions may be interpretable in terms of deviations from the av. properties of the proteins in the data set. - 56Prabhu, N. V. and Sharp, K. A. Annu. Rev. Phys. Chem. 2005, 56, 521– 548
- 57Sturtevant, J. M. Proc. Natl. Acad. Sci. U.S.A. 1977, 74, 2236– 2240
- 58Burkhalter, N. F., Dimick, S. M., and Toone, E. J. In Carbohydrates in Chemistry and Biology. Part I: Chemistry of Saccharides; Ernst, B., Hart, G. W., and Sinay, P., Eds.; Wiley-VCH: Weinheim, 2000; Vol. 2, pp 863− 914.
- 59Lemieux, R. U. Acc. Chem. Res. 1996, 29, 373– 380
Cited By
This article is cited by 45 publications.
- Falk Hoffmann, Frans A. A. Mulder, Lars V. Schäfer. How Much Entropy Is Contained in NMR Relaxation Parameters?. The Journal of Physical Chemistry B 2022, 126 (1) , 54-68. https://doi.org/10.1021/acs.jpcb.1c07786
- Johan Wallerstein, Vilhelm Ekberg, Majda Misini Ignjatović, Rohit Kumar, Octav Caldararu, Kristoffer Peterson, Sven Wernersson, Ulrika Brath, Hakon Leffler, Esko Oksanen, Derek T. Logan, Ulf J. Nilsson, Ulf Ryde, Mikael Akke. Entropy–Entropy Compensation between the Protein, Ligand, and Solvent Degrees of Freedom Fine-Tunes Affinity in Ligand Binding to Galectin-3C. JACS Au 2021, 1 (4) , 484-500. https://doi.org/10.1021/jacsau.0c00094
- Zhe Huai, Zhaoxi Shen, Zhaoxi Sun. Binding Thermodynamics and Interaction Patterns of Inhibitor-Major Urinary Protein-I Binding from Extensive Free-Energy Calculations: Benchmarking AMBER Force Fields. Journal of Chemical Information and Modeling 2021, 61 (1) , 284-297. https://doi.org/10.1021/acs.jcim.0c01217
- Mona Sarter, Doreen Niether, Bernd W. Koenig, Wiebke Lohstroh, Michaela Zamponi, Niina H. Jalarvo, Simone Wiegand, Jörg Fitter, Andreas M. Stadler. Strong Adverse Contribution of Conformational Dynamics to Streptavidin–Biotin Binding. The Journal of Physical Chemistry B 2020, 124 (2) , 324-335. https://doi.org/10.1021/acs.jpcb.9b08467
- Maria Luisa Verteramo, Olof Stenström, Majda Misini Ignjatović, Octav Caldararu, Martin A. Olsson, Francesco Manzoni, Hakon Leffler, Esko Oksanen, Derek T. Logan, Ulf J. Nilsson, Ulf Ryde, Mikael Akke. Interplay between Conformational Entropy and Solvation Entropy in Protein–Ligand Binding. Journal of the American Chemical Society 2019, 141 (5) , 2012-2026. https://doi.org/10.1021/jacs.8b11099
- Hannah E. Bruce Macdonald, Christopher Cave-Ayland, Gregory A. Ross, Jonathan W. Essex. Ligand Binding Free Energies with Adaptive Water Networks: Two-Dimensional Grand Canonical Alchemical Perturbations. Journal of Chemical Theory and Computation 2018, 14 (12) , 6586-6597. https://doi.org/10.1021/acs.jctc.8b00614
- Martin J. Thiele, Mehdi D. Davari, Melanie König, Isabell Hofmann, Niklas O. Junker, Tayebeh Mirzaei Garakani, Ljubica Vojcic, Jörg Fitter, Ulrich Schwaneberg. Enzyme–Polyelectrolyte Complexes Boost the Catalytic Performance of Enzymes. ACS Catalysis 2018, 8 (11) , 10876-10887. https://doi.org/10.1021/acscatal.8b02935
- Stefan Geschwindner, Johan Ulander, and Patrik Johansson . Ligand Binding Thermodynamics in Drug Discovery: Still a Hot Tip?. Journal of Medicinal Chemistry 2015, 58 (16) , 6321-6335. https://doi.org/10.1021/jm501511f
- Subhajit Ghosh and Joykrishna Dey . Binding of Fatty Acid Amide Amphiphiles to Bovine Serum Albumin: Role of Amide Hydrogen Bonding. The Journal of Physical Chemistry B 2015, 119 (25) , 7804-7815. https://doi.org/10.1021/acs.jpcb.5b00965
- Julien Michel, Richard H. Henchman, Georgios Gerogiokas, Michelle W. Y. Southey, Michael P. Mazanetz, and Richard J. Law . Evaluation of Host–Guest Binding Thermodynamics of Model Cavities with Grid Cell Theory. Journal of Chemical Theory and Computation 2014, 10 (9) , 4055-4068. https://doi.org/10.1021/ct500368p
- P. Ross Wilderman, Manish B. Shah, Hyun-Hee Jang, C. David Stout, and James R. Halpert . Structural and Thermodynamic Basis of (+)-α-Pinene Binding to Human Cytochrome P450 2B6. Journal of the American Chemical Society 2013, 135 (28) , 10433-10440. https://doi.org/10.1021/ja403042k
- Yi Wang, Jason R. King, Pan Wu, Daniel L. Pelzman, David N. Beratan, and Eric J. Toone . Enthalpic Signature of Methonium Desolvation Revealed in a Synthetic Host–Guest System Based on Cucurbit[7]uril. Journal of the American Chemical Society 2013, 135 (16) , 6084-6091. https://doi.org/10.1021/ja311327v
- Vidhi Mishra and Donald R. Ronning . Crystal Structures of the Helicobacter pylori MTAN Enzyme Reveal Specific Interactions between S-Adenosylhomocysteine and the 5′-Alkylthio Binding Subsite. Biochemistry 2012, 51 (48) , 9763-9772. https://doi.org/10.1021/bi301221k
- Ying-Ting Lin and Guan-Yu Chen . A Scaffold-Independent Subcellular Event-Based Analysis: Characterization of Significant Structural Modifications. Journal of Chemical Information and Modeling 2012, 52 (2) , 506-514. https://doi.org/10.1021/ci200540y
- Yue Shi, Crystal Z. Zhu, Stephen F. Martin, and Pengyu Ren . Probing the Effect of Conformational Constraint on Phosphorylated Ligand Binding to an SH2 Domain Using Polarizable Force Field Simulations. The Journal of Physical Chemistry B 2012, 116 (5) , 1716-1727. https://doi.org/10.1021/jp210265d
- Kitiyaporn Wittayanarakul, Nahoum G. Anthony, Witcha Treesuwan, Supa Hannongbua, Hasan Alniss, Abedawn I. Khalaf, Colin J. Suckling, John A Parkinson, and Simon P. Mackay . Ranking Ligand Affinity for the DNA Minor Groove by Experiment and Simulation. ACS Medicinal Chemistry Letters 2010, 1 (8) , 376-380. https://doi.org/10.1021/ml100047n
- Bharti, Maya S. Nair. Spectroscopic methods to study the thermodynamics of biomolecular interactions. 2023, 375-413. https://doi.org/10.1016/B978-0-323-99127-8.00001-5
- Maciej Przybyłek, Piotr Bełdowski, Florian Wieland, Piotr Cysewski, Alina Sionkowska. Collagen Type II—Chitosan Interactions as Dependent on Hydroxylation and Acetylation Inferred from Molecular Dynamics Simulations. Molecules 2023, 28 (1) , 154. https://doi.org/10.3390/molecules28010154
- Piotr Bełdowski, Maciej Przybyłek, Damian Bełdowski, Andra Dedinaite, Alina Sionkowska, Piotr Cysewski, Per M. Claesson. Collagen type II–hyaluronan interactions – the effect of proline hydroxylation: a molecular dynamics study. Journal of Materials Chemistry B 2022, 10 (46) , 9713-9723. https://doi.org/10.1039/D2TB01550A
- Suriya Tateing, Nuttee Suree, . Decoding molecular recognition of inhibitors targeting HDAC2 via molecular dynamics simulations and configurational entropy estimation. PLOS ONE 2022, 17 (8) , e0273265. https://doi.org/10.1371/journal.pone.0273265
- Olatomide A. Fadare, Nusrat O. Omisore, Oluwaseun B. Adegbite, Oladoja A. Awofisayo, Frank A. Ogundolie, Julius K. Adesanwo, Craig A. Obafemi. Structure based design, stability study and synthesis of the dinitrophenylhydrazone derivative of the oxidation product of lanosterol as a potential P. falciparum transketolase inhibitor and in-vivo antimalarial study. In Silico Pharmacology 2021, 9 (1) https://doi.org/10.1007/s40203-021-00097-8
- Homero Gómez-Velasco, Arturo Rojo-Domínguez, Enrique García-Hernández. Enthalpically-driven ligand recognition and cavity solvation of bovine odorant binding protein. Biophysical Chemistry 2020, 257 , 106315. https://doi.org/10.1016/j.bpc.2019.106315
- Sneha Singh, Johannes Dodt, Peter Volkers, Emma Hethershaw, Helen Philippou, Vytautus Ivaskevicius, Diana Imhof, Johannes Oldenburg, Arijit Biswas. Structure functional insights into calcium binding during the activation of coagulation factor XIII A. Scientific Reports 2019, 9 (1) https://doi.org/10.1038/s41598-019-47815-z
- Nausheen Joondan, Salma Bibi Moosun, Prakashanand Caumul, Suthananda N. Sunassee, Gerhard A. Venter, Sabina Jhaumeer-Laulloo. Effect of chain length on the interactions of sodium N-alkyl prolinates with bovine serum albumin: a spectroscopic investigation and molecular docking simulations. Colloid and Polymer Science 2018, 296 (2) , 367-378. https://doi.org/10.1007/s00396-017-4251-1
- Karpagaraj Malarkani, Ivy Sarkar, Susithra Selvam. Denaturation studies on bovine serum albumin–bile salt system: Bile salt stabilizes bovine serum albumin through hydrophobicity. Journal of Pharmaceutical Analysis 2018, 8 (1) , 27-36. https://doi.org/10.1016/j.jpha.2017.06.007
- José A. Caro, Kyle W. Harpole, Vignesh Kasinath, Jackwee Lim, Jeffrey Granja, Kathleen G. Valentine, Kim A. Sharp, A. Joshua Wand. Entropy in molecular recognition by proteins. Proceedings of the National Academy of Sciences 2017, 114 (25) , 6563-6568. https://doi.org/10.1073/pnas.1621154114
- G. Klebe. Profiling Drug Binding by Thermodynamics. 2017, 1-36. https://doi.org/10.1016/B978-0-08-101129-4.00001-1
- John F. Andersen, José M.C. Ribeiro. Salivary Kratagonists. 2017, 51-63. https://doi.org/10.1016/B978-0-12-805360-7.00004-6
- Yanan Wang, Zhuang Li, Yongzhi Zhou, Jie Cao, Houshuang Zhang, Haiyan Gong, Jinlin Zhou. Specific histamine binding activity of a new lipocalin from Hyalomma asiaticum (Ixodidae) and therapeutic effects on allergic asthma in mice. Parasites & Vectors 2016, 9 (1) https://doi.org/10.1186/s13071-016-1790-0
- Tinoush Moulaei, Kabamba B Alexandre, Shilpa R Shenoy, Joel R Meyerson, Lauren RH Krumpe, Brian Constantine, Jennifer Wilson, Robert W Buckheit, James B McMahon, Sriram Subramaniam, Alexander Wlodawer, Barry R O’Keefe. Griffithsin tandemers: flexible and potent lectin inhibitors of the human immunodeficiency virus. Retrovirology 2015, 12 (1) https://doi.org/10.1186/s12977-014-0127-3
- Subhajit Ghosh, Joykrishna Dey. Interaction of bovine serum albumin with N-acyl amino acid based anionic surfactants: Effect of head-group hydrophobicity. Journal of Colloid and Interface Science 2015, 458 , 284-292. https://doi.org/10.1016/j.jcis.2015.07.064
- Michael R. Duff, Elizabeth E. Howell. Thermodynamics and solvent linkage of macromolecule–ligand interactions. Methods 2015, 76 , 51-60. https://doi.org/10.1016/j.ymeth.2014.11.009
- Gerhard Klebe. Applying thermodynamic profiling in lead finding and optimization. Nature Reviews Drug Discovery 2015, 14 (2) , 95-110. https://doi.org/10.1038/nrd4486
- Grace-Anne Bent, Paul Maragh, Tara Dasgupta, Richard A. Fairman, Lebert Grierson. Kinetic and density functional theory (DFT) studies of in vitro reactions of acrylamide with the thiols: captopril, l -cysteine, and glutathione. Toxicology Research 2015, 4 (1) , 121-131. https://doi.org/10.1039/C4TX00070F
- Charfedinne Ayed, Samuel Lubbers, Isabelle Andriot, Yacine Merabtine, Elisabeth Guichard, Anne Tromelin. Impact of structural features of odorant molecules on their retention/release behaviours in dairy and pectin gels. Food Research International 2014, 62 , 846-859. https://doi.org/10.1016/j.foodres.2014.04.050
- Satoshi Kume, Young-Ho Lee, Masatoshi Nakatsuji, Yoshiaki Teraoka, Keisuke Yamaguchi, Yuji Goto, Takashi Inui. Fine-tuned broad binding capability of human lipocalin-type prostaglandin D synthase for various small lipophilic ligands. FEBS Letters 2014, 588 (6) , 962-969. https://doi.org/10.1016/j.febslet.2014.02.001
- Michael D. Shultz. The thermodynamic basis for the use of lipophilic efficiency (LipE) in enthalpic optimizations. Bioorganic & Medicinal Chemistry Letters 2013, 23 (21) , 5992-6000. https://doi.org/10.1016/j.bmcl.2013.08.030
- F.F.O. Sousa, A. Luzardo-Álvarez, J. Blanco-Méndez, F.J. Otero-Espinar, M. Martín-Pastor, I. Sández Macho. Use of 1H NMR STD, WaterLOGSY, and Langmuir monolayer techniques for characterization of drug–zein protein complexes. European Journal of Pharmaceutics and Biopharmaceutics 2013, 85 (3) , 790-798. https://doi.org/10.1016/j.ejpb.2013.07.008
- Teena Goel, Santosh Kumar, Souvik Maiti. Thermodynamics and solvation dynamics of BIV TAR RNA–Tat peptide interaction. Mol. BioSyst. 2013, 9 (1) , 88-98. https://doi.org/10.1039/C2MB25357G
- F.F.O. Sousa, Asteria Luzardo-Álvarez, José Blanco-Méndez, Manuel Martín-Pastor. NMR techniques in drug delivery: Application to zein protein complexes. International Journal of Pharmaceutics 2012, 439 (1-2) , 41-48. https://doi.org/10.1016/j.ijpharm.2012.09.046
- Samuel Genheden, Ulf Ryde. Will molecular dynamics simulations of proteins ever reach equilibrium?. Physical Chemistry Chemical Physics 2012, 14 (24) , 8662. https://doi.org/10.1039/c2cp23961b
- Phillip W. Snyder, Jasmin Mecinović, Demetri T. Moustakas, Samuel W. Thomas, Michael Harder, Eric T. Mack, Matthew R. Lockett, Annie Héroux, Woody Sherman, George M. Whitesides. Mechanism of the hydrophobic effect in the biomolecular recognition of arylsulfonamides by carbonic anhydrase. Proceedings of the National Academy of Sciences 2011, 108 (44) , 17889-17894. https://doi.org/10.1073/pnas.1114107108
- Richard A Bryce. Physics-based scoring of protein–ligand interactions: explicit polarizability, quantum mechanics and free energies. Future Medicinal Chemistry 2011, 3 (6) , 683-698. https://doi.org/10.4155/fmc.11.30
- Sébastien Dilly, Jean-François Liégeois. Interaction of clozapine and its nitrenium ion with rat D2 dopamine receptors: in vitro binding and computational study. Journal of Computer-Aided Molecular Design 2011, 25 (2) , 163-169. https://doi.org/10.1007/s10822-010-9407-8
- Gloria Fuentes, Shubhra Ghosh Dastidar, Arumugam Madhumalar, Chandra S. Verma. Role of protein flexibility in the discovery of new drugs. Drug Development Research 2011, 72 (1) , 26-35. https://doi.org/10.1002/ddr.20399
Abstract
Figure 1
Figure 1. Stereo view of the binding site of rRaHBP2(D24R) with bound histamine. This site corresponds to the H site of the wild-type protein. Binding-site residues are colored blue, the ligand is colored red, and ordered water molecules are shown as green spheres. Figure prepared using MOLMOL (36)
Figure 2
Figure 2. Typical ITC isotherms for the binding of histamine to (above) rRaHBP2(D24R) and (below) wild-type rRaHBP2 at 298 K.
Figure 3
Figure 3. Stereo images of the rRaHBP2(D24R)−histamine complex, showing positive (green) and negative (blue) contributions or no contribution (within the standard error, red) to the overall binding entropy from amide bond vectors on binding histamine (blue and yellow spheres).
Figure 4
Figure 4. (Left) Typical mean square displacements of solvent water molecules in the binding pocket of uncomplexed rRaHBP2(D24R) (red, green, and blue traces) in comparison with bulk water (black trace). (Right) Typical mean square displacements of solvent water molecules in the binding pocket of rRaHBP2(D24R) in complex with histamine (red, green, blue, and gray traces). In both panels the inset shows rotational autocorrelation functions for these waters using the same color scheme.
References
ARTICLE SECTIONSThis article references 59 other publications.
- 1Bingham, R., Bodenhausen, G., Findlay, J. H. B. C., Hsieh, S.-Y., Kalverda, A. P., Kjellberg, A., Perazzolo, C., Phillips, S. E. V., Seshadri, K., Turnbull, W. B., and Homans, S. W. J. Am. Chem. Soc. 2004, 126, 1675– 1681
- 2Barratt, E., Bingham, R., Warner, D. J., Laughton, C. A., Phillips, S. E. V., and Homans, S. W. J. Am. Chem. Soc. 2005, 127, 11827– 11834
- 3Malham, R., Johnstone, S., Bingham, R. J., Barratt, E., Phillips, S. E., Laughton, C. A., and Homans, S. W. J. Am. Chem. Soc. 2005, 127, 17061– 7
- 4Shimokhina, N., Bronowska, A., and Homans, S. W. Angew. Chem., Int. Ed. 2006, 45, 6374– 6376
- 5Delaglio, F., Grzesiek, S., Vuister, G. W., Zhu, G., Pfeifer, J., and Bax, A. J. Biomol. NMR 1995, 6, 277– 293[Crossref], [PubMed], [CAS], Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXhtVSmurfK&md5=a670fca5b164083e2178fafd2fb951ffNMRPipe: a multidimensional spectral processing system based on UNIX pipesDelaglio, Frank; Grzesiek, Stephan; Vuister, Geerten W.; Zhu, Guang; Pfeifer, John; Bax, AdJournal of Biomolecular NMR (1995), 6 (3), 277-93CODEN: JBNME9; ISSN:0925-2738. (ESCOM)The NMRPipe system is a UNIX software environment of processing, graphics, and anal. tools designed to meet current routine and research-oriented multidimensional processing requirements, and to anticipate and accommodate future demands and developments. The system is based on UNIX pipes, which allow programs running simultaneously to exchange streams of data under user control. In an NMRPipe processing scheme, a stream of spectral data flows through a pipeline of processing programs, each of which performs one component of the overall scheme, such as Fourier transformation or linear prediction. Complete multidimensional processing schemes are constructed as simple UNIX shell scripts. The processing modules themselves maintain and exploit accurate records of data sizes, detection modes, and calibration information in all dimensions, so that schemes can be constructed without the need to explicitly define or anticipate data sizes or storage details of real and imaginary channels during processing. The asynchronous pipeline scheme provides other substantial advantages, including high flexibility, favorable processing speeds, choice of both all-in-memory and disk-bound processing, easy adaptation to different data formats, simpler software development and maintenance, and the ability to distribute processing tasks on multi-CPU computers and computer networks.
- 6Vranken, W. F., Boucher, W., Stevens, T. J., Fogh, R. H., Pajon, A., Llinas, M., Ulrich, E. L., Markley, J. L., Ionides, J., and Laue, E. D. Proteins 2005, 59, 687– 696
- 7Farrow, N. A., Muhandiram, R., Singer, A. U., Pascal, S. M., Kay, C. M., Gish, G., Shoelson, S. E., Pawson, T., Forman-Kay, J. D., and Kay, L. E. Biochemistry (Moscow) 1994, 33, 5984– 6003[ACS Full Text
], [CAS], Google Scholar
7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXktVersbk%253D&md5=57fb1299b4acd2eb8788a83f1c322257Backbone Dynamics of a Free and a Phosphopeptide-Complexed Src Homology 2 Domain Studied by 15N NMR RelaxationFarrow, Neil A.; Muhandiram, Ranjith; Singer, Alex U.; Pascal, Steven M.; Kay, Cyril M.; Gish, Gerry; Shoelson, Steven E.; Pawson, Tony; Forman-Kay, Julie D.; Kay, Lewis E.Biochemistry (1994), 33 (19), 5984-6003CODEN: BICHAW; ISSN:0006-2960.The backbone dynamics of the C-terminal Src Homol. 2 (SH2) domain of phosphatidylinositol phospholipase Cγ1 were investigated. Two forms of the domain were studied, one in complex with a high-affinity binding peptide derived from the platelet-derived growth factor receptor and the other in the absence of this peptide. Two-dimensional 1H-15N NMR methods, employing pulsed field gradients, were used to det. steady-state 1H-15N NOE values and T1 and T2 15N relaxation times. Backbone dynamics were characterized by the overall correlation time (τm), order parameters (S2), effective correlation times for internal motions (τe), and, if required, terms to account for motions on a microsecond-to-millisecond-time scale. An extended 2-time-scale formalism was used for residues having relaxation data that could not be fit adequately using a single-time-scale formalism. The overall correlation times of the uncomplexed and complexed forms of SH2 were found to be 9.2 and 6.5 ns, resp., suggesting that the uncomplexed form was in a monomer-dimer equil. This was subsequently confirmed by hydrodynamic measurements. Anal. of order parameters revealed that residues in the so-called phosphotyrosine-binding loop exhibited higher than av. disorder in both forms of SH2. Although localized differences in order parameters were obsd. between the uncomplexed and complexed forms of SH2, overall, higher order parameters were not found in the peptide-bound form, indicating that on av., picosecond-time-scale disorder is not reduced upon binding peptide. The relaxation data of the SH2-phosphopeptide complex were fit with fewer exchange terms than the uncomplexed form. This may reflect the monomer-dimer equil. that exists in the uncomplexed form or may indicate that the complexed form has lower conformational flexibility on a microsecond-to-millisecond-time scale. - 8Johnson, B. A. and Blevins, R. A. J. Biomol. NMR 1994, 4, 603– 614[Crossref], [PubMed], [CAS], Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmt1Gkurw%253D&md5=6a93a0d9704fbbb654abc192203c68aeNMRView: a computer program for the visualization and analysis of NMR dataJohnson, Bruce A.; Blevins, Richard A.Journal of Biomolecular NMR (1994), 4 (5), 603-14CODEN: JBNME9; ISSN:0925-2738.NMRView is a computer program designed for the visualization and anal. of NMR data. It allows the user to interact with a practically unlimited no. of 2D, 3D and 4D NMR data files. Any no. of spectral windows can be displayed on the screen in any size and location. Automatic peak picking and facilitated peak anal. features are included to aid in the assignment of complex NMR spectra. NMR View provides structure anal. features and data transfer to and from structural generation programs, allowing for a tight coupling between the spectral anal. and structural generation.,. Visual correlation between structures and spectra can be done with the Mol. Data Viewer, a mol. graphics program with bidirectional communication to NMR View. The used interface can be customized and a command language is provided to allow for the automation of various tasks.
- 9García de la Torre, J., Huertas, M. L., and Carrasco, B. J. Magn. Reson. 2000, 147, 138– 146
- 10d’Auvergne, E. J. and Gooley, P. R. J. Biomol. NMR 2008, 40, 107– 119
- 11d’Auvergne, E. J. and Gooley, P. R. J. Biomol. NMR 2008, 40, 121– 133
- 12Yang, D. and Kay, L. E. J. Mol. Biol. 1996, 263, 369– 82
- 13Leslie, A. G. W. In CCP4 & ESFEACMB Newsletter On Protein Crystallography; Daresbury Laboratory: Warington, 1992.Google ScholarThere is no corresponding record for this reference.
- 14CCP4. Acta Crystallogr. 1994, D50, 760− 763Google ScholarThere is no corresponding record for this reference.
- 15Murshudov, G. N., Vagin, A. A., and Dodson, E. J. Acta Crystallogr. 1997, D53, 240– 255Google ScholarThere is no corresponding record for this reference.
- 16Emsley, P. and Cowtan, K. Acta Crystallogr. 2004, D60, 2126– 2132Google ScholarThere is no corresponding record for this reference.
- 17Laskowski, R. A., MacArthur, M. W., Moss, D. S., and Thornton, J. M. J. Appl. Crystallogr. 1993, 26, 283– 291[Crossref], [CAS], Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXit12lurY%253D&md5=fb69fc4410cd716aaaa7cc0db06b3ed2PROCHECK: a program to check the stereochemical quality of protein structuresLaskowski, Roman A.; MacArthur, Malcolm W.; Moss, David S.; Thornton, Janet M.Journal of Applied Crystallography (1993), 26 (2), 283-91CODEN: JACGAR; ISSN:0021-8898.The PROCHECK suite of programs provides a detailed check on the stereochem. of a protein structure. Its outputs comprise a no. of plots in PostScript format and a comprehensive residue-by-residue listing. These give an assessment of the overall quality of the structure as compared with well-refined structures of the same resoln. and also highlight regions that may need further investigation. The PROCHECK programs are useful for assessing the quality not only of protein structures in the process of being solved but also of existing structures and of those being modeled on known structures.
- 18Davis, I. W., Leaver-Fay, A., Chen, V. B., Block, J. N., Kapral, G. J., Wang, X., Murray, L. W., Arendall, I., W. B., Snoeyink, J., Richardson, J. S., and Richardson, D. C. Nucleic Acids Res. 2007, 35W375−W383
- 19Navaza, J. Acta Crystallogr., Sect. A 1994, 50, 157– 163
- 20Case, D. A. et al. AMBER 8; University of California: San Francisco, 2004.Google ScholarThere is no corresponding record for this reference.
- 21Cornell, W. D., Cieplak, P., Bayly, C. I., Gould, I. R., Merz, K. M., Ferguson, D. M., Spellmeyer, D. C., Fox, T., Caldwell, J. W., and Kollman, P. A. J. Am. Chem. Soc. 1995, 117, 5179– 5197[ACS Full Text
], [CAS], Google Scholar
21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXlsFertrc%253D&md5=2c8901ed614d543db51d3a28f5b6053cA Second Generation Force Field for the Simulation of Proteins, Nucleic Acids, and Organic MoleculesCornell, Wendy D.; Cieplak, Piotr; Bayly, Christopher I.; Gould, Ian R.; Merz, Kenneth M., Jr.; Ferguson, David M.; Spellmeyer, David C.; Fox, Thomas; Caldwell, James W.; Kollman, Peter A.Journal of the American Chemical Society (1995), 117 (19), 5179-97CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The authors present the derivation of a new mol. mech. force field for simulating the structures, conformational energies, and interaction energies of proteins, nucleic acids, and many related org. mols. in condensed phases. This effective two-body force field is the successor to the Weiner et al. force field and was developed with some of the same philosophies, such as the use of a simple diagonal potential function and electrostatic potential fit atom centered charges. The need for a 10-12 function for representing hydrogen bonds is no longer necessary due to the improved performance of the new charge model and new van der Waals parameters. These new charges are detd. using a 6-31G* basis set and restrained electrostatic potential (RESP) fitting and have been shown to reproduce interaction energies, free energies of solvation, and conformational energies of simple small mols. to a good degree of accuracy. Furthermore, the new RESP charges exhibit less variability as a function of the mol. conformation used in the charge detn. The new van der Waals parameters have been derived from liq. simulations and include hydrogen parameters which take into account the effects of any geminal electroneg. atoms. The bonded parameters developed by Weiner et al. were modified as necessary to reproduce exptl. vibrational frequencies and structures. Most of the simple dihedral parameters have been retained from Weiner et al., but a complex set of .vphi. and ψ parameters which do a good job of reproducing the energies of the low-energy conformations of glycyl and alanyl dipeptides was developed for the peptide backbone. - 22Hornak, V., Abel, R., Okur, A., Strockbine, B, Roitberg, A., and Simmerling, C. Proteins: Struct., Funct., Bioinf. 2006, 65, 712– 725
- 23Wang, J., Wolf, R. M., Caldwell, J., Kollman, P., and Case, D. A. J. Comput. Chem. 2004, 25, 1157– 1174[Crossref], [PubMed], [CAS], Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksFakurc%253D&md5=2992017a8cf51f89290ae2562403b115Development and testing of a general Amber force fieldWang, Junmei; Wolf, Romain M.; Caldwell, James W.; Kollman, Peter A.; Case, David A.Journal of Computational Chemistry (2004), 25 (9), 1157-1174CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)We describe here a general Amber force field (GAFF) for org. mols. GAFF is designed to be compatible with existing Amber force fields for proteins and nucleic acids, and has parameters for most org. and pharmaceutical mols. that are composed of H, C, N, O, S, P, and halogens. It uses a simple functional form and a limited no. of atom types, but incorporates both empirical and heuristic models to est. force consts. and partial at. charges. The performance of GAFF in test cases is encouraging. In test I, 74 crystallog. structures were compared to GAFF minimized structures, with a root-mean-square displacement of 0.26 Å, which is comparable to that of the Tripos 5.2 force field (0.25 Å) and better than those of MMFF 94 and CHARMm (0.47 and 0.44 Å, resp.). In test II, gas phase minimizations were performed on 22 nucleic acid base pairs, and the minimized structures and intermol. energies were compared to MP2/6-31G* results. The RMS of displacements and relative energies were 0.25 Å and 1.2 kcal/mol, resp. These data are comparable to results from Parm99/RESP (0.16 Å and 1.18 kcal/mol, resp.), which were parameterized to these base pairs. Test III looked at the relative energies of 71 conformational pairs that were used in development of the Parm99 force field. The RMS error in relative energies (compared to expt.) is about 0.5 kcal/mol. GAFF can be applied to wide range of mols. in an automatic fashion, making it suitable for rational drug design and database searching.
- 24Frisch, M. J. et al. Gaussian 98 (Revision A.9); Gaussian, Inc.: Pittsburgh, PA, 1998.Google ScholarThere is no corresponding record for this reference.
- 25Cornell, W. D., Cieplak, P., Bayly, C. I., and Kollman, P. A. J. Am. Chem. Soc. 1993, 115, 9620– 9631[ACS Full Text
], [CAS], Google Scholar
25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXmt1Gmsbk%253D&md5=448e45857da2a2fcdbb19e127e7b48a1Application of RESP charges to calculate conformational energies, hydrogen bond energies, and free energies of solvationCornell, Wendy D.; Cieplak, Piotr; Bayly, Christopher I.; Kollmann, Peter A.Journal of the American Chemical Society (1993), 115 (21), 9620-31CODEN: JACSAT; ISSN:0002-7863.The authors apply a new restrained electrostatic potential fit charge model (two-stage RESP) to conformational anal. and the calcn. of intermol. interactions. Specifically, the authors study conformational energies in butane, Me Et thioether, three simple alcs., three simple amines, and 1,2-ethanediol as a function of charge model (two-stage RESP vs std. ESP) and 1-4 electrostatic scale factor. The authors demonstrate that the two-stage RESP model with a 1-4 electrostatic scale factor ∼1/1.2 is a very good model, as evaluated by comparison with high-level ab initio calcns. For methanol and N-methylacetamide interactions with TIP3P water, the two-stage RESP model leads to hydrogen bonds only slightly weaker than found with the std. ESP changes. In tests on DNA base pairs, the two-stage RESP model leads to hydrogen bonds which are ∼1 kcal/mol weaker than those calcd. with the std. ESP charges but closer in magnitude to the best currently available ab initio calcns. Furthermore, the two-stage RESP charges, unlike the std. ESP charges, reproduce the result that Hoogsteen hydrogen bonding is stronger than Watson-Crick hydrogen bonding for adenine-thymine base pairs. The free energies of solvation of both methanol and trans-N-methylacetamide were also calcd. for the std. ESP and two-stage RESP models and both were in good agreement with expt. The authors have combined the use of two-stage RESP charges with multiple conformational fitting in studies of conformationally dependent dipole moments and energies of propylamine. The authors find that the combination of these approaches is synergistic in leading to useful charge distributions for mol. simulations. Two-stage RESP charges thus reproduce both intermol. and intramol. energies and structure quite well, making this charge model a crit. advancement in the development of a general force field for modeling biol. macromols. and their ligands, both in the gas phase and in soln. - 26Darden, T., York, D., and Pedersen, L. J. Chem. Phys. 1993, 98, 10089– 10092[Crossref], [CAS], Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXks1Ohsr0%253D&md5=3c9f230bd01b7b714fd096d4d2e755f6Particle mesh Ewald: an N·log(N) method for Ewald sums in large systemsDarden, Tom; York, Darrin; Pedersen, LeeJournal of Chemical Physics (1993), 98 (12), 10089-92CODEN: JCPSA6; ISSN:0021-9606.An N·log(N) method for evaluating electrostatic energies and forces of large periodic systems is presented. The method is based on interpolation of the reciprocal space Ewald sums and evaluation of the resulting convolution using fast Fourier transforms. Timings and accuracies are presented for three large cryst. ionic systems.
- 27Lipari, G. and Szabo, A. J. Am. Chem. Soc. 1982, 104, 4546– 4559[ACS Full Text
], [CAS], Google Scholar
27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XltVejtr8%253D&md5=7188e058f88fbaf1ba066d08a77e13a7Model-free approach to the interpretation of nuclear magnetic resonance relaxation in macromolecules. 1. Theory and range of validityLipari, Giovanni; Szabo, AttilaJournal of the American Chemical Society (1982), 104 (17), 4546-59CODEN: JACSAT; ISSN:0002-7863.A new approach to the interpretation of NMR relaxation expts. on macromols. in soln. is presented. Theor. foundations are dealt with and the range of validity of the approach is established. For both isotropic and anisotropic overall motion, the unique information on fast internal motions contained in relaxation expts. can be completely specified by 2 model-independent quantities: (1) a generalized order parameter, .SCRIPTS., which is a measure of the spatial restriction of the motion, and (2) an effective correlation time, τe, which is a measure of the rate of motion. A simple expression for the spectral d. involving these 2 parameters is derived and is exact when the internal (but not overall) motions are in the extreme narrowing limit. The model-free approach consists of using the above spectral d. to det. least-squares-fit relaxation data by treating .SCRIPTS.2 and τe as adjustable parameters. The range of validity of this approach is illustrated by analyzing error-free relaxation data generated by using sophisticated dynamic models. Empirical rules are presented that allow one to est. the accuracy of .SCRIPTS.2 and τe extd. by using the model-free approach by considering their numerical values, the resonance frequencies, and the parameters for the overall motion. For fast internal motions, it is unnecessary to use approaches based on complicated spectral densities derived within the framework of a model because all models that can give the correct value of .SCRIPTS.2 work equally well. The unique dynamic information (.SCRIPTS. and τe) can be easily extd. by using the model-free approach. If a phys. picture of the motion is desired, the numerical values of .SCRIPTS.2 and τe can be readily interpreted within a phys. reasonable model. - 28Chatfield, D. C., Szabo, A., and Brooks, B. R. J. Am. Chem. Soc. 1998, 120, 5301– 5311
- 29MacRaild, C. A., Daranas, A. H., Bronowska, A., and Homans, S. W. J. Mol. Biol. 2007, 368, 822– 832
- 30Stockmann, H., Bronowska, A., Syme, N. R., Thompson, G. S., Kalverda, A. P., Warriner, S. L., and Homans, S. W. J. Am. Chem. Soc. 2008, 130, 12420– 12426
- 31Best, R. B., Clarke, J., and Karplus, M. J. Mol. Biol. 2005, 349, 185– 203
- 32Schlitter, J. Chem. Phys. Lett. 1993, 215, 617– 621
- 33Klamt, A. and Schuurmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 5, 799– 805
- 34Smith, D. E. and Haymet, A. D. J. J. Chem. Phys. 1993, 98, 6445– 6454
- 35Paesen, G. C., Adams, P. L., Harlos, K., Nuttall, P. A., and Stuart, D. I. Mol. Cell 1999, 3, 661– 671
- 36Koradi, R., Billeter, M., and Wuthrich, K. J. Mol. Graphics 1996, 14, 51[Crossref], [PubMed], [CAS], Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xis12nsrc%253D&md5=dab669ff481b19cde8fbe2e4234711faMOLMOL: a program for display and analysis of macromolecular structuresKoradi, Reto; Billeter, Martin; Wuethrich, KurtJournal of Molecular Graphics (1996), 14 (1), 51-5, plates, 29-32CODEN: JMGRDV; ISSN:0263-7855. (Elsevier)MOLMOL is a mol. graphics program for display, anal., and manipulation of three-dimensional structures of biol. macromols., with special emphasis on NMR soln. structures of proteins and nucleic acids. MOLMOL has a graphical user interface with menus, dialog boxes, and online help. The display possibilities include conventional presentation, as well as novel schematic drawings, with the option of combining different presentations in one view of a mol. Covalent mol. structures can be modified by addn. or removal of individual atoms and bonds, and three-dimensional structures can be manipulated by interactive rotation about individual bonds. Special efforts were made to allow for appropriate display and anal. of the sets of typically 20-40 conformers that are conventionally used to represent the result of an NMR structure detn., using functions for superimposing sets of conformers, calcn. of root mean square distance (RMSD) values, identification of hydrogen bonds, checking and displaying violations of NMR constraints, and identification and listing of short distances between pairs of hydrogen atoms.
- 37Goldberg, R. N., Kishore, N., and Lennen, R. M. J. Phys. Chem. Ref. Data 2002, 31, 231– 370
- 38Ikura, M., Kay, L. E., and Bax, A. Biochemistry (Moscow) 1990, 29, 4659– 4667[ACS Full Text
], [CAS], Google Scholar
38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXitFegs7g%253D&md5=cc22f1867438dfbca353275b77773b55A novel approach for sequential assignment of proton, carbon-13, and nitrogen-15 spectra of larger proteins: heteronuclear triple-resonance three-dimensional NMR spectroscopy. Application to calmodulinIkura, Mitsuhiko; Kay, Lewis E.; Bax, AdBiochemistry (1990), 29 (19), 4659-67CODEN: BICHAW; ISSN:0006-2960.The approach is demonstrated for the protein calmodulin (16.7 kilodaltons), uniformly (∼95%) labeled with 15N and 13C. Sequential assignment of the backbone residues by std. methods was not possible because of the very narrow chem. shift distribution range of both NH and CαH protons in this largely α-helical protein. The combined use of 4 new types of heteronuclear 3-dimensional (3D) NMR spectra together with the previously described homonuclear Hartmann-Hahn-heteronuclear multiple quantum correlation 3D expt. (Marion, D. et al., 1989) can provide unambiguous sequential assignment of protein backbone resonances. Sequential connectivity is derived from 1-bond J couplings and the procedure is independent of backbone conformation. All the new 3D NMR expts. use 1H detection and rely on multiple-step magnetization transfers via well-resolved 1-bond J couplings, offering high sensitivity and requiring a total of only 9 days for the recording of all 5 3D spectra. Because the combination of 3D spectra offers at least 2 and often 3 independent pathways for detg. sequential connectivity, the new assignment procedure is easily automated. Complete assignments are reported for the 1H, C, and N backbone resonances of calmodulin, complexed with Ca. - 39Sattler, M., Schleucher, J., and Griesinger, C. Prog. Nucl. Magn. Reson. Spectrosc. 1999, 34, 93– 158
- 40Muhandiram, D. R., Yamazaki, T., Sykes, B. D., and Kay, L. E. J. Am. Chem. Soc. 1995, 117, 11536– 11544
- 41Millet, O., Muhandiram, D. R., Skrynnikov, N. R., and Kay, L. E. J. Am. Chem. Soc. 2002, 124, 6439– 6448
- 42Yu, L., Zhu, C. X., Tse-Dinh, Y. C., and Fesik, S. W. Biochemistry (Moscow) 1996, 35, 9661– 6[ACS Full Text
], [CAS], Google Scholar
42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XktVCqtLc%253D&md5=df27fe6eeb2f7baccb073af6c10d4ccbBackbone dynamics of the C-terminal domain of Escherichia coli topoisomerase I in the absence and presence of single-stranded DNAYu, Liping; Zhu, Chang-Xi; Tse-Dinh, Yuk-Ching; Fesik, Stephen W.Biochemistry (1996), 35 (30), 9661-9666CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)The backbone dynamics of the C-terminal DNA-binding domain of Escherichia coli topoisomerase I was characterized in the absence and presence of single-stranded DNA by NMR spectroscopy. 15N spin-lattice relaxation times (T1), spin-spin relaxation times (T2), and heteronuclear NOEs were detd. for the uniformly 15N-labeled protein. These data were analyzed by using the model-free formalism to derive the model-free parameters (S2, τe, and Rex) for each backbone N-H bond vector and the overall mol. rotational correlation time (τm). The mol. rotational correlation time τm was detd. to be 7.49 ns for the free and 12.7 ns for the complexed protein. Several residues were found to be much more mobile than the av., including 11 residues at the N-terminus, 2 residues at the C-terminus, and residues 25 and 31-35 which are located in a region of the protein that binds to DNA. The binding of ssDNA to the free protein caused an increase in the order parameters (S2) for a small no. of residues and a decrease in the order parameters (S2) for the majority of the residues. In particular, upon binding to ssDNA, the mobility of the 1st α-helix and the 2 β-sheets was increased, and the mobility of a few specific residues in the loops/turns was restricted. These results differed from previous studies on the backbone dynamics of mol. complexes in which reduced mobilities were typically obsd. upon ligand binding. - 43Arumugam, S., Gao, G., Patton, B. L., Semenchenko, V., Brew, K., and Van Doren, S. R. J. Mol. Biol. 2003, 327, 719– 34
- 44Yun, S., Jang, D. S., Kim, D. H., Choi, K. Y., and Lee, H. C. Biochemistry (Moscow) 2001, 40, 3967– 73[ACS Full Text
], [CAS], Google Scholar
44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXhsFKgs7s%253D&md5=89add77eb20e5813a81900488e26de1615N NMR Relaxation Studies of Backbone Dynamics in Free and Steroid-Bound Δ5-3-Ketosteroid Isomerase from Pseudomonas testosteroniYun, Sunggoo; Jang, Do Soo; Kim, Do-Hyung; Choi, Kwan Yong; Lee, Hee CheonBiochemistry (2001), 40 (13), 3967-3973CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)The backbone dynamics of Δ5-3-ketosteroid isomerase (KSI) from Pseudomonas testosteroni has been studied in free enzyme and its complex with a steroid ligand, 19-nortestosterone hemisuccinate (19-NTHS), by 15N relaxation measurements. The relaxation data were analyzed using the model-free formalism to ext. the model-free parameters (S2, τe, and Rex) and the overall rotational correlation time (τm). The rotational correlation times were 19.23 ± 0.08 and 17.08 ± 0.07 ns with the diffusion anisotropies (D‖/D.perp.) of 1.26 ± 0.03 and 1.25 ± 0.03 for the free and steroid-bound KSI, resp. The binding of 19-NTHS to free KSI causes a slight increase in the order parameters (S2) for a no. of residues, which are located mainly in helix A1 and strand B4. However, the majority of the residues exhibit reduced order parameters upon ligand binding. In particular, strands B3, B5, and B6, which have most of the residues involved in the dimer interaction, have the reduced order parameters in the steroid-bound KSI, indicating the increased high-frequency (pico- to nanosecond) motions in the intersubunit region of this homodimeric enzyme. Our results differ from those of previous studies on the backbone dynamics of monomeric proteins, in which high-frequency internal motions are typically restricted upon ligand binding. - 45Shapiro, Y. E., Kahana, E., Tugarinov, V., Liang, Z., Freed, J. H., and Meirovitch, E. Biochemistry (Moscow) 2002, 41, 6271– 81[ACS Full Text
], [CAS], Google Scholar
45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XjtVKhurg%253D&md5=cc1c988a7898412a04ac350f977d4ac4Domain Flexibility in Ligand-Free and Inhibitor-Bound Escherichia coli Adenylate Kinase Based on a Mode-Coupling Analysis of 15N Spin RelaxationShapiro, Yury E.; Kahana, Edith; Tugarinov, Vitali; Liang, Zhichun; Freed, Jack H.; Meirovitch, EvaBiochemistry (2002), 41 (20), 6271-6281CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Adenylate kinase from Escherichia coli (AKeco), consisting of a 23.6-kDa polypeptide chain folded into domains CORE, AMPbd, and LID catalyzes the reaction AMP + ATP ↔ 2ADP. The domains AMPbd and LID execute large-amplitude movements during catalysis. Backbone dynamics of ligand-free and AP5A-inhibitor-bound AKeco is studied with slowly relaxing local structure (SRLS) 15N relaxation, an approach particularly suited when the global (τm) and the local (τ) motions are likely to be coupled. For AKeco τm = 15.1 ns, whereas for AKeco*AP5A τm = 11.6 ns. The CORE domain of AKeco features an av. squared order parameter, <S2>, of 0.84 and correlation times τf = 5-130 ps. Most of the AKeco*AP5A backbone features <S2> = 0.90 and τf = 33-193 ps. These data are indicative of relative rigidity. Domains AMPbd and LID of AKeco, and loops β1/α1, α2/α3, α4/β3, α5/β4, and β8/α7 of AKeco*AP5A, feature a novel type of protein flexibility consisting of nanosecond peptide plane reorientation about the Ci-1α-Ciα axis, with correlation time τ.perp. = 5.6-11.3 ns. The other microdynamic parameters underlying this dynamic model include S2 = 0.13-0.5, τ|| on the ps time scale, and a diffusion tilt βMD ranging from 12 to 21°. For the ligand-free enzyme the τ.perp. mode was shown to represent segmental domain motion, accompanied by conformational exchange contributions Rex ≤ 4.4 s-1. Loop α4/β3 and α5/β4 dynamics in AKeco*AP5A is related to the "energetic counter-balancing of substrate binding" effect apparently driving kinase catalysis. The other flexible AKeco*AP5A loops may relate to domain motion toward product release. - 46Chervenak, M. C. and Toone, E. J. J. Am. Chem. Soc. 1994, 116, 10533– 10539
- 47Daranas, A. H., Shimizu, H., and Homans, S. W. J. Am. Chem. Soc. 2004, 126, 11870– 11876[ACS Full Text
], [CAS], Google Scholar
47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXntFWiur0%253D&md5=cbb20561d54f2b3c15f70f04f55dc368Thermodynamics of Binding of D-Galactose and Deoxy Derivatives thereof to the L-Arabinose-binding ProteinDaranas, Antonio Hernandez; Shimizu, Hiroki; Homans, Steve W.Journal of the American Chemical Society (2004), 126 (38), 11870-11876CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)We report the thermodn. of binding of D-galactose and deoxy derivs. thereof to the arabinose binding protein (ABP). The "intrinsic" (solute-solute) free energy of binding ΔG°int at 308 K for the 1-, 2-, 3-, and 6-hydroxyl groups of galactose is remarkably const. (∼-30 kJ/mol), despite the fact that each hydroxyl group subtends different nos. of hydrogen bonds in the complex. The substantially unfavorable enthalpy of binding (∼30 kJ/mol) of 1-deoxygalactose, 2-deoxygalactose, and 3-deoxygalactose in comparison with galactose, cannot be readily accounted for by differences in solvation, suggesting that solute-solute hydrogen bonds are enthalpically significantly more favorable than solute-solvent hydrogen bonds. In contrast, the substantially higher affinity for 2-deoxygalactose in comparison with either 1-deoxygalactose or 3-deoxygalactose derives from differences in the solvation free energies of the free ligands. - 48Lundquist, J. J., Debenham, S. D., and Toone, E. J. J. Org. Chem. 2000, 65, 8245– 8250
- 49Turnbull, W. B., Precious, B. L., and Homans, S. W. J. Am. Chem. Soc. 2004, 126, 1047– 1054[ACS Full Text
], [CAS], Google Scholar
49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXkt1eksg%253D%253D&md5=e453e7810b86d0e2c3a6c0299e32ee4fDissecting the Cholera Toxin-Ganglioside GM1 Interaction by Isothermal Titration CalorimetryTurnbull, W. Bruce; Precious, Bernie L.; Homans, Steve W.Journal of the American Chemical Society (2004), 126 (4), 1047-1054CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The complex of cholera toxin and ganglioside GM1 is one of the highest affinity protein-carbohydrate interactions known. Herein, the GM1 pentasaccharide is dissected into smaller fragments to det. the contribution of each of the key monosaccharide residues to the overall binding affinity. Displacement isothermal titrn. calorimetry (ITC) has allowed the measurement of all of the key thermodn. parameters for even the lowest affinity fragment ligands. Anal. of the std. free energy changes using Jencks' concept of intrinsic free energies reveals that the terminal galactose and sialic acid residues contribute 54% and 44% of the intrinsic binding energy, resp., despite the latter ligand having little appreciable affinity for the toxin. This anal. also provides an est. of 25.8 kJ mol-1 for the loss of independent translational and rotational degrees of freedom on complexation and presents evidence for an alternative binding mode for ganglioside GM2. The high affinity and selectivity of the GM1-cholera toxin interaction originates principally from the conformational preorganization of the branched pentasaccharide rather than through the effect of cooperativity, which is also reinvestigated by ITC. - 50Chang, C. A., Chen, W., and Gilson, M. K. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 1534– 1539
- 51Cabani, S., Gianni, P., Mollica, V., and Lepori, L. J. Solution Chem. 1981, 10, 563– 595
- 52Young, T., Abel, R., Kim, B., Berne, B. J., and Friesner, R. A. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 808– 813
- 53Qvist, J., Davidovic, M., Hamelberg, D., and Halle, B. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 6296– 6301
- 54Henchman, R. J. Chem. Phys. 2007, 126064504
- 55Spolar, R. S., Livingstone, J. R., and Record, M. T. J. Biochemistry (Moscow) 1992, 31, 3947– 3955[ACS Full Text
], [CAS], Google Scholar
55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XhvFSmtb4%253D&md5=18a0267fc7138758d8eff451b37006e6Use of liquid hydrocarbon and amide transfer data to estimate contributions to thermodynamic functions of protein folding from the removal of nonpolar and polar surface from waterSpolar, Ruth S.; Livingstone, Jeff R.; Record, M. Thomas, Jr.Biochemistry (1992), 31 (16), 3947-55CODEN: BICHAW; ISSN:0006-2960.This extension of the liq. hydrocarbon model seeks to quantify the thermodn. contributions to protein stability from the removal of nonpolar and polar surface from water. Thermodn. data for the transfer of hydrocarbons and org. amides from water to the pure liq. phase are analyzed to obtain contributions to the thermodn. of folding from the redn. in water-accessible surface area. Although the removal of nonpolar surface makes the dominant contribution to the std. heat capacity change of folding (ΔCfold°), here the authors show that inclusion of the contribution from removal of polar surface allows a quant. prediction of ΔCfold° within the uncertainty of the calorimetrically detd. value. Moreover, anal. of the contribution of polar surface area to the enthalpy of transfer of liq. amides provides a means of estg. the contributions from changes in nonpolar and polar surface area as well as other factors to the enthalpy of folding (ΔHfold°). In addn. to ests. of ΔHfold°, this extension of the liq. hydrocarbon model provides a thermodn. explanation for the observation that the specific enthalpy of folding (cal g-1) of a no. of globular proteins converges to a common value at approx. 383 K. Because amts. of nonpolar and polar surface area buried by these proteins upon folding are found to be linear functions of mol. mass, ests. of both ΔCfold° and ΔHfold° may be obtained given only the mol. mass of the protein of interest. Use of ΔCfold° and ΔHfold° in conjunction with a melting temp. (Tm) detd. under specified soln. conditions allows one to est. the stability (ΔGfold°) under these conditions at any temp. Calcd. values of ΔGfold° in the range 293-320 K appear generally to agree with those detd. by calorimetric and noncalorimetric methods. This anal. should provide a basis for predicting stability of proteins whose structural and thermodn. properties are similar to those in the data set used for this anal. Deviations from these predictions may be interpretable in terms of deviations from the av. properties of the proteins in the data set. - 56Prabhu, N. V. and Sharp, K. A. Annu. Rev. Phys. Chem. 2005, 56, 521– 548
- 57Sturtevant, J. M. Proc. Natl. Acad. Sci. U.S.A. 1977, 74, 2236– 2240
- 58Burkhalter, N. F., Dimick, S. M., and Toone, E. J. In Carbohydrates in Chemistry and Biology. Part I: Chemistry of Saccharides; Ernst, B., Hart, G. W., and Sinay, P., Eds.; Wiley-VCH: Weinheim, 2000; Vol. 2, pp 863− 914.
- 59Lemieux, R. U. Acc. Chem. Res. 1996, 29, 373– 380
Supporting Information
Supporting Information
ARTICLE SECTIONSComplete refs 20 and 24, backbone resonance assignments, R1, R2 and ssNOE values, 15N amide model-free parameters and TΔSconf values for backbone amide groups for rRaHBP2(D24R) and rRaHBP2(D24R) in complex with histamine. This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.