首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The primary electron donor in the photosynthetic reaction center from purple bacteria is a bacteriochlorophyll dimer containing four conjugated carbonyl groups that may form hydrogen bonds with amino acid residues. Spectroscopic analyses of a set of mutant reaction centers confirm that hydrogen bonds can be formed between each of these carbonyl groups and histidine residues in the reaction center subunits. The addition of each hydrogen bond is correlated with an increase in the oxidation potential of the dimer, resulting in a 355-mV range in the midpoint potential. The resulting changes in the free-energy differences for several reactions involving the dimer are related to the electron transfer rates using the Marcus theory. These reactions include electron transfer from cytochrome c2 to the oxidized dimer, charge recombination from the primary electron acceptor quinone, and the initial forward electron transfer.  相似文献   

2.
Optimal conditions for the conjugation of carboxyl groups on low molecular weight molecules to reactive amino groups on rabbit immunoglobulin G (IgG) using a modified carbodiimide reaction have been investigated. Reaction of [14C]hippuric acid with N-ethyl-N′-(dimethylaminopropyl) carbodiimide at pH 5 followed by adjustment to pH 8 and coupling with rabbit IgG resulted in the formation of hippuric acid-IgG conjugates with less than 10% intra- and intermolecular IgG crosslinking. More than 93% of the bonds linking hippuric acid to IgG were stable to hydroxylamine hydrolysis, indicating the peptide properties of these bonds. This two-step process permitted a defined number of hippuric acid moieties to be loaded onto a single IgG molecule and should provide a useful method for the conjugation of molecules containing carboxyl groups to amino groups on a variety of polypeptides.  相似文献   

3.
The barotropic behavior of neat and aqueous 1,2,3-triacetyl glycerol was investigated by FT-IR spectroscopy over the pressure range 0.001 to 35 kbar. The infrared spectrum in the presence of water shows bands characteristic of hydrogen bonded carbonyl groups. An increase in hydrostatic pressure leads to a strengthening of the intermolecular hydrogen bond between water and the lipid ester C = O groups. The pressure-induced formation of ice VI at 9 kbar does not affect this hydrogen bond, however, the formation, at 20 kbar, of ice VII in which the water/water hydrogen bonds are stronger than the lipid C = O/water hydrogen bonds, frees the lipid carbonyl groups from the hydrogen-bonding to water.  相似文献   

4.
We have determined the partial molar volumes and adiabatic compressibilities of a homologous series of six alpha,omega-aminocarboxylic acids over a broad pH range at 25 degrees C. We interpret the resulting data in terms of the changes in hydration associated with neutralization of amino and carboxyl termini. By combining our volumetric results with pH-dependent data on 1-anilinonaphthalene-8-sulfonic acid fluorescence we propose the following explanation to the long-standing observation that changes in volume and compressibility accompanying neutralization of a carboxyl group depend on the type of the solute in contrast to solute-independent changes in these parameters accompanying neutralization of an amino group. Unlike amino groups, neutralized carboxyl groups are capable of forming hydrogen-bonded structures stabilized by hydrogen bonds between the carbonyl oxygen of one solute molecule and the hydroxyl group of another molecule. Formation of such hydrogen-bonded structures causes an additional decrease in solute hydration with concomitant increases in volume and compressibility. Furthermore, solutes with large aliphatic moieties may form larger associates stabilized, in addition to intermolecular hydrogen bonds, by hydrophobic interactions which will result in further increases in volume and compressibility. In the aggregate, our results emphasize the need for further studies focused on developing an understanding of the role of electrostatic interactions in stabilizing/destabilizing proteins and protein complexes.  相似文献   

5.
Interaction of tetrahydrocortisol (THF) and its apolipoprotein A-I (ApoA-I) complex with a CC(GCC)5/GG(CGG)5 duplex were studied by IR spectroscopy, which revealed formation of hydrogen bonds between the OH group of the THF A-ring and the C=O groups of cytosine and guanine. In addition, THF formed hydrogen bonds with the PO2 group of the duplex and with a sugar OH group. THF and ApoA-I interacted with the duplex in the same active site involving the base C=O groups. THF increased the conformational order in the duplex, while the THF-ApoA-I complex induced an order → disorder transition.  相似文献   

6.
Sulfur atoms are present as thiol and thioether functional groups in amino acids, coenzymes, cofactors, and various products of secondary metabolic pathways. The biosynthetic pathways for several sulfur-containing biomolecules require the substitution of sulfur for hydrogen at unreactive aliphatic or electron-rich aromatic carbon atoms. Examples discussed in this review include biotin, lipoic acid, methylthioether modifications found in some nucleic acids and proteins, and thioether cross-links found in peptide natural products. Radical S-adenosyl-l-methionine (SAM) enzymes use an iron-sulfur cluster to catalyze the reduction of SAM to methionine and a highly reactive 5′-deoxyadenosyl radical; this radical can abstract hydrogen atoms at unreactive positions, facilitating the introduction of a variety of functional groups. Radical SAM enzymes that catalyze sulfur insertion reactions contain a second iron-sulfur cluster that facilitates the chemistry, either by donating the cluster''s endogenous sulfide or by binding and activating exogenous sulfide or sulfur-containing substrates. The use of radical chemistry involving iron-sulfur clusters is an efficient anaerobic route to the generation of carbon-sulfur bonds in cofactors, secondary metabolites, and other natural products.  相似文献   

7.
Manikandan K  Ramakumar S 《Proteins》2004,56(4):768-781
A comprehensive database analysis of C--H...O hydrogen bonds in 3124 alpha-helices and their corresponding helix termini has been carried out from a nonredundant data set of high-resolution globular protein structures resolved at better than 2.0 A in order to investigate their role in the helix, the important protein secondary structural element. The possible occurrence of 5 --> 1 C--H...O hydrogen bond between the ith residue CH group and (i - 4)th residue C==O with C...O < or = 3.8 A is studied, considering as potential donors the main-chain Calpha and the side-chain carbon atoms Cbeta, Cgamma, Cdelta and Cepsilon. Similar analysis has been carried out for 4 --> 1 C--H...O hydrogen bonds, since the C--H...O hydrogen bonds found in helices are predominantly of type 5 --> 1 or 4 --> 1. A total of 17,367 (9310 of type 5 --> 1 and 8057 of type 4 --> 1) C--H...O hydrogen bonds are found to satisfy the selected criteria. The average stereochemical parameters for the data set suggest that the observed C--H...O hydrogen bonds are attractive interactions. Our analysis reveals that the Cgamma and Cbeta hydrogen atom(s) are frequently involved in such hydrogen bonds. A marked preference is noticed for aliphatic beta-branched residue Ile to participate in 5 --> 1 C--H...O hydrogen bonds involving methylene Cgamma 1 atom as donor in alpha-helices. This may be an enthalpic compensation for the greater loss of side-chain conformational entropy for beta-branched amino acids due to the constraint on side-chain torsion angle, namely, chi1, when they occur in helices. The preference of amino acids for 4 --> 1 C--H...O hydrogen bonds is found to be more for Asp, Cys, and for aromatic residues Trp, Phe, and His. Interestingly, overall propensity for C--H...O hydrogen bonds shows that a majority of the helix favoring residues such as Met, Glu, Arg, Lys, Leu, and Gln, which also have large side-chains, prefer to be involved in such types of weak attractive interactions in helices. The amino acid side-chains that participate in C--H...O interactions are found to shield the acceptor carbonyl oxygen atom from the solvent. In addition, C--H...O hydrogen bonds are present along with helix stabilizing salt bridges. A novel helix terminating interaction motif, X-Gly with Gly at C(cap) position having 5 --> 1 Calpha--H...O, and a chain reversal structural motif having 1 --> 5 Calpha-H...O have been identified and discussed. Our analysis highlights that a multitude of local C--H...O hydrogen bonds formed by a variety of amino acid side-chains and Calpha hydrogen atoms occur in helices and more so at the helix termini. It may be surmised that the main-chain Calpha and the side-chain CH that participate in C--H...O hydrogen bonds collectively augment the cohesive energy and thereby contribute together with the classical N--H...O hydrogen bonds and other interactions to the overall stability of helix and therefore of proteins.  相似文献   

8.
The crystal and molecular structure of a dipeptide containing a blocked gamma-carboxyglutamyl (Gla) residue is presented. Two intermolecular hydrogen bonds link the amides with carbonyl groups in the dipeptide backbone, but the protected gamma-carboxy groups on the modified glutamic acid are not hydrogen bonded.  相似文献   

9.
Proteins which have been exposed to the hydroxyl radical (.OH) or to the combination of .OH plus the superoxide anion radical and oxygen (.OH + O2- + O2) exhibit altered primary structure and increased proteolytic susceptibility. The present work reveals that alterations to primary structure result in gross distortions of secondary and tertiary structure. Denaturation/increased hydrophobicity of bovine serum albumin (BSA) by .OH, or by .OH + O2- + O2 was maximal at a radical/BSA molar ratio of 24 (all .OH or 50% .OH + 50% O2-). BSA exposed to .OH also underwent progressive covalent cross-linking to form dimers, trimers, and tetramers, partially due to the formation of intermolecular bityrosine. In contrast, .OH + O2- + O2 caused spontaneous BSA fragmentation. Fragmentation of BSA produced new carbonyl groups with no apparent increase in free amino groups. Fragmentation may involve reaction of (.OH-induced) alpha-carbon radicals with O2 to form peroxyl radicals which decompose to fragment the polypeptide chain at the alpha-carbon, rather than at peptide bonds. BSA fragments induced by .OH + O2- + O2 exhibited molecular weights of 7,000-60,000 following electrophoresis under denaturing conditions, but could be visualized as hydrophobic aggregates in nondenaturing gels (confirmed with [3H]BSA following treatment with urea or acid). Combinations of various chemical radical scavengers (mannitol, urate, t-butyl alcohol, isopropyl alcohol) and gases (N2O, O2, N2) revealed that .OH is the primary species responsible for alteration of BSA secondary and tertiary structure. Oxygen, and O2- serve only to modify the outcome of .OH reaction. Furthermore, direct studies of O2- + O2 (in the absence of .OH) revealed no measurable changes in BSA structure. The process of denaturation/increased hydrophobicity was found to precede either covalent cross-linking (by .OH) or fragmentation (by .OH + O2- + O2). Denaturation was half-maximal at a radical/BSA molar ratio of 9.6, whereas half-maximal aggregation or fragmentation occurred at a ratio of 19.4. Denaturation/hydrophobicity may hold important clues for the mechanism(s) by which oxygen radicals can increase proteolytic susceptibility.  相似文献   

10.
The single crystals of coordinated complexes of neutral erythritol (C4H10O4) with various transition metal ions were synthesized and studied using FT-IR and single crystal X-ray diffraction analysis. Two CuCl2-erythritol complexes (denoted as CuE(I) and CuE(II)) were obtained. In CuE(I), Cu2+ coordinates with two chloride ions and four OH groups from two erythritol molecules. Two copper centers are linked by one erythritol molecule to form a zigzag chain. For CuE(II), each Cu2+ coordinates with two OH groups from an erythritol molecule and two chloride ions. The crystal of CuE(II) contains complexed and free erythritol, the dimers of [Cu2Cl4(C4H10O4)] further form a [Cu2Cl4(C4H10O4)]infinity chain via secondary Cu...Cl bonds, both the dimer unit of [Cu2Cl4.(C4H10O4)] and non-coordinated C4H10O4 unit exist side by side in the crystal. MnCl2-erythritol complex whose structure is similar to CuE(I) is also acquired. The OH groups of erythritol act as ligand to coordinate to metal ions on one hand, one the other hand, OH groups form hydrogen bonds network that link chain and layer together to build three-dimensional structures.  相似文献   

11.
The self-association of several N-urethanyl-L-amino acids (N-t-Boc-glycine, N-t-Boc-L-alanine, N-t-Boc-L-methionine, and N-t-Boc-O-Bz-L-tyrosine has been investigated in carbon tetrachloride by Fourier transform (FTIR) spectrometry. The fractions of nonbonded OH and NH groups have been determined from the intensities of the free OH and NH-stretching vibrations. The FTIR spectra have been examined in the carbonyl stretching regions by using Fourier self-deconvolution techniques. The results show that intermolecular hydrogen bonds are formed between the carboxylic acid group of one molecule and the urethane group of another molecule, suggesting that the leading factor for the self-association of N-t-Boc amino acids is the acidity of the OH groups. In N-t-Boc glycine, more heterodimers are formed than in the other amino acids. In N-t-Boc-O-Bz-L-tyrosine where the steric hindrance of the substituent implanted on the Cα atom is higher, the concentration of homodimers and heterodimers is lower than in N-t-Boc-L-alanine or N-t-Boc-L methionine. © 1997 John Wiley & Sons, Inc.  相似文献   

12.
The conformation and the interaction of CHF2OCF2CHF2 (desflurane II) with one water molecule is investigated theoretically using the ab initio MP2/aug-cc-pvdz and DFT-based M062X/6-311++G(d,p) methods. The calculations include the optimized geometries, the harmonic frequencies of relevant vibrational modes along with a natural bond orbital (NBO) analysis including the NBO charges, the hybridization of the C atom and the intra- and intermolecular hyperconjugation energies. In the two most stable conformers, the CH bond of the F2HCO- group occupies the gauche position. The hyperconjugation energies are about the same for both conformers and the conformational preference depends on the interaction between the non-bonded F and H atoms. The deprotonation enthalpies of the CH bonds are about the same for both conformers, the proton affinity of the less stable conformer being 3 kcal mol?1 higher. Both conformers of desflurane II interact with water forming cyclic complexes characterized by CH…O and OH…F hydrogen bonds. The binding energies are moderate, ranging from ?2.4 to ?3.2 kcal mol?1 at the MP2 level. The origin of the blue shifts of the ν(CH) vibrations is analyzed. In three of the complexes, the water molecule acts as an electron donor. Interestingly, in these cases a charge transfer is also directed to the non bonded OH group of the water molecule. This effect seems to be a property of polyfluorinated ethers.  相似文献   

13.
The novel purine and pyrimidine derivatives of 1-aminocyclopropane-1-carboxylic acid 1 and 2 were obtained by alkylation of 6-(N-pyrrolyl)purine and thymine with methyl 1-benzamido-2-chloromethylcyclopropanecarboxylate. X-ray crystal structure analysis shows that the cyclopropane rings in 1 and 2 posses Z-configuration. The cyclopropane ring atoms and attached atoms of the benzamido and methoxycarbonyl moiety of both molecules are disposed perpendicularly to each other. The carbonyl oxygen of the methoxycarbonyl moiety adopts in both compounds a synperiplanar conformation with respect to the midpoint of the distal bond of the cyclopropane ring. The torsion angles Phi and psi for the 1-aminocyclopropane-1-carboxylic acid residue in 1 and 2 correspond to a folded conformation, while the torsion angles omega define antiperiplanar conformation. Intermolecular hydrogen bonds connect the molecules of 1 into dimers. Each dimer is hydrogen-bonded with four ethanol molecules, thus forming discrete unit. On the contrary, intermolecular hydrogen bonds link the molecules of 2 generating three-dimensional network.  相似文献   

14.
Raman spectra are reported for crystalline nonactin, monactin, dinactin, trinactin, and tetranactin and their solutions in CCl4, CHCl3, CH3OH, and 4:1 (v/v) CH3OH:CHCl3. The macrotetrolide nactins selectively bind a wide variety of cations, and are important model compounds for the study of ion complexation. The conformations of nonactin, monactin, and dinactin in solution are similar. Their conformations are found to be sufficiently open to permit the ester carbonyl groups to form hydrogen bonds with CH3OH; this gives rise to characteristic changes in the vibration frequencies associated with the ester groups. Nonactin, which is the least soluble of the nactins in CH3OH, is also the least effective at forming hydrogen bonds with CH3OH. The greater ability of the higher nactins to form hydrogen bonds with CH3OH may be due to the increased inductive effect of ethyl over methyl side chains, which may increase the dipole moment of the ester carbonyl groups. Spectra of crystalline nonactin, monactin, and tetranactin are fairly similar, while the spectra of dinactin and trinactin comprise a second, distinct family. This is consistent with X-ray crystallographic studies, which show that nonactin and tetranactin form monoclinic crystals, while trinactin is triclinic.  相似文献   

15.
The structure of α-chitin has been determined by X-ray diffraction, based on the intensity data from deproteinized lobster tendon. Least-squares refinement shows that adjacent chains have alternating sense (i.e. are antiparallel). In addition, there is a statistical distribution of side-chain orientations, such that all the hydroxyl groups form hydrogen bonds. The unit cell is orthorhombic with dimensions a = 0.474 ± 0.001 nm, b = 1.886 ± 0.002 nm and c = 1.032 ± 0.002 nm (fiber axis); the space group is P212121 and the cell contains disaccharide sections of the two chains passing through the center and corner of the ab projection. The chains form hydrogen-bonded sheets linked by CO…HN bonds approximately parallel to the a axis, and each chain has an O-3′H…O.5 intramolecular hydrogen bond, similar to that in cellulose. Adjacent chains along the ab diagonal have different conformations for the CH2OH groups: on one chain these groups form O.6H…O.6′ intermolecular hydrogen bonds to the CH2OH group on the adjacent chain along the ab diagonal. The latter group is oriented to form an intramolecular O.6′H…O.7 bond to the carboxyl oxygen on the next residue. The results indicate that a statistical mixture of CH2OH orientations is present, equivalent to half oxygens on each residue, each forming inter- and intramolecular hydrogen bonds. As a result the structure contains two types of amide groups, which differ in their hydrogen bonding, and account for the splitting of the amide I band in the infrared spectrum. The Inability of this chitin polymorph to swell on soaking in water is explained by the extensive intermolecular hydrogen bonding.  相似文献   

16.
This article reports, for the first time, on the absolute configuration of (+)-9-benzyloxy-α-dihydrotetrabenazine ( 8 ), as determined from the perspective of X-ray crystallography. Compound 8 was prepared by a six-step reaction using 3-benzyloxy-4-methoxybenzaldehyde ( 1 ) as a starting material. The X-ray crystal diffraction structure of two compounds, racemic 9-benzyloxy-tetrabenazine ( 5 ) and the diastereomeric salt of compound 8 , is also described for the first time in this article. The X-ray results and the chiral HPLC helped elucidate that compound 8 has an absolute configuration as 2R,3R,11bR. The crystal structure of racemic compound 5 contains two symmetry- independent molecules in the unit cell. Interestingly, while they are structural isomers, they are enantiomers, too, i.e., in solution, because they are not mirror images of each other in the crystal lattice. In order to elucidate the intermolecular interaction mechanism of the diastereomeric salt of compound 8 , its crystal packing was investigated with regard to the weak interactions, such as salt bridge, OH…O and CH…O hydrogen bonds, and intermolecular CH…π interaction. The results showed that the carbonyl-assisted salt bridges and the OH…O hydrogen bonds formed polar columns in the crystal structure of the diastereomeric salt of compound 8 , resembling butterflies with open wings as viewed along the c-axis. These polar columns were extended to three-dimensional network by intermolecular CH…O hydrogen bonds and intermolecular CH…π interactions. Chirality, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

17.

Melinjo seed extract contains melinjo resveratrol compounds that has antioxidant activity. The radical-scavenging sites required for the antioxidant activity, however, is yet to be located. We report a computational study that aims to locate scavenging sites of the simplest resveratrol dimer, gnetin C. We consider the reaction of gnetin C and hydroperoxyl radical energetically with the basis of density-functional calculations, to be compared with the reaction of the resveratrol monomer, trans-resveratrol, and hydroperoxyl radical. The results show that OH group at the para position is the most reactive scavenging site for both molecules. Besides the OH group, gnetin C also provides two CH groups in the furan ring that are favorable as scavenging sites. Therefore, furan ring plays an important role in the scavenging activity, which is contrary to the experimental speculation that propose resorcinol ring. Our study shows the prospect of density-functional calculation for studying the radical-scavenging reaction.

  相似文献   

18.
Summary 2D 1H NMR spectroscopy of two -helical peptides which differ in their amphipathicity has been used to investigate the relationships between amide-proton chemical shifts, amide-proton exchange rates, temperature, and trifluoroethanol (TFE) concentration. In 50% TFE, in which the peptides are maximally helical, the amide-proton chemical shift and temperature coefficient patterns are very similar to each other in each peptide. Temperature coefficients from –10 to –6 ppb/K, usually indicative of the lack of intramolecular hydrogen bonds, were observed even for hydrophobic amino acids in the center of the -helices. However, slow hydrogen isotope exchange for residues from 4 to 16 in both 18-mer helices indicates intact intramolecular hydrogen bonds over most of the length of these peptides. Based on these anomalous observations, we suggest that the pattern of amide-proton shifts in -helices in H2O/TFE solvents is dominated by bifurcated intermolecular hydrogen-bond formation between the backbone carbonyl groups and TFE. The amide-proton chemical shift changes with increasing temperature may be interpreted by a disruption of intermolecular hydrogen bonds between carbonyl groups and the TFE in TFE/water rather than by the length of intramolecular hydrogen bonds in -helices. Supplementary Material is available upon request, comprising seven pages with listings of experimental details and the NMR shift data for the two peptides.  相似文献   

19.
The aim of this study was to characterize the exopolysaccharide (EPS) produced by Bacillus velezensis SN-1 (B. velezensis SN-1) (EPS-SN-1), which was isolated from the fermented Da jiang. The microbe made crude exopolysaccharides EPS-SN-1 was produced throughout the bacterial growth period, and the highest yield (2.7 g/L) was obtained with sucrose as the carbon source. As per high performance liquid chromatography (HPLC), EPS-SN-1 is a heteropolysaccharide consisting of glucose, mannose and fructose, with a high molecular weight of 2.21 × 105 Da. FTIR spectra further indicated the presence of hydroxyl and carbonyl groups, and NMR analysis confirmed both α- and β-glycosidic bonds. Furthermore, differential scanning calorimetry (DSC) showed that EPS-SN-1 has high thermal stability with fusion point of 270.7 °C. Finally, EPS-SN-1 demonstrated strong antioxidant capacity via its ability to scavenge hydroxyl radical (•OH), 1,1-diphenyl-2-picrylhydrazyl (DPPH•) radical, ABTS radical (ABTS•+) and oxygen radical (O2−•). Taken together, EPS-SN-1 is a promising natural antioxidant and probiotic with potential applications in the food industry.  相似文献   

20.
Breton J  Chitnis PR  Pantelidou M 《Biochemistry》2005,44(14):5402-5408
P700, the primary electron donor of photosystem I, is an asymmetric dimer made of one molecule of chlorophyll a' (P(A)) and one of chlorophyll a (P(B)) that are bound to the homologous PsaA and PsaB polypeptides. While the carbonyl groups of P(A) are involved in hydrogen-bonding interactions with several surrounding amino acid side chains and a water molecule, P(B) does not engage hydrogen bonds with the protein. Notably, the residue Thr A739 is donating a strong hydrogen bond to the 9-keto C=O group of P(A) and the homologous residue Tyr B718 is free from interaction with P(B). Light-induced FTIR difference spectroscopy of the photooxidation of P700 has been combined with a site-directed mutagenesis attempt to introduce hydrogen bonds to the carbonyl groups of P(B) in Synechocystis sp. PCC 6803. The FTIR study of the Y(B718)T mutant provides evidence that the 9-keto C=O group of P(B) and P(B)(+) engages a relatively strong hydrogen-bonding interaction with the surroundings in a significant fraction (40 +/-10%) of the reaction centers. Additional mutations on the two PsaB residues homologous to those involved in the main interactions between the PsaA polypeptide and the 10a-carbomethoxy groups of P(A) affect only marginally the vibrational frequency of the 10a-ester C=O group of P(B). The FTIR data on single, double, and triple mutants at these PsaB sites indicate a plasticity of the interactions of the carbonyl groups of P(B) with the surrounding protein. However, these mutations do not perturb the hydrogen-bonding interactions assumed by the 9-keto and 10a-ester C=O groups of P(A) and P(A)(+) with the protein and have only a limited effect on the relative charge distribution between P(A)(+) and P(B)(+).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号