scispace - formally typeset
Search or ask a question

Showing papers on "Hydrogen bond published in 1991"


Book
17 Jul 1991
TL;DR: In this article, the van der Waals Radii cut-off criterion is used to define the strong and weak hydrogen-bond configurations, as well as the relationship between two-center and three-center hydrogen bonds.
Abstract: IA Basic Concepts.- 1 The Importance of Hydrogen Bonds.- 1.1 Historical Perspective.- 1.2 The Importance of Hydrogen Bonds in Biological Structure and Function.- 1.3 The Role of the Water Molecules.- 1.4 Significance of Small Molecule Crystal Structural Studies.- 1.5 The Structural Approach.- 2 Definitions and Concepts.- 2.1 Definition of the Hydrogen Bond - Strong and Weak Bonds.- 2.2 Hydrogen-Bond Configurations: Two- and Three-Center Hydrogen Bonds Bifurcated and Tandem Bonds.- 2.3 Hydrogen Bonds Are Very Different from Covalent Bonds.- 2.4 The van der Waals Radii Cut-Off Criterion Is Not Useful.- 2.5 The Concept of the Hydrogen-Bond Structure.- 2.6 The Importance of ? and ? Cooperativity.- 2.7 Homo-, Anti- and Heterodromic Patterns.- 2.8 Hydrogen Bond Flip-Flop Disorder: Conformational and Configurational.- 2.9 Proton-Deficient Hydrogen Bonds.- 2.10 The Excluded Region.- 2.11 The Hydrophobic Effect.- 3 Experimental Studies of Hydrogen Bonding.- 3.1 Infrared Spectroscopy and Gas Electron Diffraction.- 3.2 X-Ray and Neutron Crystal Structure Analysis.- 3.3 Treatment of Hydrogen Atoms in Neutron Diffraction Studies.- 3.4 Charge Density and Hydrogen-Bond Energies.- 3.5 Neutron Powder Diffraction.- 3.6 Solid State NMR Spectroscopy.- 4 Theoretical Calculations of Hydrogen-Bond Geometries.- 4.1 Calculating Hydrogen-Bond Geometries.- 4.2 Ab-Initio Molecular Orbital Methods.- 4.3 Application to Hydrogen-Bonded Complexes.- 4.4 Semi-Empirical Molecular Orbital Methods.- 4.5 Empirical Force Field or Molecular Mechanics Methods.- 5 Effect of Hydrogen Bonding on Molecular Structure.- IB Hydrogen-Bond Geometry.- 6 The Importance of Small Molecule Structural Studies.- 6.1 Problems Associated with the Hydrogen-Bond Geometry.- 6.2 The Hydrogen Bond Can Be Described Statistically.- 6.3 The Problems of Measuring Hydrogen-Bond Lengths and Angles in Small Molecule Crystal Structures.- 7 Metrical Aspects of Two-Center Hydrogen Bonds.- 7.1 The Metrical Properties of O-H *** O Hydrogen Bonds.- 7.1.1 Very Strong and Strong OH *** O Hydrogen Bonds Occur with Oxyanions, Acid Salts, Acid Hydrates, and Carboxylic Acids.- 7.1.2 OH *** O Hydrogen Bonds in the Ices and High Hydrates.- 7.1.3 Carbohydrates Provide the Best Data for OH ... O Hydrogen Bonds: Evidence for the Cooperative Effect.- 7.2 N-H *** O Hydrogen Bonds.- 7.3 N-H *** N Hydrogen Bonds.- 7.4 O-H *** N Hydrogen Bonds.- 7.5 Sequences in Lengths of Two-Center Hydrogen Bonds.- 7.6 H/D Isotope Effect.- 8 Metrical Aspects of Three- and Four-Center Hydrogen Bonds.- 8.1 Three-Center Hydrogen Bonds.- 8.2 Four-Center Hydrogen Bonds.- 9 Intramolecular Hydrogen Bonds.- 10 Weak Hydrogen-Bonding Interactions Formed by C-H Groups as Donors and Aromatic Rings as Acceptors.- 11 Halides and Halogen Atoms as Hydrogen-Bond Acceptors.- 12 Hydrogen-Bond Acceptor Geometries.- II Hydrogen Bonding in Small Biological Molecules.- 13 Hydrogen Bonding in Carbohydrates.- 13.1 Sugar Alcohols (Alditols) as Model Cooperative Hydrogen-Bonded Structures.- 13.2 Influence of Hydrogen Bonding on Configuration and Conformation in Cyclic Monosaccharides.- 13.3 Rules to Describe Hydrogen-Bonding Patterns in Monosaccharides.- 13.4 The Water Molecules Link Hydrogen-Bond Chains into Nets in the Hydrated Monosaccharide Crystal Structures.- 13.5 The Disaccharide Crystal Structures Provide an Important Source of Data About Hydrogen-Bonding Patterns in Polysaccharides.- 13.6 Hydrogen Bonding in the Tri- and Tetrasaccharides Is More Complex and Less Well Defined.- 13.7 The Hydrogen Bonding in Polysaccharide Fiber Structures Is Poorly Defined.- 14 Hydrogen Bonding in Amino Acids and Peptides: Predominance of Zwitterions.- 15 Purines and Pyrimidines.- 15.1 Bases Are Planar and Each Contains Several Different Hydrogen-Bonding Donor and Acceptor Groups.- 15.2 Many Tautomeric Forms Are Feasible But Not Observed.- 15.3 ?-Bond Cooperativity Enhances Hydrogen-Bonding Forces.- 15.4 General, Non-Base-Pairing Hydrogen Bonds.- 16 Base Pairing in the Purine and Pyrimidine Crystal Structures.- 16.1 Base-Pair Configurations with Purine and Pyrimidine Homo-Association.- 16.2 Base-Pair Configurations with Purine-Pyrimidine Hetero-Association: the Watson-Crick Base-Pairs.- 16.3 Base Pairs Can Combine to Form Triplets and Quadruplets.- 17 Hydrogen Bonding in the Crystal Structures of the Nucleosides and Nucleotides.- 17.1 Conformational and Hydrogen-Bonding Characteristics of the Nucleosides and Nucleotides.- 17.2 A Selection of Cyclic Hydrogen-Bonding Patterns Formed in Nucleoside and Nucleotide Crystal Structures.- 17.3 General Hydrogen-Bonding Patterns in Nucleoside and Nucleotide Crystal Structures.- III Hydrogen Bonding in Biological Macromolecules.- 18 O-H *** O Hydrogen Bonding in Crystal Structures of Cyclic and Linear Oligoamyloses: Cyclodextrins, Maltotriose, and Maltohexaose.- 18.1 The Cyclodextrins and Their Inclusion Complexes.- 18.2 Crystal Packing Patterns of Cyclodextrins Are Determined by Hydrogen Bonding.- 18.3 Cyclodextrins as Model Compounds to Study Hydrogen-Bonding Networks.- 18.4 Cooperative, Homodromic, and Antidromic Hydrogen-Bonding Patterns in the ?-Cyclodextrin Hydrates.- 18.5 Homodromic and Antidromic O-H *** O Hydrogen-Bonding Systems Analyzed Theoretically.- 18.6 Intramolecular Hydrogen Bonds in the ?-Cyclodextrin Molecule are Variable - the Induced-Fit Hypothesis.- 18.7 Flip-Flop Hydrogen Bonds in ?-Cyclodextrin * 11 H2O.- 18.8 From Flip-Flop Disorder to Ordered Homodromic Arrangements at Low lbmperature: The Importance of the Cooperative Effect.- 18.9 Maltohexaose Polyiodide and Maltotriose - Double and Single Left-Handed Helices With and Without Intramolecular O(2) *** O(3?) Hydrogen Bonds.- 19 Hydrogen Bonding in Proteins.- 19.1 Geometry of Secondary-Structure Elements: Helix, Pleated Sheet, and Turn.- 19.2 Hydrogen-Bond Analysis in Protein Crystal Structures.- 19.3 Hydrogen-Bonding Patterns in the Secondary Structure Elements.- 19.4 Hydrogen-Bonding Patterns Involving Side-Chains.- 19.5 Internal Water Molecules as Integral Part of Protein Structures.- 19.6 Metrical Analysis of Hydrogen Bonds in Proteins.- 19.7 Nonsecondary-Structure Hydrogen-Bond Geometry Between Main-Chains, Side-Chains and Water Molecules.- 19.8 Three-Center (Bifurcated) Bonds in Proteins.- 19.9 Neutron Diffraction Studies on Proteins Give Insight into Local Hydrogen-Bonding Flexibility.- 19.10 Site-Directed Mutagenesis Gives New Insight into Protein Thermal Stability and Strength of Hydrogen Bonds.- 20 The Role of Hydrogen Bonding in the Structure and Function of the Nucleic Acids.- 20.1 Hydrogen Bonding in Nucleic Acids is Essential for Life.- 20.2 The Structure of DNA and RNA Double Helices is Determined by Watson-Crick Base-Pair Geometry.- 20.3 Systematic and Accidental Base-Pair Mismatches: "Wobbling" and Mutations.- 20.4 Noncomplementary Base Pairs Have a Structural Role in tRNA.- 20.5 Homopolynucleotide Complexes Are Stabilized by a Variety of Base-Base Hydrogen Bonds - Three-Center (Bifurcated) Hydrogen Bonds in A-Tracts.- 20.6 Specific Protein-Nucleic Acid Recognition Involves Hydrogen Bonding.- IV Hydrogen Bonding by the Water Molecule.- 21 Hydrogen-Bonding Patterns in Water, Ices, the Hydrate Inclusion Compounds, and the Hydrate Layer Structures.- 21.1 Liquid Water and the Ices.- 21.2 The Hydrate Inclusion Compounds.- 21.3 Hydrate Layer Structures.- 22 Hydrates of Small Biological Molecules: Carbohydrates, Amino Acids, Peptides, Purines, Pyrimidines, Nucleosides and Nucleotides.- 23 Hydration of Proteins.- 23.1 Characterization of "Bound Water" at Protein Surfaces - the First Hydration Shell.- 23.2 Sites of Hydration in Proteins.- 23.3 Metrics of Water Hydrogen Bonding to Proteins.- 23.4 Ordered Water Molecules at Protein Surfaces - Clusters and Pentagons.- 24 Hydration of Nucleic Acids.- 24.1 Two Water Layers Around the DNA Double Helix.- 24.2 Crystallographically Determined Hydration Sites in A-, B-, Z-DNA. A Statistical Analysis.- 24.3 Hydration Motifs in Double Helical Nucleic Acids.- 24.3.1 Sequence-Independent Motifs.- 24.3.2 Sequence-Dependent Motifs.- 24.4 DNA Hydration and Structural Transitions Are Correlated: Some Hypotheses.- 25 The Role of Three-Center Hydrogen Bonds in the Dynamics of Hydration and of Structure Transition.- References.- Refcodes.

2,739 citations




ComponentDOI
TL;DR: The 3D structure of the Maltodextrin-binding protein (Mr = 40,622) with bound maltose has been obtained by crystallographic analysis at 2.8-A resolution as discussed by the authors.

478 citations


Journal ArticleDOI
01 Jan 1991-Nature
TL;DR: Inomata et al. as discussed by the authors reported a phase transition in an interpenetrating polymer network of poly(acrylamide) and poly (acrylic acid) that is controlled by cooperative 'zipping' interactions between the molecules which result from hydrogen bonding.
Abstract: INTERACTIONS between macromolecules fall into four categories: ionic, hydrophobic, van der Waals and hydrogen bonding. Phase transitions in polymer gels provide a means of studying these interactions. Many gels will undergo reversible, discontinuous volume changes in response to changes in, for example, temperature, gel composition or light irradiation1–5. These transitions result from the competition between repulsive intermolecular forces, usually electrostatic in nature, that act to expand the polymer network, and an attractive force that acts to shrink it. Volume transitions in gels have been observed that are driven by all of the above-mentioned forces except hydrogen bonding (ref 6–10; T.T. et al, unpublished data; H. Inomata et al., personal communication). Here we report on a phase transition in an interpenetrating polymer network of poly(acrylamide) and poly(acrylic acid) that completes this picture—it is controlled by cooperative 'zipping' interactions between the molecules which result from hydrogen bonding. Cooperativity is an essential feature of the interactions, in that independent hydrogen bonds would not provide a sufficient driving force for the transition. A further novel characteristic of this phase transition is that the swelling (in water) is induced by an increase rather than a decrease in temperature.

463 citations



Journal ArticleDOI
01 Jan 1991-Polymer
TL;DR: In this article, the Young's modulus along the chain axis was calculated for cellulose crystalline forms I and II (native and regenerated celluloses, respectively), and anisotropy of the modulus and linear compressibility in the planes perpendicular to the molecular chain axis were also calculated.

443 citations


Journal ArticleDOI
TL;DR: The X-ray crystal structure of recombinant wild-type azurin from Pseudomonas aeruginosa was determined by difference Fourier techniques using phases derived from the structure of the mutant His35Leu, which involves a concomitant Pro36-Gly37 main-chain peptide bond flip.

437 citations



Journal ArticleDOI
TL;DR: The crystal structure of Cd5,Zn2-metallothionein from rat liver has been refined and the previously reported metal positions and the protein model derived therefrom are incorrect.

342 citations


Journal ArticleDOI
01 Feb 1991-Nature
TL;DR: In this paper, the authors performed X-ray diffraction studies on crystalline Na4[calix[4]arene sulphonate and found evidence for hydrogen bonding between water and the aromatic π electrons in the solid state.
Abstract: THEinteraction of water with aromatic moieties is of importance in biological systems, as most encounter an aqueous environment during their normal functions. For example, common constituents of globular proteins such as phenylalanine, tryptophan and tyrosine possess aromatic side-chains1 that may encounter water molecules inside the protein structure2. As a model for hydrogen-bonding interactions with aromatics, we have performed X-ray diffraction studies on crystalline Na4[calix[4]arene sulphonate].13.5H2O. Calixarene molecules3,4contain hydrophobic cavities comprised of aromatic groups, rimmed, in the case of the water-soluble sulphonates (R = -SO3Na), by hydrophilic groups5,6. In the absence of a hydrophobic guest, the cavity invariably contains a water molecule. The low-temperature X-ray crystal structure of this compound (see Table 1 and Fig. 1) shows direct evidence for hydrogen bonding between water and the aromatic π electrons in the solid state.

Journal ArticleDOI
10 Oct 1991-Nature
TL;DR: This structure emphasizes the importance of specific hydrogen bonding between not only the nucleotide bases, but also the ribose hydroxyls, phosphate oxygens and tightly bound waters in stabilization of the intramolecular and intermolecular structures of double helical RNA.
Abstract: The crystal structure of the RNA dodecamer duplex (r-GGACUUCGGUCC)2 has been determined. The dodecamers stack end-to-end in the crystal, simulating infinite A-form helices with only a break in the phosphodiester chain. These infinite helices are held together in the crystal by hydrogen bonding between ribose hydroxyl groups and a variety of donors and acceptors. The four noncomplementary nucleotides in the middle of the sequence did not form an internal loop, but rather a highly regular double-helix incorporating the non-Watson-Crick base pairs, G.U and U.C. This is the first direct observation of a U.C (or T.C) base pair in a crystal structure. The U.C pairs each form only a single base-base hydrogen bond, but are stabilized by a water molecule which bridges between the ring nitrogens and by four waters in the major groove which link the bases and phosphates. The lack of distortion introduced in the double helix by the U.C mismatch may explain its low efficiency of repair in DNA. The G.U wobble pair is also stabilized by a minor-groove water which bridges between the unpaired guanine amino and the ribose hydroxyl of the uracil. This structure emphasizes the importance of specific hydrogen bonding between not only the nucleotide bases, but also the ribose hydroxyls, phosphate oxygens and tightly bound waters in stabilization of the intramolecular and intermolecular structures of double helical RNA.


Journal ArticleDOI
TL;DR: In this paper, the crystal structure of Cd5,Zn2-metallothionein from rat liver has been refined at 2.0Aresolution of a R-value of 0.176 for all observed data.

Journal ArticleDOI
TL;DR: It is suggested that the two pK a, values of the active site aspartyls arise from a structure not unlike that in maleic acid with a hydrogen‐bonded intermediate species and a dicarboxylate characterised by electrostatic repulsions between the two negatively charged groups.

Journal ArticleDOI
TL;DR: The three-dimensional structure of cyclosporin A bound to cyclophilin was calculated with the distance geometry program DISMAN, and the structures resulting from 181 converged calculations were energy refined with the program FANTOM.
Abstract: Cyclosporin A bound to the presumed receptor protein cyclophilin was studied in aqueous solution at pH 6.0 by nuclear magnetic resonance spectroscopy using uniform {sup 15}N- or {sup 13}C-labeling of cyclosporin A and heteronuclear spectral editing techniques. With an input of 108 intramolecular NOEs and four vicinal {sup 3}J{sub HN{alpha}} coupling constants, the three-dimensional structure of cyclosporin A bound to cyclophilin was calculated with the distance geometry program DISMAN, and the structures resulting from 181 converged calculations were energy refined with the program FANTOM. A group of 120 conformers was selected on the basis of the residual constraint violations and energy criteria to represent the solution structure. The average of the pairwise root-mean-square distances calculated for the backbone atoms of the 120 structures was 0.58 {angstrom}. The structure represents a novel conformation of cyclosporin A, for which the backbone conformation is significantly different from the previously reported structures in single crystals and in chloroform solution. The structure has all peptide bonds in the trans form, contains no elements of regular secondary structure and no intramolecular hydrogen bonds, and exposes nearly all polar groups to its environment. The root-mean-square distance between the backbone atoms of the crystal structure of cyclosporin Amore » and the mean of the 120 conformers representing the NMR structure of cyclosporin A bound to cyclophilin is 2.5 {angstrom}.« less

Journal ArticleDOI
TL;DR: In this article, the polarity of mixtures of water and cyanomethane was established from measurements of Kamlet and Taft's (KT's) π* their hydrogen bond acceptance ability from measurements, β, and α, with suitable solvatochromic indicators.
Abstract: The polarity of mixtures of water and cyanomethane was established from measurements of Kamlet and Taft's (KT's) π* their hydrogen bond acceptance ability from measurements of KT's β, and their hydrogen bond donation ability from measurements of KT's α, with suitable solvatochromic indicators. The structure of these mixtures was explored by use of thermodynamic data from the literature and the application of the quasi-lattice quasi-chemical (QLQC) and the inverse Kirkwood-Buff integral (IKBI) methods

Journal ArticleDOI
01 Feb 1991-Proteins
TL;DR: A set of 85 high‐resolution protein structures is surveyed in order to evaluate the prevalence and geometry of sulfur‐containing hydrogen bonds and this information should be of interest to experimentalists and theoreticians intersted in protein structure and protein engineering.
Abstract: Intrachain hydrogen bonds are a hallmark of globular proteins. Traditionally, these involve oxygen and nitrogen atoms. The electronic structure of sulfur is compatible with hydrogen bond formation as well. We surveyed a set of 85 high-resolution protein structures in order to evaluate the prevalence and geometry of sulfur-containing hydrogen bonds. This information should be of interest to experimentalists and theoreticians interested in protein structure and protein engineering.


Journal ArticleDOI
TL;DR: The resulting atomic co-ordinates indicate that the NAM residue in site D is distorted from the full 4C1 chair conformation to one in which the ring atoms C-1, C-2, O-5 and C-5 are approximately coplanar, and the hydroxymethyl group is positioned axially (a conformation best described as a sofa).


Journal ArticleDOI
TL;DR: X-ray crystal structures have been determined for complexes of cytochrome P-450CAM with the substrates camphane, adamantane, and thiocamphor, and the most unexpected orientation in the active site was given, which may correspond to a nonproductive complex.
Abstract: X-ray crystal structures have been determined for complexes of cytochrome P-450CAM with the substrates camphane, adamantane, and thiocamphor. Unlike the natural substrate camphor, which hydrogen bonds to Tyr96 and is metabolized to a single product, camphane, adamantane and thiocamphor do not hydrogen bond to the enzyme and all are hydroxylated at multiple positions. Evidently the lack of a substrate-enzyme hydrogen bond allows substrates greater mobility in the active site, explaining this lower regiospecificity of metabolism as well as the inability of these substrates to displace the distal ligand to the heme iron. Tyr96 is a ligand, via its carbonyl oxygen atom, to a cation that is thought to stabilize the camphor-P-450CAM complex [Poulos, T. L., Finzel, B. C., & Howard, A. J. (1987) J. Mol. Biol. 195, 687-700]. The occupancy and temperature factor of the cationic site are lower and higher, respectively, in the presence of the non-hydrogen-bonding substrates investigated here than in the presence of camphor, underscoring the relationship between cation and substrate binding. Thiocamphor gave the most unexpected orientation in the active site of any of the substrates we have investigated to date. The orientation of thiocamphor is quite different from that of camphor. That is, carbons 5 and 6, at which thiocamphor is primarily hydroxylated [Atkins, W. M., & Sligar, S. G. (1988) J. Biol. Chem. 263, 18842-18849], are positioned near Tyr96 rather than near the heme iron. Therefore, the crystallographically observed thiocamphor-P-450CAM structure may correspond to a nonproductive complex. Disordered solvent has been identified in the active site in the presence of uncoupling substrates that channel reducing equivalents away from substrate hydroxylation toward hydrogen peroxide and/or "excess" water production. A buried solvent molecule has also been identified, which may promote uncoupling by moving from an internal location to the active site in the presence of highly mobile substrates.


Journal ArticleDOI
TL;DR: The three-dimensional structure of TIM complexed with a reactive intermediate analogue, phosphoglycolohydroxamate (PGH), is solved at 1.9-A resolution and the structure suggests that His-95 is neutral rather than cationic in the ground state and therefore would have to function as an imidazole acid instead of the usualImidazolium.
Abstract: The glycolytic enzyme triosephosphate isomerase (TIM) catalyzes the interconversion of the three-carbon sugars dihydroxyacetone phosphate (DHAP) and D-glyceraldehyde 3-phosphate (GAP) at a rate limited by the diffusion of substrate to the enzyme. We have solved the three-dimensional structure of TIM complexed with a reactive intermediate analogue, phosphoglycolohydroxamate (PGH), at 1.9-A resolution and have refined the structure to an R-factor of 18%. Analysis of the refined structure reveals the geometry of the active-site residues and the interactions they make with the inhibitor and, by analogy, the substrates. The structure is consistent with an acid-base mechanism in which the carboxylate of Glu-165 abstracts a proton from carbon while His-95 donates a proton to oxygen to form an enediol (or enediolate) intermediate. The conformation of the bound substrate stereoelectronically favors proton transfer from substrate carbon to the syn orbital of Glu-165. The crystal structure suggests that His-95 is neutral rather than cationic in the ground state and therefore would have to function as an imidazole acid instead of the usual imidazolium. Lys-12 is oriented so as to polarize the substrate oxygens by hydrogen bonding and/or electrostatic interaction, providing stabilization for the charged transition state. Asn-10 may play a similar role.

Journal ArticleDOI
TL;DR: Detailed analyses of the unfolding pathways at high temperature have revealed that the formation or disappearance of main-chain helical hydrogen bonds occurs frequently through an alpha in equilibrium with 3(10) inilibrium with no hydrogen bond sequence.
Abstract: Molecular dynamics simulations of the S-peptide analogue AETAAAKFLREHMDS have been conducted in aqueous solution for 300 ps at 278 K and for 500 ps in two different runs at 358 K. The results show agreement with experimental observations in that at low temperature, 5 degrees C, the helix is stable, while unfolding is observed at 85 degrees C. In the low-temperature simulation a solvent-separated ion pair was formed between Glu-2 and Arg-10, and the side chain of His-12 reoriented toward the C-terminal end of the alpha-helix. Detailed analyses of the unfolding pathways at high temperature have also revealed that the formation or disappearance of main-chain helical hydrogen bonds occurs frequently through an alpha in equilibrium with 3(10) in equilibrium with no hydrogen bond sequence.

Journal ArticleDOI
TL;DR: In this paper, the potential functions for simple amides, several peptides and a small protein have been worked out for the MM3 force field by fitting the vibrational spectra of the simple amide and examining more carefully electrostatic interactions including charge-charge and charge-dipole interactions.
Abstract: The potential functions for simple amides, several peptides and a small protein have been worked out for the MM3 force field. Structures and energies were fit as previously with MM2, but additionally, we fit the vibrational spectra of the simple amides (average rms error over four compounds, 34 cm−1), and examined more carefully electrostatic interactions, including charge-charge and charge-dipole interactions. The parameters were obtained and tested by examining four simple amides, five electrostatic model complexes, two dipeptides, six crystalline cyclic peptides, and the protein Crambin. The average root-mean-square deviation from the X-ray structures for the six cyclic peptide crystals was only 0.10 A for the nonhydrogen atomic positions, and 0.011 A, 1.0°, and 4.9° for bond lengths, bond angles, and torsional angles, respectively. The parameter set was then further tested by minimizing the high resolution crystal structure of the hydrophobic protein Crambin. The resultant root-mean-square deviations for the non-hydrogen atomic data, in the presence of the crystal lattice, are 0.22 A, 0.023 A, 2.0°, and 6.4° for coordinates, bond lengths, bond angles, and torsional angles, respectively.

Journal ArticleDOI
TL;DR: Extension of the approach will allow, in those cases where there is good complementarity between ligand and receptor, the prediction of approximate peptide-peptide binding constants in aqueous solution.
Abstract: An expression is presented for the estimation of approximate binding constants for bimolecular associations in solution. The consequences of the approach have been examined for the bimolecular association of two peptide components in aqueous solution: specifically for the binding of two vancomycin group antibiotics, vancomycin itself and ristocetin A, to the peptide cell wall analogue ^V-Ac-D-AIa-D-AIa and related ligands. Uncertainties in the treatment are relatively large, but the physical insights gained into the binding process (in part with the aid of calorimetric data obtained by others) are enlightening. We conclude that for amide-amide hydrogen bond formation in aqueous solution at room temperature, the intrinsic binding energy is ca. 24 kJ mol"1 (an intrinsic binding constant of ca. 104); this process is almost completely driven by a favorable entropy change associated with the release of water molecules from the amide NH and CO groups involved in hydrogen bond formation. The bimolecular association of N-Ac-D-AIa-D-Ala with ristocetin A has a remarkably small entropy change at 298 K (TAS = 3 ± 1.5 kJ mol"1)- We conclude that the release of water from polar and hydrocarbon groups involved in the binding almost exactly compensates for (i) the unfavorable entropy change due to the freezing out of four rotors of /V-Ac-D-AIa-D-AIa upon binding and (ii) the unfavorable entropy change of a bimolecular association. A crude quantitation of these effects is presented. We also present an estimate of the increase in translational plus rotational free energy, as a function of the ligand mass, occurring when a ligand binds to a larger receptor. This quantity, fundamental to all binding processes, is relatively insensitive to the shape of the ligand. Extension of the approach will allow, in those cases where there is good complementarity between ligand and receptor, the prediction of approximate peptide-peptide binding constants in aqueous solution.



Journal ArticleDOI
TL;DR: BINAP-Ru(II) catalyzed hydrogenation of β-substituted (E )-β-(acylamino)acrylic acids allows efficient enantioselective synthesis of βamino acids as mentioned in this paper.
Abstract: BINAP—Ru(II) catalyzed hydrogenation of β-substituted ( E )-β-(acylamino)acrylic acids allows efficient enantioselective synthesis of β-amino acids. The Z double bond isomers which possess an intramolecular hydrogen bond between amide and ester groups are more reactive but are hydrogenated with poor enantioselectivity. BINAP—Rh(I) complexes afford only moderate stereoselectivity with the opposite sense of enantioselection.