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 46 毫秒
1.
In this work, the translational self-diffusion constants, DT's, of 12 amino acids (Ala, Arg, Asn, Asp, Cys, Glu, His, Ile, Lys, Met, Phe, and Ser) are measured by field gradient NMR and extrapolated to infinite dilution. The experiments were carried out in D2O at 298 K at pD approximately =3.5 in 50 mM sodium phosphate buffer. Of these 12 amino acids, 6 are being reported for the first time (Asp, Cys, Glu, His, Lys, and Met) and the remaining 6 (Ala, Arg, Asn, Ile, Phe, and Ser) are compared with DT's from the literature. When corrected for differences in solvent viscosity and temperature, the discrepancy between DT's measured in the present work and those reported previously is always <8%, which is reasonable given the range of values reported previously by different groups. With the present work, DT's for all of the amino acids are now available. These diffusion constants are then used in modeling studies of the diffusion and free solution electrophoretic mobility, mu, of several model peptides. For this set of peptides, it is shown that modeling using revised input parameters results in improved agreement between model and experimental mobilities.  相似文献   

2.
Chiral sulfoxides are widely used in organic synthesis as chiral auxiliaries. There are numerous strategies for the preparation of enantiomerically pure sulfoxides, based either on the enantioselective oxidation of sulphides or the enantiospecific reduction of sulfoxides. For both cases, bioconversion techniques have been developed and proposed for large-scale synthesis. Methionine sulfoxide reductase enzymes (MsrA and MsrB) catalyse the stereoselective conversion of methionine sulfoxide to methionine. MsrA can also catalyse the reduction of other exogenous sulfoxides, including p-tolyl methyl sulfoxide. However, the stereoselectivity towards this type of substrate is not yet well characterized. The activity and enantioselectivity of MsrA toward several aryl methyl sulfoxides is presented in this paper.  相似文献   

3.
A dual‐selector system employing achiral crown ethers in combination with cyclodextrins has been developed for the separation of peptide diastereomers that contain methionine sulfoxide. The combinations of the crown ethers 15‐crown‐5, 18‐crown‐6, Kryptofix® 21 and Kryptofix® 22 and β‐cyclodextrin, carboxymethyl‐β‐cyclodextrin, and sulfated β‐cyclodextrin were screened at pH 2.5 and pH 8.0 using a 40/50.2 cm, 50 μm id fused‐silica capillary and a separation voltage of 25 kV. No diastereomer separation was observed in the sole presence of crown ethers, while only sulfated β‐cyclodextrin was able to resolve some peptide diastereomers at pH 8.0. Depending on the amino acid sequence of the peptide and the applied cyclodextrin, the addition of crown ethers, especially the Krpytofix® diaza‐crown ethers, resulted in significantly enhanced chiral recognition. Keeping one selector of the dual system constant, increasing concentrations of the second selector resulted in increased peak resolution and analyte migration time for peptide‐crown ether‐cyclodextrin combinations. The simultaneous diastereomer separation of three structurally related peptides was achieved using the dual selector system.  相似文献   

4.
The proton affinities (PA) of the 20 naturally occurring alpha-amino acids (AA) have been determined computationally by means of density functional theory (DFT) and high-level G2(MP2) calculations. These theoretical PAs, together with data that have appeared since 1997 in the literature, are used to validate the most reasonable currently available PA scale for AAs (Harrison, A. G. Mass Spectrom. Rev. 1997, 16, 201-217.). Significant scatter is observed for the PAs of Ser, Asp, Phe, Asn, Met, Pro, Gln, Glu, Trp, His, Lys, and Arg, many of which have a basic side-chain functionality. Critical review of the available data leads to new consensus PAs for Asn, Gln, Met, and Arg of 222.4, 230.5, 223.7, and 250.2 kcal/mol, respectively.  相似文献   

5.
Amino acid residue-specific backbone and side-chain dissociations of peptide z ions in MS(3) spectra were elucidated for over 40 pentapeptides with arginine C-terminated sequences of the AAXAR and AAHXR type, nonapeptides of the AAHAAXX"AR and AAHAXAX"AR type, and AAHAAXX"AAR decapeptides. Peptide z(n) ions containing amino acid residues with readily transferrable benzylic or tertiary β-hydrogen atoms (Phe, Tyr, His, Trp, Val) underwent facile backbone cleavages to form dominant z(n-2) or z(n-3) ions. These backbone cleavages are thought to be triggered by a side-chain β-hydrogen atom transfer to the z ion C(α) radical site followed by homolytic dissociation of the adjacent C(α)-CO bond, forming x(n-2) cation-radicals that spontaneously dissociate by loss of HNCO. Amino acid residues that do not have readily transferrable β-hydrogen atoms (Gly, Ala) do not undergo the z(n) → z(n-2) dissociations. The backbone cleavages compete with side-chain dissociations in z ions containing Asp and Asn residues. Side-chain dissociations are thought to be triggered by α-hydrogen atom transfers that activate the C(β)-C(γ) or C(β)-heteroatom bonds for dissociations that dominate the MS(3) spectra of z ions from peptides containing Leu, Cys, Lys, Met, Ser, Arg, Glu, and Gln residues. The Lys, Arg, Gln, and Glu residues also participate in γ-hydrogen atom transfers that trigger other side-chain dissociations.  相似文献   

6.
Ricin and its corresponding polypeptides (A & B chain) were purified from castor seed. The molecular weight of ricin subunits were 29,000 and 28,000 daltons. The amino acids in ricin determined were Asp45 The22 Ser40 Glu53 Cys4 Gly96 His5 Ile21 Leu33 Lys20 Met4 Phe13 Pro37 Tyr11 Ala45 Val23 Arg20 indicating that ricin contains approximately 516 amino acid residues. The amino acids of the two subunits of ricin A and B chains were Asp23 The12 Ser21 Glu29 Cys2 Gly48 His3 Ile12, Leu17 Lys10 Met2 Phe6 Pro17 Tyr7 Ala35 Val13 Arg13 while in B chain the amino acids were Asp22 The10 Ser19 Glu25 Cys2 Gly47 His1 Ile10, Leu15 Lys11 Met1 Phe7 Pro6 Tyr5 Ala32Val11 Arg10. The total helical content of ricin came around 53.6% which is a new observation.  相似文献   

7.
The thermodynamic stability and oligomerization status of the tumor suppressor p53 tetramerization domain have been studied experimentally and theoretically. A series of hydrophilic mutations at Met-340 and Leu-344 of human p53 were designed to disrupt the hydrophobic dimer-dimer interface of the tetrameric oligomerization domain of p53. Meanfield calculations of the free energy of the solvated mutants as a function of interdimer distance were compared with experimental data on the thermal stability and oligomeric state [tetramer, dimer, or equilibrium mixture of both] of each mutant. The calculations predicted a decreasing stability and oligomeric state for the following amino acids at residue 340: Met [tetramer] > Ser Asp, His, Gin, > Glu, Lys [dimer], whereas the experimental results showed the following order: Met [tetramer] > Ser > Gln > His, Lys > Asp, Glu [dimers]. For residue 344, the calculated trend was Leu [tetramer] > Ala > Arg, Gln, Lys [dimer], and the experimental trend was Leu [tetramer] > Ala, Arg, Gln, Lys [dimer]. The discrepancy for the lysine side chain at residue 340 is attributed to the dual nature of lysine, both hydrophobic and charged. The incorrect prediction of stability of the mutant with Asp at residue 340 is attributed to the fact that within the meanfield approach, we use the wild-type backbone configuration for all mutants, but low melting temperatures suggest a softening of the α-helices at the dimer-dimer interface. This initial application of meanfield theory toward a protein-solvent system is encouraging for the application of the theoretical model to more complex systems.  相似文献   

8.
Xylanases from Bacillus circulans (BCX) are known as configuration-retaining glycoside hydrolases, which hydrolyze xylans with two glutamic acid residues (Glu78 and Glu172) serving as catalytic active residues according to a double displacement mechanism. Existing experimental researches show that mutating the asparagines (Asn) to aspartic acid (Asp) at position 35 next to Glu172 can obviously improve the catalytic activity of BCX. To better understand the inherent mechanism for the experimental finding, we performed quantum chemistry calculations on two model systems to mimic the catalyses of wild-type and mutant BCXs. Geometrical structures and relative energies of intermediates and transition states involved in the hydrolysis reactions are given in detail. It is found that in the wild-type model system Asn35 interacts with Glu172 via a loose hydrogen bond, while in the mutant model system Asp35 forms a very tight hydrogen bond with Glu172. The glycosidic bond cleavage is proposed to be the rate-determining step for the hydrolysis reaction, whose barrier varies from 98 to 65 kJ mol−1 when Asn35 is replaced by Asp35, showing the presence of Asp35 remarkably reduces the energy demand for the hydrolysis reaction. The present result provides a theoretical elucidation for why a single amino acid substitution can importantly influences catalytic activity of BCX.  相似文献   

9.
The adhesive polydecapeptide poly(Lys‐Pro‐Thr‐Gln‐Tyr‐Ser‐Asp‐Glu‐Tyr‐Lys) (average repeating number, n = 5), which is the consensus sequence of the Asian freshwater mussel Limnoperna fortunei adhesive protein (Lffp), has been synthesized by the polycondensation of the active esters. The surface chemical experiments revealed the following characteristics of the freshwater adhesive protein: (i) wettability of the Lffp solution is affected by the polar component value (γsp) of the surface free energy of the substrate, and a substrate having a γsp less than 10 mJ·m–2 exhibits a reduced wettability of the Lffp solution; (ii) the comparison of wettability of native Lffp with synthetic Lffp suggests that the decapeptide sequence, ‐Lys‐Pro‐Thr‐Gln‐Tyr‐Ser‐Asp‐Glu‐Tyr‐Lys‐, contributes to the interaction with the underwater surface; (iii) the Lffp tends to adsorb on nonpolar surfaces that have a low γsp value; and (iv) the adsorption ability of the freshwater adhesive protein is less than that of the marine adhesive protein because of the higher hydrophilicity of the freshwater adhesive protein. An antifouling examination indicated that a γsp value of the substrate surface of less than 10 mJ·m–2 should achieve a higher antifouling effect towards the L. fortunei attachment. These results are the first findings for the development of a freshwater antifouling strategy based on the molecular mechanism underlying the attachment of L. fortunei.  相似文献   

10.
The role of the second shell in the process of metal binding and selectivity in metalloproteins has been elucidated by combining Protein Data Bank (PDB) surveys of Mg, Mn, Ca, and Zn binding sites with density functional theory/continuum dielectric methods (DFT/CDM). Peptide backbone groups were found to be the most common second-shell ligand in Mg, Mn, Ca, and Zn binding sites, followed (in decreasing order) by Asp/Glu, Lys/Arg, Asn/Gln, and Ser/Thr side chains. Aromatic oxygen- or nitrogen-containing side chains (Tyr, His, and Trp) and sulfur-containing side chains (Cys and Met) are seldom found in the second coordination layer. The backbone and Asn/Gln side chain are ubiquitous in the metal second coordination layer as their carbonyl oxygen and amide hydrogen can act as a hydrogen-bond acceptor and donor, respectively, and can therefore partner practically every first-shell ligand. The second most common outer-shell ligand, Asp/Glu, predominantly hydrogen bonds to a metal-bound water or Zn-bound histidine and polarizes the H-O or H-N bond. In certain cases, a second-shell Asp/Glu could affect the protonation state of the metal ligand. It could also energetically stabilize a positively charged metal complex more than a neutral ligand such as the backbone and Asn/Gln side chain. As for the first shell, the second shell is predicted to contribute to the metal selectivity of the binding site by discriminating between metal cations of different ionic radii and coordination geometries. The first-shell-second-shell interaction energies decay rapidly with increasing solvent exposure of the metal binding site. They are less favorable but are of the same order of magnitude as compared to the respective metal-first-shell interaction energies. Altogether, the results indicate that the structure and properties of the second shell are dictated by those of the first layer. The outer shell is apparently designed to stabilize/protect the inner-shell and complement/enhance its properties.  相似文献   

11.
Highly efficient catalytic asymmetric Claisen rearrangements of O‐propargyl β‐ketoesters and O‐allyl β‐ketoesters have been accomplished under mild reaction conditions. In the presence of the chiral N,N′‐dioxide/NiII complex, a wide range of allenyl/allyl‐substituted all‐carbon quaternary β‐ketoesters was obtained in generally good yield (up to 99 %) and high diastereoselectivity (up to 99:1 d.r.) with excellent enantioselectivity (up to 99 % ee).  相似文献   

12.
Ion formation may be made more efficient than in normal electrospray ionization (ESI) for certain classes of compounds, such as the polar amino acids Glu, Asn, His, Ser, Asp, Arg, Tyr and Lys, by adjusting the voltage of a normal ESI interface needle to zero voltage. For aspartic acid (Asp) the gain in signal-to-noise (S/N) ratio of the liquid chromatography/mass spectrometry (LC/MS) chromatograms obtained in the selective ion monitoring (SIM) mode (m/z 134) with zero needle potential was 40-50 times higher than detection at 4 kV. Ion formation at zero potential is likely to follow a mechanism related to sonic spray ionization. The utility of the zero needle voltage ESI was illustrated by determining the age of a human tooth by the aspartic acid epimerization method. The procedure involved separating the D- and L-aspartic acid of a tooth extract on a chiral HPLC column and detection by zero voltage ESI-MS3.  相似文献   

13.
A novel selectively deprotectable triazacyclophane scaffold was used for the design and split-mix synthesis of two libraries of solid-phase bound tripodal synthetic receptors possessing three different amino acid or peptidic arms. In the synthesis of the first library, the two outer arms consisted of amino acid Ala, Arg, Asp, Gln, Gly, Lys, Phe, Ser, Tyr, or Val and the middle arm consisted of amino acid Asn, Glu, His, Leu, or Pro. The second library contained amino acid and/or (di)peptide arms. The arms were different in all library members. The first outer arm consisted of amino acid(s) Ala, Arg, Gln, Phe, or Ser, the second outer arm consisted of amino acid(s) Asp, Gly, Lys, Tyr, or Val, and the middle arm consisted of amino acid(s) Asn, Glu, His, Leu, or Pro, leading to a 27 000 member library of synthetic tripodal receptor molecules. In on-bead screening experiments, a remarkable selectivity of some library members for Fe(3+) was observed and decoding of their structures by Edman degradation revealed consensus sequences with structural resemblance to non-heme iron proteins.  相似文献   

14.
Molecular simulation by using force field parameters has been widely applied in the fields of peptide and protein research for various purposes. We recently proposed a new all‐atom protein force field, called the SAAP force field, which utilizes single amino acid potentials (SAAPs) as the fundamental elements. In this article, whole sets of the SAAP force field parameters in vacuo, in ether, and in water have been developed by ab initio calculation for all 20 proteinogenic amino acids and applied to Monte Carlo molecular simulation for two short peptides. The side‐chain separation approximation method was employed to obtain the SAAP parameters for the amino acids with a long side chain. Monte Carlo simulation for Met‐enkephalin (CHO‐Tyr‐Gly‐Gly‐Phe‐Met‐NH2) by using the SAAP force field revealed that the conformation in vacuo is mainly controlled by strong electrostatic interactions between the amino acid residues, while the SAAPs and the interamino acid Lennard‐Jones potentials are predominant in water. In ether, the conformation would be determined by the combination of the three components. On the other hand, the SAAP simulation for chignolin (H‐Gly‐Tyr‐Asp‐Pro‐Glu‐Thr‐Gly‐Thr‐Trp‐Gly‐OH) reasonably reproduced a native‐like β‐hairpin structure in water although the C‐terminal and side‐chain conformations were different from the native ones. It was suggested that the SAAP force field is a useful tool for analyzing conformations of polypeptides in terms of intrinsic conformational propensities of the single amino acid units. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009  相似文献   

15.
The synthesis of two cyclic pentapeptides cyclo (-Arg-Lys-Xxx-D-Val-Tyr-) (Xxx=Asp or Glu) with thymopentin-analogue sequences is described. Cyclization was achieved by the carbodiimide/DMAP method. The results of the NMR investigations performed on the protected pentapeptides suggest a βII′/γ-structure in DMSO solution.  相似文献   

16.
Human serum albumin (HSA) is involved in the transport of metal ions and potential metallodrugs. Depending on the metal, several sites are available, among which are N-terminal (NTS) and multi-metal binding sites (MBS). Despite the large number of X-ray determinations for albumins, only one structure with Zn2+ is available. In this work, the binding to HSA of the VIVO2+ ion was studied by an integrated approach based on spectroscopic and computational methods, which allowed the systems to be characterized even in the absence of X-ray analysis. The behavior depends on the type of albumin, defatted (HSAd) or fatted (HSAf). With HSAd ‘primary’ and ‘secondary’ sites were revealed, NTS with (His3, His9, Asp13, Asp255) and MBS with (His67, His247, Asp249, Asn99 or H2O); with increasing the ratio VIVO2+/HSAd, ‘tertiary’ sites, with one His-N and other donors (Asp/Glu-O or carbonyl-O) are populated. With HSAf, fatty acids (FAs) cause a rotation of the subdomains IA and IIA, which results in the formation of a dinuclear ferromagnetic adduct (VIVO)2D(HSAf) with a μ1,1-Asp249 and the binding of His247, Glu100, Glu252, and His67 or Asn99. FAs hinder also the binding of VIVO2+ to the MBS.  相似文献   

17.
An optimized HPLC/MS/MS method was established to quantify glutamate (Glu) and aspartic acid (Asp) in rat hippocampus with glutamate‐d5 (Glu‐d5) as internal standard. The mass spectrometry was operated under the multiple reaction monitoring mode using electrospray ionization in the positive ion mode for Glu and negative ion mode for Asp. The retention times of Glu, Asp and Glu‐d5 were 1.53, 2.07 and 1.52 min, respectively. The linearity of calibration curves was good, with r2 > 0.99 and a lower limit of quantitation of 10 ng/mL. The intra‐day precisions (relative standard deviation, RSD) of Glu and Asp were in the range of 3.61–8.17 and 4.22–10.09%, respectively; the inter‐day precisions (RSD) of Glu and Asp were in the range of 3.57–5.19 and 2.49–5.04%, respectively. The accuracies of Glu and Asp were in the range of ?2.10–6.20 and ?0.90–10.00%, respectively. The recovery rates of 10, 100 and 1000 ng/mL were found to be 0.89 ± 0.24, 1.01 ± 0.10 and 0.90 ± 0.12 for Glu and 0.99 ± 0.26, 0.93 ± 0.07 and 1.13 ± 0.13 for Asp, respectively. This optimized method was successfully applied to quantify the concentration of Glu and Asp in rat hippocampus in brain ischemia/reperfusion research. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

18.
李轶  李琳  黄开勋 《化学进展》2012,24(7):1398-1404
蛋氨酸(Met)是生物体内很容易被氧化的氨基酸之一,氧化应激条件下,生成S型和R型蛋氨酸亚砜(MetO), 晶状体蛋白中MetO的增加与晶状体老化和白内障形成相关。生物体内存在着两种不同的蛋氨酸亚砜还原酶(Msr),即MsrA和B,分别能特异性地作用于自由或结合在蛋白质中的S-MetO和R-MetO,将MetO修复为Met,从而避免了蛋白质结构和功能的改变。在哺乳动物中,MsrA以单基因形式存在,而MsrB有3种异构体,分别为MsrB1,MsrB2和MsrB3,其中MsrB1是一个硒蛋白,又被称为硒蛋白R(SelR)。本文介绍了Msrs的基因表达、分布和亚细胞定位,比较了MsrA和MsrBs蛋白结构和催化机制的异同,讨论了晶状体蛋白Met残基的氧化与白内障形成和发展的关系。现有的这些研究结果表明Msrs作为一类特异性的抗氧化还原酶,通过对MetO的修复,在抑制晶状体的损伤方面发挥重要作用。此外,MsrB1作为一个硒蛋白受机体硒水平的调节,因此,通过补硒保持晶状体适当的硒浓度以维持MsrB1的活性,对白内障的形成和发展可能具有一定的预防作用。  相似文献   

19.
Nucleoside phosphoramidates (NPs) are a class of nucleotide analogues that has been developed as potential antiviral/antitumor prodrugs. Recently, we have shown that some amino acid nucleoside phosphoramidates (aaNPs) can act as substrates for viral polymerases like HIV‐1 RT. Herein, we report the synthesis and hydrolysis of a series of new aaNPs, containing either natural or modified nucleobases to define the basis for their differential reactivity. Aqueous stability, kinetics, and hydrolysis pathways were studied by NMR spectroscopy at different solution pD values (5–7) and temperatures. It was observed that the kinetics and mechanism (P? N and/or P? O bond cleavage) of the hydrolysis reaction largely depend on the nature of the nucleobase and amino acid moieties. Aspartyl NPs were found to be more reactive than Gly or β‐Ala NPs. For aspartyl NPs, the order of reactivity of the nucleobase was 1‐deazaadenine>7‐deazaadenine>adenine>thymine≥3‐deazaadenine. Notably, neutral aqueous solutions of Asp‐1‐deaza‐dAMP degraded spontaneously even at 4 °C through exclusive P? O bond hydrolysis (a 50‐fold reactivity difference for Asp‐1‐deaza‐dAMP vs. Asp‐3‐deaza‐dAMP at pD 5 and 70 °C). Conformational studies by NMR spectroscopy and molecular modeling suggest the involvement of the protonated N3 atom in adenine and 1‐ and 7‐deazaadenine in the intramolecular catalysis of the hydrolysis reaction through the rare syn conformation.  相似文献   

20.
The metal-binding ability of human ubiquitin (hUb) towards a selection of biologically relevant metal ions and complexes has been probed. Different techniques have been used to obtain crystals suitable for crystallographic analysis. In the first type of experiments, crystals of hUb have been soaked in solutions containing copper(II) acetate and two metallodrugs, Zeise salt (K[PtCl(3)(η(2)-C(2)H(4))]·H(2)O) and cisplatin (cis-[PtCl(2)(NH(3))(2)]). The Zeise salt is used in a test for hepatitis, whereas cisplatin is one of the most powerful anticancer drugs in clinical use. The Zeise salt readily reacts with hUb crystals to afford an adduct with three platinum residues per protein molecule, Pt(3)-hUb. In contrast, copper(II) acetate and cisplatin were found to be unreactive for contact times up to one hour and to cause degradation of the hUb crystals for longer times. In the second type of experiments, hUb was cocrystallized with a solution of copper(II) or zinc(II) acetate or cisplatin. Zinc(II) acetate gives, at low metal-to-protein molar ratios (8:1), crystals containing one metal ion per three molecules of protein, Zn-hUb(3) (already reported in previous work), whereas at high metal-to-protein ratios (70:1) gives crystals containing three Zn(II) ions per protein molecule, Zn(3)-hUb. In contrast, once again, copper(II) acetate and cisplatin, even at low metal-to-protein ratios, do not give crystalline material. In the soaking experiment, the Zeise anion leads to simultaneous platination of His68, Met1, and Lys6. Present and previous results of cocrystallization experiments performed with Zn(II) and other Group 12 metal ions allow a comprehensive understanding of the metal-ion binding properties of hUb with His68 as the main anchoring site, followed by Met1 and carboxylic groups of Glu16, Glu18, Glu64, Asp21, and Asp32, to be reached. In the case of platinum, Lys6 can also be a binding site. The amount of bound metal ion, with respect to that of the protein, appears to be a relevant parameter influencing crystal packing.  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