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.
We present calculations of the possible magnitude of propinquity which has been proposed to play an important role in enzymic catalysis. The effect has been evaluated as in the past by calculating the ratio of bimolecular to intramolecular reaction rates. The ratio is estimated for intramolecular catalysis in rigid systems as well as for systems with five and six rotatable bonds. Our method differs from others mainly in the way cyclization has been treated. The reaction rate (in all cases) is proportional to the probability that the reactive units have the appropriate spatial and orientational positioning for reaction. This probability is obtained by evaluating a distance distribution function within the spatial and angular intervals to which the units are constrained after reaction. For the bimolecular case we have made the usual assumption that the distribution function is uniform. For the intramolecular reaction, neither the spatial nor the angular part of the distribution function is uniform. The pertinent parameters in this case are the bond lengths and angles, and the statistical weight matrices describing torsional rotation. The difficulty in obtaining analytical expressions for the distribution function is circumvented by using Monte Carlo methods. It is argued that the spatial contribution to rate accelerations in rigid systems can be as high as 107 M, depending upon the size of the volume to which the reactive units are constrained after reaction. The limitation on the smallest physically reasonable volume is estimated from considerations of energy requirements and vibrational amplitudes. Accelerations by five- and six-membered ring cyclizations were estimated at 103 M, the six-membered ring exhibiting the smaller rate enhancement.  相似文献   

2.
Clorobiocin is an aminocoumarin antibiotic containing a 5-methylpyrrolyl-2-carboxyl moiety, attached by an ester bond to a deoxysugar. This pyrrolyl moiety is important for the binding of the antibiotic to its biological target, the B subunit of gyrase. Inactivation experiments had shown that two putative acyl carrier proteins, CloN5 and CloN1, and two putative acyl transferases, CloN2 and CloN7, are involved in the transfer of the pyrrolyl-2-carboxyl moiety to the deoxysugar. In this study, pyrrolyl-2-carboxyl-N-acetylcysteamine thioester was synthesized and fed to cloN1 , cloN2 and cloN7 mutants, and secondary metabolite formation was analyzed by HPLC and HPLC-MS. Transfer of the pyrrolyl-2-carboxyl moiety was observed in the cloN1 and cloN2 mutants, but not in the cloN7 mutant, suggesting that CloN7 is responsible for this reaction. The product of this transfer, novclobiocin 109, was not further methylated to the 5-methylpyrrolyl-2-carboxyl compound, i.e. clorobiocin, suggesting that methylation does not take place after the acyl transfer. Additional investigations for the presence of 5-methylpyrrolyl-2-carboxylic acid in the mutants, and inactivation experiments with the methyltransferase gene cloN6, suggested that methylation by CloN6 and acyl transfer by CloN7 take place in a concerted fashion, requiring the presence of both proteins for efficient product formation. A mechanism for the methylation/acyl transfer process in the late steps of clorobiocin biosynthesis, involving CloN1, CloN2, CloN5, CloN6 and CloN7 is suggested.  相似文献   

3.
Imide capture of a C-terminal peptidylazide with a side-chain thioacid derivative of an N-terminally protected aspartyl peptide leads to the formation of an imide bond bringing the two peptide ends into close proximity. Unmasking of the Nα protecting group and intramolecular acyl migration results in the formation of a native peptide bond to asparagine.  相似文献   

4.
U Azzena  P L Luisi 《Biopolymers》1986,25(4):555-570
Sequences 74–91 and 77–91 of E. coli thioredoxin, which according to x-ray structure contain an irregular β-turn, a hairpinlike structural element, have been synthesized and their conformational properties in solution have been investigated by means of CD spectroscopy. In addition, analogs of these sequences, containing the regular β-turn element Gly-Pro-(Gly)2, have also been prepared and investigated. These are BOC-Ile-Gly-Pro-(Gly)2-Val-OMe (III) and BOC-(Ile)3Gly-Pro-(Gly)2-(Val)5-OMe (IV) that on the basis of probability, should form hairpin structures stabilized by intramolecular interactions. While the natural sequences were shown to be unable to adopt structures characterized by an intrinsic conformational stability, the two analogs showed evidence of intramolecular folding in methanol and trifluoroethanol–water solution. In particular, the CD spectra are indicative of β-structure. The most interesting case was observed for compound IV, as the highest degree of conformational order was present in solutions containing a large proportion of water. In addition, the formation of this structure took place in a highly cooperative manner. The results are utilized to discuss whether and to what extent conformationally stable folding peptide units of small size can be formed in aqueous solution.  相似文献   

5.
Synthetic peptides reproducing the helix‐loop‐helix (HLH) domains of the Id proteins fold into highly stable helix bundles upon self‐association. Recently, we have shown that the replacement of the dipeptide Val‐Ser at the loop–helix‐2 junction with the corresponding O‐acyl iso‐dipeptide leads to a completely unfolded state that only refolds after intramolecular ON acyl migration. Herein, we report on an Id HLH analog based on the substitution of the Pro‐Ser motif at the helix‐1–loop junction with the corresponding O‐acyl iso‐dipeptide. This analog has been successfully synthesized by solid‐phase Fmoc chemistry upon suppression of DKP formation. No secondary structure could be detected for the O‐acyl iso‐peptide before its conversion into the native form by ON acyl shift. These results show that the loop–helix junctions are determinant for the folded/unfolded state of the Id HLH domain. Further, despite the high risk of DKP formation, peptides containing O‐acyl iso‐Pro‐Ser/Thr units are synthetically accessible by Fmoc chemistry. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
S‐Acyl cysteine peptides containing α‐, β‐ or γ‐amino acid residues undergo long‐range S‐ to N‐acyl transfer to give analogs of native tripeptides and tetrapeptides containing additional carbon atoms in the chain. The ease of intramolecular SN‐acyl transfer relative to intermolecular transacylation is favored increasingly for 9 < 12 < 13 ~ 10‐membered cyclic transition states; the observed order is explained on conformational and intermolecular interaction considerations. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
The 4,5-dimethoxy-2-mercaptobenzyl (Dmmb) group attached to a main chain amide in a peptide is easily transformed into an S-peptide via an intramolecular NS acyl shift reaction under acidic conditions, and the S-peptide produces a peptide thioester through an intermolecular thiol–thioester exchange reaction. In order to develop a method for efficiently preparing peptide thioesters based on the NS acyl shift reaction, the factors involved in this process were analyzed in detail. The general features of the transformation at the Dmmb group attached amide bond in a trifluoroacetic acid (TFA) solution and the generation of a peptide thioester were examined by 13C-NMR spectral measurements, reversed-phase (RP) HPLC analyses, mass measurements, and amino acid analyses. The methoxy group of the Dmmb group was not essential for the NS acyl shift reaction, but played a role in stabilizing the thioester form. The addition of water to the TFA solution accelerated the NS acyl shift reaction mediated by the Dmmb group and also suppressed the acid-catalyzed cleavage of the Dmmb group. A peptide thioester was produced from the S-peptide via an intermolecular thiol–thioester exchange reaction with minimal epimerization of the amino acid residue that constituted the thioester bond. Undesirable side reactions, such as the hydrolysis of the thioester bond and an S–N acyl shift reaction occurred during the synthetic process, which is a subject of further investigation.  相似文献   

8.
The molecular conformations of the linear oligopeptides H-(L -Ala)n-L -Pro-OH, with n = 1,2 and 3, have been investigated. 13C nmr observation of the equilibrium between the cis and trans forms of the Ala-Pro peptide bond indicated the occurrence of nonrandom conformations in solutions of these flexible peptides. The formation of the nonrandom species containing the cis form of the Ala-Pro bond was found to depend on the deprotonation of the carboxylic acid group of proline, the solvent, and the ionic strength in aqueous solution. The influence of intramolecular hydrogen bonding on the relative conformational energies of the species containing the cis and trans Ala-Pro peptide bond was studied by comparison of the peptides H-(Ala)n-Pro-OH with analogous molecules where hydrogen bond formation was excluded by the covalent structure. In earlier work a hydrogen bond between the protonated terminal carboxylic acid group and the carbonyl oxygen of the penultimate amino acid residue had been suggested to stabilize conformations including trans proline. For the systems described here this hypothesis can be ruled out, since the cis:trans ratio is identical for molecules with methyl ester protected and free protonated terminal carboxylic acid groups of proline. Direct evidence for hydrogen bond formation between the deprotonated terminal carboxylic acid group and the amide proton of the penultimate amino acid residue in the molecular species containing cis proline was obtained from 1H nmr studies. However, the cis:trans ratio of the Ala-Pro bond was not affected by N-methylation of the penultimate amino acid residue, which prevents formation of this hydrogen bond. Overall the experimental observations lead to the conclusion that the relative energies of the peptide conformations including cis or trans proline are mainly determined by intramolecular electrostatic interactions, whereas in the molecules considered, intramolecular hydrogen bonding is a consequence of specific peptide backbone conformations rather than a cause for the occurrence of energetically favored species. Independent support for this conclusion was obtained from model consideration which indicated that electrostatic interactions between the terminal carboxylic acid group and the carbonyl oxygen of the penultimate amino acid residue could indeed account for the observed relative conformational energies of the species containing cis and trans proline, respectively.  相似文献   

9.
Buczek A  Wałęsa R  Broda MA 《Biopolymers》2012,97(7):518-528
The tendency to adopt β‐turn conformation by model dipeptides with α,β‐dehydrophenylalanine (ΔPhe) residue in the gas phase and in solution is investigated by theoretical methods. We pay special attention to a dependence of conformational properties on the side‐chain configuration of dehydro residue and the influence of N‐methylation on β‐turn stability. An extensive computational study of the conformational preferences of Z and E isomers of dipeptides Ac‐Gly‐(E/Z)‐ΔPhe‐NHMe ( 1a / 1b ) and Ac‐Gly‐(E/Z)‐ΔPhe‐NMe2 ( 2a / 2b ) by B3LYP/6‐311++G(d,p) and MP2/6‐311++G(d,p) methods is reported. It is shown that, in agreement with experimental data, Ac‐Gly‐(Z)‐ΔPhe‐NHMe has a great tendency to adopt β‐turn conformation. In the gas phase the type II β‐turn is preferred, whereas in the polar environment, the type I. On the other hand, dehydro residue in Ac‐Gly‐(E)‐ΔPhe‐NHMe has a preference to adopt extended conformations in all environments. N‐methylation of C‐terminal amide group, which prevents the formation of 1←4 intramolecular hydrogen bond, change dramatically the conformational properties of studied dehydropeptides. Especially, the tendency to adopt β‐turn conformations is much weaker for the N‐methylated Z isomer (Ac‐Gly‐(Z)‐ΔPhe‐NMe2), both in vacuo and in the polar environment. On the contrary, N‐methylated E isomer (Ac‐Gly‐(E)‐ΔPhe‐NMe2) can easier adopt β‐turn conformation, but the backbone torsion angles (?1, ψ1, ?2, ψ2) are off the limits for common β‐turn types. © 2012 Wiley Periodicals, Inc. Biopolymers 97:518–528, 2012.  相似文献   

10.
Benzyloxycarbonyl-L-proline p-guanidinophenyl ester is an "inverse substrate" for trypsin; i.e., the cationic center is included in the leaving group instead of being in the acyl moiety. This substrate can be used in trypsin-catalyzed acyl-transfer reactions leading to the synthesis of Pro-Xaa peptide bonds. The reaction proceeds about 20 times slower than reaction with similar alanine-containing substrates, but the ratio between synthesis and hydrolysis is more favorable. The investigation of a series of nucleophiles led to information about the specificity of the process. Nucleophiles differing only in the P(1)'-position show an increasing acyl transfer efficiency in the order Phe < Gly < Ley < Ser < Ala < lle. C terminal elongation of the nucleophiles is of minor influence on their efficiency. The formation of an H bond between the acyl-enzyme and the nucleophile seems to play an important role in the aminolysis of the acyl-enzyme.  相似文献   

11.
Kinetics of the acyl transfer catalyzed by Xanthomonas α-amino acid ester hydrolase was studied. The enzyme hydrolyzed d-α-phenylglycine methyl ester (d-PG-OMe) to give equimolar amounts of d-α-phenylglycine and methanol. With d-PG-OMe as an acyl donor and 7-amino-3-deacetoxy-cephalosporanic acid (7-ADCA) as an acyl acceptor, the enzyme transferred the acyl group from d-PG-OMe to 7-ADCA in competition with water. The addition of amine nucleophiles (7-ADCA and 6-aminopenicillanic acid) decreased the molecular activity (ko) of the enzyme-catalyzed hydrolysis of d-PG-OMe, whereas it did not alter the Michaelis constant (KM), and plots of l/ko against the initial concentration of a nucleophile (no) gave a straight line. These results support the assumptions that the overall process for hydrolysis and acyl transfer proceeds through a common acyl-enzyme intermediate, that the acylation step of the enzyme is rate-limiting, and that the transfer competes with the hydrolysis of the acyl donor.  相似文献   

12.
The penicillin amidase (PA) from Escherichia coli belongs to a group of proteolytically processed bacterial enzymes. The mechanism of the maturation of the single polypeptide proenzyme has been studied for the PA from E. coli using a slowly processing mutant proenzyme. The mutant proenzyme was constructed by replacing Thr with Gly in the Thr(263)-Ser(264) bond that must be hydrolysed in active PA. The mutant proenzyme was purified by biospecific affinity chromatography using an immobilized monoclonal antibody against PA. The maturation of the free and covalently immobilized purified proenzyme was studied in vitro. For the free proenzyme the same products with PA activity as observed in homogenates of wild-type PA-producing E. coli cells were found to be formed during this process. A kinetic analysis of the possible inter- and intramolecular processes involved in the maturation demonstrated that unambiguous evidence for the existence of intramolecular processes can only be obtained in systems where intermolecular processes are excluded. The Gly(263)-Ser(264) bond was found to be hydrolysed first in the free and immobilized mutant proenzyme, based on determinations of mass spectra, N-terminal sequences and active site concentrations. In the system with immobilized proenzyme intermolecular processes are excluded, demonstrating that this bond is hydrolysed by intramolecular autoproteolysis. Based on the known three-dimensional structure of the PA from E. coli the same maturation mechanism should apply for the wild-type proenzyme.  相似文献   

13.
The mechanism of luminol chemiluminescence is a special case of nucleophilic addition to carbonyl compounds. The breakdown of the key intermediate, an alpha hydroxy hydroperoxide, produces a peracid ortho to an acyl diazene group. After intramolecular addition of the peracid, the energy from nitrogen expulsion is utilized in the formation of an anti-aromatic endoperoxide. Rupture along the O,O bond leaves a substantial part of the ensuing phthalate in its excited state. The emitter is shown to be a mono-protonated phthalate unaccessible by photoexcitation. The dark reaction is a concerted decomposion of the alpha hydroxy hydroperodixe to yield ground-state phthalate.  相似文献   

14.
 A series of oxoiron(IV) porphyrin cation radical complexes was investigated as compound I analogs of cytochrome P-450. Both the spectroscopic features and the reactivities of the complexes in oxygen atom transfer to olefins were examined as a function of only one variable, the axial ligand trans to the oxoiron(IV) bond. The results disclosed two important kinetic steps – electron transfer from olefin to oxoiron(IV) and intramolecular electron transfer from metal to porphyrin radical – which are affected differently by the axial ligands. The large kinetic barrier of the latter step in the reaction of olefins with the perchlorato-bound oxoiron(IV) porphyrin cation radical complex enabled the trapping of a reaction intermediate in which the metal, but not the porphyrin radical, is reduced. The first electron transfer step is probably followed by σ-bond formation, which readily accounts for formation of isomerized organic products at low temperatures. It is finally postulated that part of the enhanced oxygenation activities of cytochrome P-450 monooxygenases and chloroperoxidases is due to a lowering of the energy barrier for the second electron transfer step via participation of their redox-active cysteinate ligand. Received: 16 January 1997 / Accepted: 24 May 1997  相似文献   

15.
Transglutaminases (protein-glutamine:amine γ-glutamyltransferase, EC 2.3.2.13) are a family of calcium-dependent enzymes that catalyze an acyl transfer between glutamine residues and a wide variety of primary amines. When a lysine residue acts as the acyl-acceptor substrate, a γ-glutamyl-ε-lysine isopeptide bond is formed. This isopeptide bond formation represents protein cross-linking, which is critical to several biological processes. Microbial transglutaminase (mTG) is a bacterial variant of the transglutaminase family, distinct by virtue of its calcium-independent catalysis of the isopeptidic bond formation. Furthermore, mTG’s promiscuity in acyl-acceptor substrate preference highlights its biocatalytic potential. The acyl-donor substrate, however, is limited in its scope; the amino acid sequences flanking glutamine residues dramatically affect substrate specificity and activity. Here, we have developed and optimized a modified glutamate dehydrogenase assay with the intention of analyzing potential high-affinity peptides. This direct continuous assay presents significant advantages over the commonly used hydroxamate assay, including generality, sensitivity, and ease of manipulation. Furthermore, we identified 7M48 (WALQRPH), a high-affinity peptide that shows greater affinity with mTG (KM = 3 mM) than the commonly used Cbz-Gln-Gly (KM = 58 mM), attesting to its potential for application in biocatalysis and bioconjugation.  相似文献   

16.
A conformational analysis of protected glutamate homo-oligopeptides Z-[Glu(OEt)]n-OEt (n = 2–7) was carried out in chloroform solution using high-resolution 1H-nmr spectroscopy. At dilute peptide concentrations, the backbone NH and α-CH resonances are well resolved and can be assigned by combining extensive homonuclear decoupling experiments with data for co-oligopeptide derivatives. The structure of these peptides in solution was then assessed using information from chemical shifts, coupling constants, temperature coefficients, and titration of each oligomer with trifluoroacetic acid (TFA). The di- and tripeptides are found to be in disordered forms in deuterochloroform (CDCl3) and CDCl3/TFA mixtures. The tetrapeptide exhibits a folded structure with intramolecular hydrogen bonding at Glu2 in CDCl3 and undergoes a transition to increasingly disordered forms as TFA is added. The pentamer to heptamer show a folded structure with a strong intramolecular hydrogen bond at Glu2 and a weaker hydrogen bond at Glu3, which are disrupted as these peptides go to random coils at high TFA/CDCl3 ratios. In addition, the N-terminal portions of these glutamate peptides appear to be involved in side chain–main chain interactions. The results support the hypothesis that protected linear homo-oligopeptides may possess two or more segments of conformation with intramolecular folding preferred near the N-terminal portion.  相似文献   

17.
Farazi TA  Waksman G  Gordon JI 《Biochemistry》2001,40(21):6335-6343
MyristoylCoA:protein N-myristoyltransferase (Nmt) attaches myristate to the N-terminal Gly residue of proteins involved in a variety of signal transduction cascades, and other critical cellular functions. To gain insight about the structural basis of substrate recognition and catalysis, we determined the structures of a binary complex of Saccharomyces cerevisiae Nmt1p with myristoylCoA to 2.2 A resolution and of a ternary complex of Nmt1p with a nonhydrolyzable myristoylCoA analogue [S-(2-oxo)pentadecylCoA] and an octapeptide substrate (GLYASKLA) to 2.5 A resolution. The binary complex reveals how myristoylCoA alters the conformation of the enzyme to promote binding of both myristoylCoA and peptide and identifies the backbone amides of F170 and L171 as an oxyanion hole which polarizes the reactive thioester carbonyl. The ternary complex structure reveals details of the enzyme's peptide binding specificity and illuminates its mechanism of acyl transfer. The N-terminal Gly ammonium is positioned in close proximity to the C-terminal carboxylate of the protein, where it is poised to undergo the required deprotonation to an amine. In this conformation, the nucleophile is 6.3 A away from the thioester carbonyl. A catalytic mechanism is proposed whereby, once deprotonation is initiated, the N-terminal Gly amine can approximate the thioester carbonyl by rotating along Psi. This motion is facilitated by a H-bond network and leads to reaction between the glycine nitrogen nucleophile and the carbonyl. Loss of CoA from the tetrahedral intermediate may be facilitated by intramolecular H-bonding of the sulfur to the adenylamine of CoA. This affords a compact leaving group and lends a role for the observed bends in the CoA structure. The absolute requirement for Gly at the N-terminus of substrates is explained by the requirement for flexible rotation of its amine.  相似文献   

18.
James P. Tam  Qitao Yu 《Biopolymers》1998,46(5):319-327
In biological systems, both proteolysis and aminolysis of amide bonds produce activated intermediates through acyl transfer reactions either inter- or intramolecularly. Protein splicing is an illustrative example that proceeds through a series of catalyzed acyl transfer reactions and culminates at an O- or S-acyl intermediate. This intermediate leads to an uncatalyzed acyl migration to form an amide bond in the spliced product. A ligation method mimicking the uncatalyzed final steps in protein splicing has been developed utilizing the acyl transfer amide-bond feature for the blockwise coupling of unprotected, free peptide segments at methionine (Met). The latent thiol moiety of Met can be exploited using homocysteine at the α-amino terminal position of a free peptide for transthioesterification with another free peptide containing an α-thioester to give an S-acyl intermediate. A subsequent, proximity-driven S- to N-acyl migration of this acyl intermediate spontaneously rearranges to form a homocysteinyl amide bond. S-methylation with excess p-nitrobenezensulfonate yields Met at the ligation site. The methionine ligation is selective and orthogonal, and is usually completed within 4 h when performed at slightly basic pH and under strongly reductive conditions. No side reactions due to acylation were observed with any other α-amines of both peptide segments as seen in the synthesis of parathyroid hormone peptides. Furthermore, cyclic peptide can also be obtained through the same strategy by placing both homocysteine at the amino terminus and the thioester at the carboxyl terminus in an unprotected peptide precursor. These biomimetic ligation strategies hold promise for engineering novel peptides and proteins. © 1998 John Wiley & Sons, Inc. Biopoly 46: 319–327, 1998  相似文献   

19.
The effect of Clostridium perfringens alpha-toxin on liposomes prepared from phosphatidylcholine (PC) containing the fatty acyl residues of 18 carbon atoms was investigated. The toxin-induced carboxyfluorescein (CF) leakage and phosphorylcholine release from multilamellar liposomes increased as the phase transition temperature of the phosphatidylcholines containing unsaturated fatty acyl residues decreased. However, there was no difference between the sensitivity of the different phosphatidylcholines solubilized by deoxycholate to the phospholipase C (PLC) activity of the toxin. However, the toxin did not hydrolyze solubilized distearoyl-l -α-phosphatidylcholine (DSPC) or phosphatidylcholine containing saturated fatty acyl residue, and caused no effect on liposomes composed of DSPC. These results suggest that the activity of the toxin is closely related to the membrane fluidity and double bond in PC. The N-terminal domain of alpha-toxin (AT1-246) and variant H148G did not induce CF leakage from liposomes composed of dioleoyl-l -α-phosphatidylcholine (DOPC). H148G bound to the liposomes, but AT1-246 did not. However, the C-terminal domain (AT251-370) conferred binding to liposomes and the membrane-damaging activity on AT1-246. These observations suggest that the membrane-damaging action of alpha-toxin is due to the binding of the C-terminal domain of the toxin to the double bond in the PC in the bilayer and hydrolysis of the PC by the N-terminal domain.  相似文献   

20.
The solid state conformations of cyclo[Gly–Proψ[CH2S]Gly–D –Phe–Pro] and cyclo[Gly–Proψ[CH2–(S)–SO]Gly–D –Phe–Pro] have been characterized by X-ray diffraction analysis. Crystals of the sulfide trihydrate are orthorhombic, P212121, with a = 10.156(3) Å, b = 11.704(3) Å, c = 21.913(4) Å, and Z = 4. Crystals of the sulfoxide are monoclinic, P21, with a = 10.662(1) Å, b = 8.552(3) Å, c = 12.947(2) Å, β = 94.28(2), and Z = 2. Unlike their all-amide parent, which adopts an all-trans backbone conformation and a type II β-turn encompassing Gly-Pro-Gly-D -Phe, both of these peptides contain a cis Gly1-Pro2 bond and form a novel turn structure, i.e., a type II′ β-turn consisting of Gly–D –Phe–Pro–Gly. The turn structure in each of these peptides is stabilized by an intramolecular H bond between the carbonyl oxygen of Gly1 and the amide proton of D -Phe4. In the cyclic sulfoxide, the sulfinyl group is not involved in H bonding despite its strong potential as a hydrogen-bond acceptor. The crystal structure made it possible to establish the absolute configuration of the sulfinyl group in this peptide. The two crystal structures also helped identify a type II′ β-turn in the DMSO-d6 solution conformers of these peptides. © 1993 John Wiley & Sons, Inc.  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