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.
N.m.r. and c.d. spectroscopy have been used to study the interactions of cyclic hexapeptide cyclo(Pro-Sar-Sar)2 with metal ions and ammonium ions. Cyclo(Pro-Sar-Sar)2 was found to form complexes with Li+, K?, Ba2+ and Cu2+, accompanying the conformational change into a single conformer, and the conformation of cyclo(Pro-Sar-Sar)2 in the Li+-complex was different from that in the Cu2+-complex. These findings indicate conformational flexibility of cyclo(Pro-Sar-Sar)2. The equilibrium constant for the complexation with Li+ was 2.3 × 102l mol?1, and cyclo(Pro-Sar-Sar)2 adopted an asymmetric conformation in the complex. The addition of α-amino acid ester hydrochloride also caused the conformational change of cyclo(Pro-Sar-Sar)2), but in this case it did not converge into a single conformation. This type of interaction was strengthened with aromatic α-amino acid ester hydrochloride due to the aromatic-amide interactions. Finally, the rates of exchange between unbound α-amino acid ester hydrochlorides and those complexed with cyclo(Pro-Sar-Sar)2 were found to be different, according to the nature of α-amino acid.  相似文献   

2.
Cyclic octapeptides, cyclo(X-Pro)4, where X represents Phe, Leu, or Lys(Z), were synthesized and their conformations investigated. A C2-symmetric conformer containing two cis peptide bonds was found in all of these cyclic octapeptides. The numbers of available conformations due to the cistrans isomerization of Pro peptide bonds depended on the nature of the solvent and X residue: they decreased in the following order: cyclo[Lys(Z)-Pro]4 > cyclo(Leu-Pro)4 > cyclo(Phe-Pro)4 in CDCl3. 13C spin-lattice relaxation times (T1) of these cyclic octapeptides were measured, and the contribution of segmental mobility to T1 was found to vary with the nature of the X residue.  相似文献   

3.
In order to establish whether p.m.r. spectroscopy is useful for identifying Amadori- and Heyns-rearrangement products, the p.m.r. spectra at 220 MHz of 16 rearrangement products derived from d-glucose or d-fructose and amino acids have been investigated. At pH 3, the protons of the NCH2 group of N-substituted 1-amino-1-deoxy-d-fructose (Amadori-rearrangement products) resonate at δ 3.25–3.60 in D2O and are shifted upfield by 0.3–0.6 p.p.m. at pH 9. These protons exchange with deuterium. Also, in D2O there is an equilibrium of the acyclic, furanose, and pyranose structures, the last being favoured. At pH ? 7, the equilibrium is completely shifted to the β-pyranose form, which adopts exclusively the 2C5 conformation. At pH 3, the equilibrium favours the β-furanose form. At pH 3, H-1e and H-1a of N-substituted 2-amino-2-deoxy-d-glucoses (Heyns-rearrangement products) resonate at δ 5.55 and 5.04, respectively. At pH 9, the signal for H-2 is shifted upfield by 0.2–0.7 p.p.m. In D2O solution, these compounds exist as an equilibrium of α- and β-pyranose forms in the 4C1 conformation. The α anomer is stabilised by the amino acid group at position 2. At pH 3, the αβ-ratio is 2–4:1, and, at pH 9, 1.0–1.1:1.  相似文献   

4.
C M Deber  P D Adawadkar 《Biopolymers》1979,18(10):2375-2396
We have synthesized and characterized a series of cation-binding cyclic octapeptides which may function as potential ionophoric substances. The materials contain varying degrees of hydrophobic character, which was controlled systematically through the incorporation of N-alkylglycine residues where N-alkyl = methyl, n-hexyl, cyclohexyl, or n-decyl. The peptides reported include cyclo(Phe-Sar-Gly-Sar)2, cyclo(Glu(OBzl)-Sar-Gly-Sar-Glu(OBzl)-Sar-Gly-(N-decyl)Gly), cyclo(Glu(OBzl)-Sar-Gly-(N-decyl)Gly)2, cyclo(Glu(OBzl)-Sar-Gly-(N-hexyl)Gly)2, cyclo(Glu(OBzl)-Sar-Gly-(N-cyclohexyl)Gly)2, and the corresponding free diacid forms of the Glu-containing compounds. Using 13C- and 1H-nmr spectra, we demonstrated that the mixture of cis/trans peptide bond-isomer conformers, characteristic of the free-peptide benzyl esters in solution, was converted to unique C2-symmetric, presumably all-trans conformers on complexation with calcium ions. Cation-transport experiments, using the thick-liquid model of transport in a Pressman cell, established that these compounds transport a variety of cations and that one peptide examined in detail, cyclo(Glu(OBzl)-Sar-Gly-(N-decyl)Gly)2 (selectivity Ca2+ > Na+ > K+ > Mn2+ > Cu2+ > Mg2+ > Co2+ > Zn2+), transports calcium about an order of magnitude more efficiently than magnesium.  相似文献   

5.
The conformational analysis of a number of N-hexopyranosylimidazoles and their tetra-O-acetyl derivatives has been carried out using 1H-n.m.r. data obtained after computer simulation of spectra. Evidence is presented that, for some compounds, mixtures of 4C1 and 1C4 conformers are present in solution, and the possible contributions of steric effects and polar, reverse anomeric effects are discussed. It is concluded that the results can in large part be accounted for by steric factors, but that the operation of additional polar factors is likely. Present rationalisations of the reverse anomeric effect are discussed and a stereoelectronic interpretation is presented. The conformations of the exocyclic hydroxymethyl groups are analysed and shown to give additional information about the presence of alternative chair conformations.  相似文献   

6.
The net charges on various atoms of aldohexopyranose pentaacetates were computed by using the MO-LCAO method of Del Re for σ-charges and the Hückel MO method for π-charges. The potential and free energies of sixteen aldohexopyranose pentaacetates in the C1(D) and 1C(D) conformations were estimated. Minimization of the energies of these conformations was studied by suitably tilting the axial CC and CO bonds. As with the free sugars, considerable release of strain is achieved when tilts of 4.5 and 2° are given to the axial CH2OAc and the axial OAc groups, respectively, involved in the Hassel—Ottar effect in the 1C(D) conformations. In the case of C1(D) conformations, the ideal models have the minimum energy even when the acetate groups are involved in syn-axial interactions, indicating that strain induced by axial acetate groups is less than that of axial hydroxyl groups. The calculated free-energies agree well with the experimental values after adding a value of 0.9 kcal.mole-1 for the anomeric effect of the acetoxyl group. The free-energy calculations also predict that α-D-idohexopyranose pentaacetate and α-D-altrose pentaacetate favour the C1(D) conformation and β-D-idose pentaacetate a C1?1C equilibrium in solution, in agreement with n.m.r. studies.  相似文献   

7.
Conformational analysis of muscimol, a GABA agonist   总被引:3,自引:0,他引:3  
The potential energy barriers for rotation around the C5C6 bond in muscimol and two related isoxazoles have been calculated using the MINDO/3 molecular orbital method. The preferred conformations have N7C6C5C4 torsion angles near ± 100 °, in agreement with crystallographic data. The activities of muscimol and related isoxazoles as bicuculline-sensitive inhibitors of neuronal firing, however, are best accounted for in terms of “active conformations” with N7C6C5C4 torsion angles in the range +(32–46) °. These findings predict a limited range of possible “active conformations” for the flexible neurotransmitter GABA at postsynaptic receptors common to GABA, (+)-bicuculline salts and muscimol.  相似文献   

8.
Conformation in aqueous solution at pH 6.95 of tripeptides having cyclic dipeptide backbones, cyclo[l-Glu(l-Leu-OBzl)-l-His] and cyclo[l-Glu(l-Leu-OH)-l-His], was investigated by u.v., c.d. and n.m.r. spectroscopy and by the lanthanide probe method. In the major conformation of cyclo[l-Glu(l-Leu-OBzl)-l-His], the cyclic dipeptide backbone takes a flagpole-boat conformation in which the sidechain of the l-His residue is nearly parallel with the backbone plane and the sidechain of the l-Glu residue protrudes outside the backbone plane. In the major conformation of cyclo[l-Glu(l-Leu-OH)-l-His], the cyclic dipeptide backbone takes a flagpole-boat conformation in which the sidechains of the l-His and l-Glu residues are accommodated in the same side of the backbone plane so that the imidazolyl sidechain of l-His residue is twisted slightly. Tripeptides were not found to change the conformation when metal salts or ammonium salts such as Cl?H3N?(CH2)11 COOEt, Gly-OEt-HCl, dl-Val-OEt-HCl and l-Leu-OEt-HCl were added, but a significant conformation change occurred upon adding d-Leu-OEt·HCl. If the same situation holds with the addition of α-amino acid p-nitrophenyl ester hydrochlorides, the previously reported enantiomer-selective catalysis by the tripeptides which hydrolysed d-Leu-OPh(NO2·HCl faster than l-Leu-OPh(NO2)·HCl can be explained; that is, the tripeptides change the conformation only when d-Leu-OPh(NO2)·HCl is bound and consequently the intramolecular reaction is facilitated. This phenomenon may be compared with that of ‘induced fit’ in enzyme catalysis.  相似文献   

9.
Cyclic hexapeptides, cyclo (L-Leu-L-Phe-L-Pro)2 and cyclo[L-Cys(Acm)-L-Phe-L-Pro]2, in which Acm represents an acetoamide-methyl group, were synthesized, and the conformation and complexation with metal ions were investigated. Cooperation of the carbonyl groups of the Cys(Acm) side chains with those of the cyclic skeleton in complexation was especially examined. Cyclo(L-Leu-L-Phe-L-Pro)2, which possesses no functional groups on side chains, was taken as the reference compound. 13C- and two-dimensional n.m.r. measurements revealed that cyclo(L-Leu-L-Phe-L-Pro)2 and cyclo[L-Cys(Acm)-L-Phe-L-Pro]2 took a C2-symmetric conformation containing cis L-Phe-L-Pro bonds in chloroform and acetonitrile. Both cyclic hexapeptides were found to complex selectively with Ba2+ and Ca2+ in acetonitrile. On complexation the conformation of either cyclic hexapeptide changed into a similar one. However, the binding constant of cyclo[L-Cys(Acm)-L-Phe-L-Pro]2 was higher than that of cyclo(L-Leu-L-Phe-L-Pro)2. The n.m.r. measurements showed that the amide carbonyl groups of Cys(Acm) side chains as well as those of cyclic skeleton in cyclo[L-Cys(Acm)-L-Phe-L-Pro]2 cooperatively bound the cations.  相似文献   

10.
Abstract

An 15N-NMR study at natural abundance of 04/N3-substituted pyrimidine and C6-substituted purine ribonucleosides has shown that the exact location of the protecting group (substituent) on either 04 or N3 in pyrimidines has a strong influence on the electronic properties of the resultant pyrimidine system, mainly because of the change of state of hybridization of the N3-nitrogen. The basicity of N3 in some C4-substituted pyrimidines has been studied by following the 15N chemical shifts of protonated species in the presence of CF3COOH both in DMSO and in CH2Cl2 solution. A comparison of the basic character of N3 in C4-substituted pyrimidine and of N1 in C6-substituted purine nucleosides has shown that the magnitude of the 15N shift of N3 (or N1) upon protonation is governed mainly by the electronic properties of the heteroatom linked to C4 (or C6). It also clearly emerged in this study that there is very litle difference in basicities of N3 of pyrimidine and N1 of purine nucleosides despite the presence of the fused imidazole moiety in the latter.  相似文献   

11.
The conformations of heparin in aqueous solution in the presence of sodium, potassium, magnesium and calcium cations were studied using circular dichroism, optical rotation, nuclear magnetic resonance and equilibrium dialysis. Potassium and magnesium cations, when added to sodium heparinate solutions, cause small chiroptical changes. Binding of calcium ions gives rise to large changes in both optical rotation and circular dichroism. This is indicative of a major change in chain conformation, which is also manifest in 13C and 1H n.m.r.4Equilibrium dialysis suggests one mole of calcium bound per mole of tetrasaccharide, which n.m.r. indicates to be appropriately sulphated iduronateglucosamine-iduronate-glucosamine. The calcium is chelated by two iduronate carboxyl groups. Proton-proton coupling constants, determined by convolution difference spectroscopy and Carr-Purcell sequences, indicate that, over the temperature range 285 to 353 K, the iduronate ring is best described as 1C4(l) and the glucosamine residue as 4C1(d) for both sodium and calcium forms.The conformational change induced by calcium is ascribed to rotation around the glycosidic linkages. The binding process is co-operative and the binding constant of 103 to 104m?1 is biologically significant. The findings are consistent with intramolecular binding. Hence, this study represents the first report of a polysaccharide undergoing a cation-induced intramolecular disorder-order process. The authors postulate that a function of the post-polymerization epimerization of d-glucuronate to l-iduronate is the attainment of the precise geometry required for co-operative calcium binding with consequent modulation of the flexibility of the tetrasaccharide units.  相似文献   

12.
CHOLINERGIC substrates have been found in a gauche conformation (G), skewed about the Cα–Cβ bond of the (CH3)3N+–CH2–CH2–O cholinic fragment in a number of crystal structures1–11. Sulphur and selenium isologues, on the other hand, with the (CH3)3N–CH2–CH2–(S,Se) group, are normally in the extended trans conformations (T)10,12,13. In agreement with the assumption that the reduced spectrum of biological activity of many rigid analogues and Cα or Cβ substituted derivatives of acetylcholine can be partially ascribed to the reduction in conformational flexibility14–17, a theoretical investigation18 predicted the existence of four almost isoenergetic conformations TTTT, TGTT, TTGT and TGGT about the ψ0, ψ1, ψ2 and ψ3 internal rotation angles schematically represented in Fig. 1.  相似文献   

13.
Twelve-membered head-to-tail cyclic tetrapeptides (CTPs) are rigid molecules found in nature and possess a diverse range of biological activities. A possible reason may be due to their ability to adopt rigid conformations in solution mimicking reverse-turns. Reverse-turns are common structural motifs which serve as molecular recognition sites in many protein-receptor interactions. In this paper, we describe the solid-phase synthesis of the antibacterial cyclic tetrapeptide cyclo[Gly-Ser-Pro-Glu] (cyclo[GSPE]), first isolated from the Ruegeria strain of marine bacteria by Mitova et al. (J Nat Prod 67:1178–1181, 2004). Our NMR experiments in H2O:D2O:DMSO (18:1:1) revealed that it possessed three conformations in an approximate ratio of 4:2:1 based on NMR amide peak intensities. 2D NMR studies and computer calculations revealed that the major conformer adopted a reverse-turn conformation and have ω torsion angles twisted by up to 2°, with two transoid amide bonds between Gly-Ser, Pro-Glu and two cisoid amide bonds between Ser-Pro, Glu-Gly in a cistrans-cistrans (ctct) pattern. This supports previous reports that majority of CTPs adopt a ctct pattern when dissolved in hydrogen-bond disrupting solvents (Che and Marshall in J Med Chem 49:111–124, 2006 and references cited therein). An ensemble of ten lowest-energy-minimised 3D structures generated using XPLOR-NIH software revealed that cyclo[GSPE] possessed a rigid backbone ring scaffold. The remaining two minor conformers were present in quantities too low for NMR structural studies.  相似文献   

14.
Cyclic dipeptide cyclo(l- or d-Glu-l-His) carrying an anionic site and a nucleophilic site has been synthesized and used as a catalyst for the solvolysis of cationic esters in aqueous alcohols. In the solvolysis of 3-acyloxy-N-trimethylanilinium iodide (S+n, n = 2 and 10) and Cl?H3N+(CH2)11COOPh(NO2), no efficient nucleophilic catalysis was observed. On the other hand, in the solvolysis of Gly-OPh(NO2)·HCl, Val-OPh(NO2)·HCl and Leu-OPh(NO2)·HCl a very efficient general base-type catalysis by cyclo(l-Glu-l-His) was observed. In particular, with the latter two substrates the catalysis by cyclo(l-Glul-His) was more efficient than that by imidazole, although the catalysis was not enantiomer-selective. The diastereomeric cyclic dipeptide cyclo(d-Glu-l-His) was almost inactive under the same conditions. Confomation of cyclo(l- or d-Glu-l-His) in aqueous solution was investigated and the structure/catalysis relationship is discussed.  相似文献   

15.
Previously, we isolated a new enzyme, N-substituted formamide deformylase, that catalyzes the hydrolysis of N-substituted formamide to the corresponding amine and formate (H. Fukatsu, Y. Hashimoto, M. Goda, H. Higashibata, and M. Kobayashi, Proc. Natl. Acad. Sci. U. S. A. 101:13726–13731, 2004, doi:10.1073/pnas.0405082101). Here, we discovered that this enzyme catalyzed the reverse reaction, synthesizing N-benzylformamide (NBFA) from benzylamine and formate. The reverse reaction proceeded only in the presence of high substrate concentrations. The effects of pH and inhibitors on the reverse reaction were almost the same as those on the forward reaction, suggesting that the forward and reverse reactions are both catalyzed at the same catalytic site. Bisubstrate kinetic analysis using formate and benzylamine and dead-end inhibition studies using a benzylamine analogue, aniline, revealed that the reverse reaction of this enzyme proceeds via an ordered two-substrate, two-product (bi-bi) mechanism in which formate binds first to the enzyme active site, followed by benzylamine binding and the subsequent release of NBFA. To our knowledge, this is the first report of the reverse reaction of an amine-forming deformylase. Surprisingly, analysis of the substrate specificity for acids demonstrated that not only formate, but also acetate and propionate (namely, acids with numbers of carbon atoms ranging from C1 to C3), were active as acid substrates for the reverse reaction. Through this reaction, N-substituted carboxamides, such as NBFA, N-benzylacetamide, and N-benzylpropionamide, were synthesized from benzylamine and the corresponding acid substrates.  相似文献   

16.
The conformational characteristics of the peptide sequence X-l-Pro, where X  Gly or l-Ala and the peptide bond joining X and l-Pro is cis, are evaluated. Semi-empirical potential functions are used to estimate the contributions to the conformational energy made by the non-bonded van der Waals' and electrostatic interactions and the intrinsic torsional potentials about the NCa and CaC′ bonds. Rotations φ1 and ψ1 about the NCa and CaC′ bonds in residue X and rotation ψ2 about the CaC′ bond in l-Pro are permitted, while the angle of rotation φ2 about the NCa bond in l-Pro is fixed at 120 ° by the pyrrolidine ring. The presence of the cis peptide bond connecting X and l-Pro renders the backbone rotations φ1, ψ1 in X dependent upon the rotation ψ2 about the CaC′ bond in l-Pro. (Interdependence of rotations in neighboring residues joined by a cis peptide bond was previously observed in l-alanine oligomers.) The number of energetically allowed conformations for the Gly and l-Ala residues preceding a cis peptide bond l-Pro residue are found to be substantially reduced from those permitted when the peptide bond is trans or when l-Pro is replaced by an amino acid residue. On the other hand, ψ2 = 100 to 160 ° (cis′) and 300 to 0 ° (trans′) are found to be the lowest energy conformations of the l-Pro residue irrespective of the cis or trans conformation of the X-l-Pro peptide bond.  相似文献   

17.
A cell-free supernatant of lysates of Lactobacillus plantarum catalyses the synthesis of lipids from [2-14C]mevalonate. Of the added mevalonate, 7.5% is incorporated into lipids, which were fractionated into five components. About 4% of the radioactivity in these lipids co-chromatographs with compounds shown by mass spectrometry, n.m.r. and i.r. spectroscopy to be C55 polyprenols, and about 2% co-chromatographs with a hexamer. The rest of the radioactivity is in more complex fractions. Analysis by mass spectrometry, n.m.r. and i.r. spectroscopy shows that the major C55 polyprenol is undecaprenol, accompanied by an isomer containing one reduced isoprene unit. A Kuhn–Roth degradation of [14C]polyprenols indicates that the supernatant catalyses synthesis of these compounds de novo.  相似文献   

18.
Previous work has shown that the maximum fluorescence yield from PS 2 of Synechococcus PCC 7942 occurs when the cells are at the CO2 compensation point. The addition of inorganic carbon (Ci), as CO2 or HCO3 , causes a lowering of the fluorescence yield due to both photochemical (qp) and non-photochemical (qN) quenching. In this paper, we characterize the qN that is induced by Ci addition to cells grown at high light intensities (500 mol photons m–2 s–1). The Ci-induced qN was considerably greater in these cells than in cells grown at low light intensities (50 mol photons m–2 s–1), when assayed at a white light (WL) intensity of 250 mol photons m–2 s–1. In high-light grown cells we measured qN values as high as 70%, while in low-light grown cells the qN was about 16%. The qN was relieved when cells regained the CO2 compensation point, when cells were illuminated by supplemental far-red light (FRL) absorbed mainly by PS 1, or when cells were illuminated with increased WL intensities. These characteristics indicate that the qN was not a form of energy quenching (qE). Supplemental FRL illumination caused significant enhancement of photosynthetic O2 evolution that could be correlated with the changes in qp and qN. The increases in qp induced by Ci addition represent increases in the effective quantum yield of PS 2 due to increased levels of oxidized QA. The increase in qN induced by Ci represents a decrease in PS 2 activity related to decreases in the potential quantum yield. The lack of diagnostic changes in the 77 K fluorescence emission spectrum argue against qN being related to classical state transitions, in which the decrease in potential quantum yield of PS 2 is due either to a decrease in absorption cross-section or by increased spill-over of excitation energy to PS 1. Both the Ci-induced qp (t 0.5<0.5 s) and qN (t 0.51.6 s) were rapidly relieved by the addition of DCMU. The two time constants give further support for two separate quenching mechanisms. We have thus characterized a novel form of qN in cyanobacteria, not related to state transitions or energy quenching, which is induced by the addition of Ci to cells at the CO2-compensation point.Abbreviations BTP- 1,3-bis[tris(hydroxymethyl)-methylaminopropane] - Chl- chlorophyll - Ci- inorganic carbon (CO2+HCO3 +CO3 2–) - DCMU- 3-(3,4-dichlorophenyl)-, 1-dimethylurea) - F- chlorophyll fluorescence measured at any time in the absence of a saturating flash - Fo- chlorophyll fluorescence with only the weak modulated measuring beam on - FM'- chlorophyll fluorescence during a saturating flash - FM- maximum chlorophyll fluorescence, measured in the presence of WL and FRL at the CO2-compensation point or in the presence of DCMU - FV- variable fluorescence (= FM'–F0) - FRL- supplemental illumination with far red light - MB- modulated measuring beam of the PAM fluorometer - MV- methyl viologen - PAM- pulse amplitude modulation - PFD- incident photon flux density - PS 1, 2- Photosystems 1 and 2 - QA- primary electron-accepting plastoquinione of PS 2 - qN- non-photochemical quenching of chlorophyll fluorescence - qp- photochemical quenching of chlorophyll fluorescence; rubisco-ribulose bisphosphate carboxylase/oxygenase - SF- saturating flash (600 ms duration) - WL- white light illumination  相似文献   

19.
X-Ray crystallographic analysis was performed on the compound to which had been assigned the structure 1,2,3,5-tetra-O-acetyl-4-deoxy-4-C-[(S)-ethylphosphinyl]-α-d-ribofuranose. The results showed that the compound has the proposed configuration, the five-membered ring is in the 3T2 conformation with a tendency towards the E2 form, the substituents on C-1, C-4, and P-5 are linked bisectionally, and the acetoxyl groups on C-2 and C-3 are respectively attached axially and equatorially. Based on the X-ray crystallographic and 1H-n.m.r.-spectral data, favored conformations of P-in-ring analogs of aldopentofuranose peracetates in solution are discussed.  相似文献   

20.
Depending on growth conditions, some species of purple photosynthetic bacteria contain peripheral light-harvesting (LH2) complexes that are heterogeneous owing to the presence of different protomers (containing different αβ-apoproteins). Recent spectroscopic studies of Rhodopseudomonas palustris grown under low-light conditions suggest the presence of a C 3-symmetric LH2 nonamer comprised of two distinct protomers. The software program Cyclaplex, which enables generation and data-mining of virtual libraries of molecular rings formed upon combinatorial reactions, has been used to delineate the possible number and type of distinct nonamers as a function of numbers of distinct protomers. The yield of the C 3-symmetric nonamer from two protomers (A and B in varying ratios) has been studied under the following conditions: (1) statistical, (2) enriched (preclusion of the B-B sequence), and (3) seeded (pre-formation of an A-B-A block). The yield of C 3-symmetric nonamer is at most 0.98 % under statistical conditions versus 5.6 % under enriched conditions, and can be dominant under conditions of pre-seeding with an A-B-A block. In summary, the formation of any one specific nonamer even from only two protomers is unlikely on statistical grounds but must stem from enhanced free energy of formation or a directed assembly process by as-yet unknown factors.  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