共查询到20条相似文献,搜索用时 0 毫秒
1.
Resolution of mandelic acid with (-)-(1R,2S)-ephedrine in water and ethanol produces intermediate diastereomeric salts with greatly disparate solubilities and melting points. Single crystal X-ray analysis of the less (L) and more (M) soluble (-)-ephedrinium mandelates (I, II) shows crystal structures which are isosteric, each crystallizing in the monoclinic system, space group C2. Protonated ephedrines occupy the same relative positions in the L- and M-salts, and mandelates are in the same general locations. Hydrogen bonds link alternating protonated ephedrine nitrogens and mandelate carboxylate oxygens in each salt forming columns of ions. The helical H-bonded chain winds down the crystallographic 2-fold screw axis. Additional H-bonds form between 2-fold related mandelates in the L-salt. Mixed crystals, containing both mandelate isomers, (2R)- and (2S)-mandelates, are obtained from the resolving system partly depleted of the L-salt. A specimen with nearly equal amounts of the mandelates (III) is also isosteric with the commensurate structures. I (294K), L-salt: a = 18.160(7), b = 6.538(2), c = 13.898(4) A, beta = 92.02(3) degrees, V = 1649.1(9) A3; IIa (294K), M-salt: a = 17.978(11), b = 7.164(4), c = 13.574(6)A, beta = 96.41(4) degrees, V = 1737.3(16) A3; IIb (223K), M-salt: a = 17.805(8), b = 7.115(2), c = 13.50(5) A, beta = 96.89(3) degrees, V = 1697.9(15) A3; III (294K), mixed-salt: a = 18.184(22), b = 6.792(7), c = 13.808(19) A, beta = 93.74(10) degrees, V = 1701.7(35) A3. 相似文献
2.
The crystal packing surfaces comprising protein-RNA interactions were analyzed for 50 RNA-protein crystal structures in the Protein Data Bank database. Protein-RNA crystal contacts, which represent nonspecific protein-RNA interfaces, were investigated for their amino acid propensities, hydrogen bond patterns, and backbone and side chain interactions. When compared to biologically relevant interactions, the protein-RNA crystal contacts exhibit similarities as well as differences with respect to the principles of protein-RNA interactions. Similar to what was observed at cognate protein-RNA interfaces, positively charged amino acids have high propensities at noncognate protein-RNA interfaces and preferentially form hydrogen bonds with RNA phosphate groups. In contrast, nonpolar residues are less frequently associated with noncognate interactions. These results highlight the important roles of both electrostatic and hydrogen bonding interactions, facilitated by positively charged amino acids, in mediating both specific and nonspecific protein-RNA interactions. 相似文献
3.
Isolde Le Trong Stefanie Freitag Lisa A. Klumb Vano Chu Patrick S. Stayton Ronald E. Stenkamp 《Acta Crystallographica. Section D, Structural Biology》2003,59(9):1567-1573
An elaborate hydrogen‐bonding network contributes to the tight binding of biotin to streptavidin. The specific energetic contributions of hydrogen bonds to the biotin ureido oxygen have previously been investigated by mapping the equilibrium and activation thermodynamic signatures of N23A, N23E, S27A, Y43A and Y43F site‐directed mutants [Klumb et al. (1998), Biochemistry, 37 , 7657–7663]. The crystal structures of these variants in the unbound and biotin‐bound states provide structural insight into the energetic alterations and are described here. High (1.5–2.2 Å) to atomic resolution (1.14 Å) structures were obtained and structural models were refined to R values ranging from 0.12 to 0.20. The overall folding of streptavidin as described previously has not changed in any of the mutant structures. Major deviations such as side‐chain shifts of residues in the binding site are observed only for the N23A and Y43A mutations. In none of the mutants is a systematic shift of biotin observed when one of the hydrogen‐bonding partners to the ureido oxygen of biotin is removed. Recent thermodynamic studies report increases of ΔΔG° of 5.0–14.6 kJ mol−1 for these mutants with respect to the wild‐type protein. The decreasing stabilities of the complexes of the mutants are discussed in terms of their structures. 相似文献
4.
Hydrogen bonding and π‐π interactions take special part in the enantioselectivity task. In this regard, because of having both hydrogen acceptor and hydrogen donor groups, melamine derivatives become more of an issue for enantioselectivity. In the light of such information, triazine‐based chiral, fluorescence active novel thiazole derivatives L1 and L2 were designed and synthesized from (S)‐(?)‐2‐amino‐1‐butanol and (1S,2R)‐(+)‐2‐amino‐1,2‐diphenylethanol. The structural establishment of these compounds was made by spectroscopic methods such as FTIR, 1H, and 13C NMR. While the solution of these compounds in DMSO did not show any fluorescence emission, it was observed that the emission increased 44‐fold for L1 and 55‐fold for L2 in 95% water, similar to the aggregation‐induced emission (AIE) characterized compounds. In this regard, enantioselective capabilities of these compounds against carboxylic acids were tested, and in experiments carried out at a ratio of 40/60 DMSO/H2O, it was determined that R‐2ClMA increased the fluorescence emission of L1 chiral receptor by 2.59 times compared to S‐isomer. 相似文献
5.
Translocation of negatively charged ions across cell membranes by ion pumps raises the question as to how protein interactions control the location and dynamics of the ion. Here we address this question by performing extensive molecular dynamics simulations of wild type and mutant halorhodopsin, a seven-helical transmembrane protein that translocates chloride ions upon light absorption. We find that inter-helical hydrogen bonds mediated by a key arginine group largely govern the dynamics of the protein and water groups coordinating the chloride ion. 相似文献
6.
7.
Koellner G Steiner T Millard CB Silman I Sussman JL 《Journal of molecular biology》2002,320(4):721-725
The crystal structure of acetylcholinesterase from Torpedo californica complexed with the uncharged inhibitor, PEG-SH-350 (containing mainly heptameric polyethylene glycol with a terminal thiol group) is determined at 2.3 A resolution. This is an untypical acetylcholinesterase inhibitor, since it lacks the cationic moiety typical of the substrate (acetylcholine). In the crystal structure, the elongated ligand extends along the whole of the deep and narrow active-site gorge, with the terminal thiol group bound near the bottom, close to the catalytic site. Unexpectedly, the cation-binding site (formed by the faces of aromatic side-chains) is occupied by CH(2) groups of the inhibitor, which are engaged in C-H...pi interactions that structurally mimic the cation-pi interactions made by the choline moiety of acetylcholine. In addition, the PEG-SH molecule makes numerous other weak but specific interactions of the C-H...O and C-H...pi types. 相似文献
8.
Hydrogen Evolution Reaction: A New Platinum‐Like Efficient Electrocatalyst for Hydrogen Evolution Reaction at All pH: Single‐Crystal Metallic Interweaved V8C7 Networks (Adv. Energy Mater. 23/2018)
下载免费PDF全文

Haitao Xu Jing Wan Huijuan Zhang Ling Fang Li Liu Zhengyong Huang Jian Li Xiao Gu Yu Wang 《Liver Transplantation》2018,8(23)
9.
Haitao Xu Jing Wan Huijuan Zhang Ling Fang Li Liu Zhengyong Huang Jian Li Xiao Gu Yu Wang 《Liver Transplantation》2018,8(23)
Exploring low‐cost hydrogen evolution reaction (HER) catalysts with remarkable activity over wide pH range (0–14) still remains an enormous challenge. Herein, for the first time, a novel platinum‐like, double‐deck carbon coated V8C7 networks with the highly active (110) facet exposed as a new efficient HER electrocatalyst is reported. The single‐crystal interweaved V8C7 networks are designed and fabricated based on a low crystal‐mismatch strategy and confinement effect of double‐deck carbon coating. In addition, electrochemical tests and theoretical simulation indicate that the metallic character of V8C7, high‐activity of exposed facet, and low barrier energy for water dissociation can contribute to highly catalytic activity of HER. Impressively, the HER performances of the interweaved V8C7 networks can be comparable to those of Pt at an all‐pH environment, with Tafel slopes of 44, 64, and 34.5 mV dec?1and overpotential of 47, 77, and 38 mV at ?10 mA cm?2 in 1 m KOH, 0.1 m phosphate buffer, and 0.5 m H2SO4, respectively. This work provides a blueprint for exploring new‐type platinum‐like catalysts for various energy conversion systems. 相似文献
10.
Three new coordination complexes, [Cu(L1)(H2O)] (1), [Ni(L2)2]·CH3CN (2) and [Co(HL3)(L3)] (3) [where H2L1, N,N′-bis(3-methoxysalicylidenimino)-1,3-diamino-propane; HL2, 2-((E)-(1,3-dihydroxy-2-methylpropan-2-ylimino)methyl)phenol; H2L3, 2-((E)-(2-hydroxyethylimino)methyl)-4-bromophenol] have been synthesized and systematically characterized by elemental analyses, FT-IR, electronic spectroscopy, cyclic voltammetry and thermogravimetric analyses. Single crystal X-ray diffraction studies confirm that the metal center in complex 1 has distorted square-pyramidal geometry while it is distorted octahedral in the other two complexes. In all the complexes O-H?O hydrogen bondings assemble the molecular units leading to ordered supramolecular architectures. While both complexes 1 and 2 form infinite one-dimensional arrays through the self organisation of hydrogen bonded ring motifs, complex 3 is a unique star-shaped cyclic hexamer generated through intermolecular hydrogen bonding. 相似文献
11.
Loops are integral components of protein structures, providing links between elements of secondary structure, and in many cases contributing to catalytic and binding sites. The conformations of short loops are now understood to depend primarily on their amino acid sequences. In contrast, the structural determinants of longer loops involve hydrogen-bonding and packing interactions within the loop and with other parts of the protein. By searching solved protein structures for regions similar in main chain conformation to the antigen-binding loops in immunoglobulins, we identified medium-sized loops of similar structure in unrelated proteins, and compared the determinants of their conformations. For loops that form compact substructures the major determinant of the conformation is the formation of hydrogen bonds to inward-pointing main chain atoms. For loops that have more extended conformations, the major determinant of their structure is the packing of a particular residue or residues against the rest of the protein. The following picture emerges: Medium-sized loops of similar conformation are stabilized by similar interactions. The groups that interact with the loop have very similar spatial dispositions with respect to the loop. However, the residues that provide these interactions may arise from dissimilar parts of the protein: The conformation of the loop requires certain interactions that the protein may provide in a variety of ways. 相似文献
12.
《Biophysical journal》2022,121(11):2168-2179
Cysteine residues perform a dual role in mammalian hairs. The majority help stabilize the overall assembly of keratins and their associated proteins, but a proportion of inter-molecular disulfide bonds are assumed to be associated with hair mechanical flexibility. Hair cortical microstructure is hierarchical, with a complex macro-molecular organization resulting in arrays of intermediate filaments at a scale of micrometres. Intermolecular disulfide bonds occur within filaments and between them and the surrounding matrix. Wool fibers provide a good model for studying various contributions of differently situated disulfide bonds to fiber mechanics. Within this context, it is not known if all intermolecular disulfide bonds contribute equally, and, if not, then do the disproportionally involved cysteine residues occur at common locations on proteins? In this study, fibers from Romney sheep were subjected to stretching or to their breaking point under wet or dry conditions to detect, through labeling, disulfide bonds that were broken more often than randomly. We found that some cysteines were labeled more often than randomly and that these vary with fiber water content (water disrupts protein-protein hydrogen bonds). Many of the identified cysteine residues were located close to the terminal ends of keratins (head or tail domains) and keratin-associated proteins. Some cysteines in the head and tail domains of type II keratin K85 were labeled in all experimental conditions. When inter-protein hydrogen bonds were disrupted under wet conditions, disulfide labeling occurred in the head domains of type II keratins, likely affecting keratin-keratin-associated protein interactions, and tail domains of the type I keratins, likely affecting keratin-keratin interactions. In contrast, in dry fibers (containing more protein-protein hydrogen bonding), disulfide labeling was also observed in the central domains of affected keratins. This central “rod” region is associated with keratin-keratin interactions between anti-parallel heterodimers in the tetramer of the intermediate filament. 相似文献
13.
Effect of a single aspartate on helix stability at different positions in a neutral alanine-based peptide. 总被引:1,自引:7,他引:1
下载免费PDF全文

B. M. Huyghues-Despointes J. M. Scholtz R. L. Baldwin 《Protein science : a publication of the Protein Society》1993,2(10):1604-1611
A single aspartate residue has been placed at various positions in individual peptides for which the alanine-based reference peptide is electrically neutral, and the helix contents of the peptides have been measured by circular dichroism. The dependence of peptide helix content on aspartate position has been used to determine the helix propensity (s-value). Both the charged (Asp-) and uncharged (Asp0) forms of the aspartate residue are strong helix breakers and have identical s-values of 0.29 at 0 degree C. The interaction of Asp- with the helix dipole affects helix stability at positions throughout the helix, not only near the N-terminus, where the interaction is helix stabilizing, and the C-terminus, where it is destabilizing. Comparison of the helix contents at acidic pH (Asp0) and at neutral pH (Asp-) shows that the charge-helix dipole interaction is screened slowly with increasing NaCl concentration, and screening is not complete even at 4.8 M NaCl. Lastly, a helix-stabilizing hydrogen-bond interaction between glutamine and aspartate (spacing i, i + 4) has been found. This side-chain interaction is specific for both the orientation and spacing of the glutamine and aspartate residues and is resistant to screening by NaCl. 相似文献
14.
Scott A. Lee Nong-Jian Tao Allan Rupprecht 《Journal of biomolecular structure & dynamics》2013,31(11):1337-1342
Raman spectroscopy is used to probe the nature of the hydrogen bonds which hold the water of hydration to DNA. The ~ 3450?cm?1 molecular O–H stretching mode shows that the first six water molecules per base pair of the primary hydration shell are very strongly bound to the DNA. The observed shift in the peak position of this mode permits a determination of the length of the hydrogen bonds for these water molecules. These hydrogen bonds appear to be about 0.3?Å shorter than the hydrogen bonds in bulk water. The linewidth of this mode shows no significant changes above water contents of about 15 water molecules per base pair. This technique of using a vibrational spectroscopy to obtain structural information about the hydration shells of DNA could be used to study the hydration shells of other biomolecules. 相似文献
15.
We have analyzed the interstitial water (ISW) structures in 1500 protein crystal structures deposited in the Protein Data Bank that have greater than 1.5 Å resolution with less than 90% sequence similarity with each other. We observed varieties of polygonal water structures composed of three to eight water molecules. These polygons may represent the time‐ and space‐averaged structures of “stable” water oligomers present in liquid water, and their presence as well as relative population may be relevant in understanding physical properties of liquid water at a given temperature. On an average, 13% of ISWs are localized enough to be visible by X‐ray diffraction. Of those, averages of 78% are water molecules in the first water layer on the protein surface. Of the localized ISWs beyond the first layer, almost half of them form water polygons such as trigons, tetragons, as well as expected pentagons, hexagons, higher polygons, partial dodecahedrons, and disordered networks. Most of the octagons and nanogons are formed by fusion of smaller polygons. The trigons are most commonly observed. We suggest that our observation provides an experimental basis for including these water polygon structures in correlating and predicting various water properties in liquid state. 相似文献
16.
13C‐nmr chemical shift tensor components are reported for a 13C‐labeled Gly1 amide carbonyl carbon of a glycylglycine (Gly1Gly2) single crystal, a GlyGly · HNO3 single crystal and a GlyGly · HCl · H2O single crystal, for which the three‐dimensional crystal structures have already been determined by x‐ray diffraction. The tensor components were measured by changing the angle between the crystal plane and the applied magnetic field by using a goniometer designed in this work for use in conventional 13C cross‐polarization/magic angle spinning nmr probe. From these experimental data, the principal values of the 13C chemical shift tensor and its directions for the Gly1 amide carbonyl carbon were determined. It was found that the 13C chemical shift tensor components (δ11, δ22, and δ33) for the Gly1 amide carbonyl carbon in GlyGly and GlyGly · HNO3 with a >NH · · · OC< type of hydrogen bond depend on the hydrogen‐bond length and the directions of the δ22 components of these peptides are along the hydrogen‐bonded >CO bond axis. In addition, the magnitude of the deviation from the bond axis depends on the hydrogen‐bond angle. Further, the experimental result for GlyGly · HCl · H2O with a O H · · ·OC< type of hydrogen bond was discussed. © 1999 John Wiley & Sons, Inc. Biopoly 50: 61–69, 1999 相似文献
17.
More recently, tremendous progress has been achieved in the development of two‐dimensional semiconductor materials applied in catalyst, energy application, sensor device and bioengineering since the birth of graphene isolated from graphite. Layered molybdenum disulfide (MoS2) as an indirect gap semiconductor can efficiently emit photoluminescence (PL) excited by visible light, which shows a great potential in adaptive biological imaging. However, 1 photon PL of MoS2 for cell imaging purposes suffers from strong autofluorescence and ion‐induced PL quenching. Herein, we report single layer small chitosan decorated MoS2 nanosheets as a nonbleaching, nonblinking optical nanoprobe under near infrared femtosecond laser excitation and their applications for strong 2 photon luminescence (TPL) and strong second harmonic generation (SHG) bioimaging. Furthermore, the TPL can resist the ion‐induced quenching on the cellular membrane. The proposed TPL and SHG of single‐layer MoS2 show great potential for real‐time, deep, multiphoton and three‐dimensional bioimaging under low‐power laser excitation. 相似文献
18.
Annabelle Varrot Gideon J. Davies 《Acta Crystallographica. Section D, Structural Biology》2003,59(3):447-452
Non‐covalent interactions between protein and ligand at the active centre of glycosidases play an enormous role in catalysis. Dissection of these hydrogen‐bonding networks is not merely important for an understanding of enzymatic catalysis, but is also increasingly relevant for the design of transition‐state mimics, whose tautomeric state, hydrogen‐bonding interactions and protonation contribute to tight binding. Here, atomic resolution (∼1 Å) analysis of a series of complexes of the 34 kDa catalytic core domain of the Bacillus agaradhaerens endoglucanase Cel5A is presented. Cel5A is a `retaining' endoglucanase which performs catalysis via the formation and subsequent breakdown of a covalent glycosyl‐enzyme intermediate via oxocarbenium‐ion‐like transition states. Previous medium‐resolution analyses of a series of enzymatic snapshots has revealed conformational changes in the substrate along the reaction coordinate (Davies et al., 1998). Here, atomic resolution analyses of the series of complexes along the pathway are presented, including the `Michaelis' complex of the unhydrolysed substrate, the covalent glycosyl‐enzyme intermediate and the complex with the reaction product, cellotriose. These structures reveal intimate details of the protein–ligand interactions, including most of the carbohydrate‐associated H atoms and the tautomeric state of crucial active‐centre groups in the pH 5 orthorhombic crystal form and serve to illustrate the potential for atomic resolution analyses to inform strategies for enzyme inhibition. 相似文献
19.
Zinc endopeptidase thermolysin can be inhibited by a series of phosphorus-containing peptide analogues, Cbz-Gly-psi (PO2)-X-Leu-Y-R (ZGp(X)L(y)R), where X = NH, O, or CH2; Y = NH or O; R = Leu, Ala, Gly, Phe, H, or CH3. The affinity correlation as well as an X-ray crystallography study suggest that these inhibitors bind to thermolysin in an identical mode. In this work, we calculate the electrostatic binding free energies for a series of 13 phosphorus-containing inhibitors with modifications at X, Y, and R moieties using finite difference solution to the Poisson-Boltzmann equation. A method has been developed to include the solvation entropy changes due to binding different ligands to a macromolecule. We demonstrate that the electrostatic energy and empirically derived solvation entropy can account for most of the binding energy differences in this series. By analyzing the binding contribution from individual residues, we show that the energy of a hydrogen bond is not confined to the donor and acceptor. In particular, the positive charges on Zn and Arg 203, which are not the acceptors, contribute significantly to the hydrogen bonds between two amides of ZGpLL and the thermolysin. 相似文献
20.
During the first few minutes of fibrillation of a 14-residue peptide homologous to the hydrophobic C-terminal part of the Abeta-peptide, EM micrographs reveal small crystalline areas (100 to 150 nm, repeating unit 47 A) scattered in more amorphous material. On a longer time scale, these crystalline areas disappear and are replaced by tangled clusters resembling protofilaments (hours), and eventually by more regular amyloid fibrils of 60 A to 120 A diameter (days). The transient population of the crystalline areas indicates the presence of ordered substructures in the early fibrillation process, the diameter of which matches the length of the 14-mer peptide in an extended beta-strand conformation. 相似文献