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 15 毫秒
1.
Summary The amino acid permeability of membranes is of interest because they are one of the key solutes involved in cell function. Membrane permeability coefficients (P) for amino acid classes, including neutral, polar, hydrophobic, and charged species, have been measured and compared using a variety of techniques. Decreasing lipid chain length increased permeability slightly (5-fold), while variations in pH had only minor effects on the permeability coefficients of the amino acids tested in liposomes. Increasing the membrane surface charge increased the permeability of amino acids of the opposite charge, while increasing the cholesterol content decreased membrane permeability. The permeability coefficients for most amino acids tested were surprisingly similar to those previously measured for monovalent cations such as sodium and potassium (approximately 10–12–10–13 cm · s–1). This observation suggests that the permeation rates for the neutral, polar and charged amino acids are controlled by bilayer fluctuations and transient defects, rather than partition coefficients and Born energy barriers. Hydrophobic amino acids were 102 more permeable than the hydrophilic forms, reflecting their increased partition coefficient values.External pH had dramatic effects on the permeation rates for the modified amino acid lysine methyl ester in response to transmembrane pH gradients. It was established that lysine methyl ester and other modified short peptides permeate rapidly (P = 10–2 cm · s–1) as neutral (deprotonated) molecules. It was also shown that charge distributions dramatically alter permeation rates for modified di-peptides. These results may relate to the movement of peptides through membranes during protein translocation and to the origin of cellular membrane transport on the early Earth.Abbreviations DCP dicetylphosphate - DMPC dimyristoyl phosphatidylcholine - EPC egg phosphatidylcholine - LUV large unilamellar vesicle - MLV multilamellar vesicle - PLM planar lipid membrane - SUV small unilamellar vesicle - pH transmembrane pH gradient  相似文献   

2.
The temperature dependence of the partition of a neuropeptide, Substance P (SP), and its [Tyr8] analogue in a widely used membrane mimic, dodecylphosphocholine micelles, was studied by using a pulsed field gradient nmr diffusion technique. The partition coefficient was found to decrease when the temperature is increased, indicating a favorable (negative) enthalpy change upon partitioning of the peptides. Thermodynamic functions of the partitioning were determined. The enthalpy of partition ΔHpart, was found to be in the −2.5 to −3.0 kcal/mol range, which is between 2 and 3 times higher than the entropic term −TΔSpart. The free energy of partitioning is consistent with a model in which the SP peptides interact with the micelles mainly through the hydrophobic side chains of the residues Phe7, Phe8 (or Tyr8), Leu10, and Met11, and without the insertion of a major portion of the peptide into the hydrophobic core of the micelles. © 1998 John Wiley & Sons, Inc. Biopoly 45: 395–403, 1998  相似文献   

3.
The research draws on experimental and theoretical data about energetics and kinetics of mass spectrometric (MS) reactions of glycyl homopenta– ( G5 ) and glycyl homohexapeptides ( G6 ). It shows the great applicability of the methods of quantum chemistry to predict MS profile of peptides using energetics of collision induced dissociation (CID) fragment species. Mass spectrometry is among irreplaceable methods, providing unambiguous qualitative, quantitative and structural information about analytes, applicable to many scientific areas like environmental chemistry; food chemistry; medicinal chemistry; and more. Our study could be considered of substantial interdisciplinary significance, where MS proteomics is widely used. The experimental design involves electrospray ionization (ESI) and CID MS/MS. Theoretical design is based on ab initio and density functional theory (DFT) methods. Experimental MS and theoretical free Gibbs energies as well as rate constants of fragment reactions are compared. The thermodynamic encompasses gas–phase and polar continuum analysis, including polar protic and aprotic solvents within temperature T = 10–500 K; dielectric constant ε = 0–78, pH, and ionic strengths μ = 0.001–1.0 mol dm?1. There are computed and discussed 39 protonated forms of peptides at amide N– and –(NHC)= O centers; corresponding fragment ions studying their thermodynamic stability depending on experimental conditions. A correlation analysis between molecular conformations of parent ions and fragment species; their proton accepting ability and internal energy distribution is carried out. Data about ionization potentials (IPs) and electron affinities (EAs) are discussed, as well.  相似文献   

4.
The model proposed by Ichimaru for calculating transport coefficients is generalized to describe a plasma containing neutral atoms and ions with different charges. Ichimaru's model was developed for a fully ionized two-component (electrons and a single ion species) plasma with a temperature above 105 K. Taking into account several species of positive ions and neutral atoms makes it possible to extend Ichimaru's model to a partially ionized plasma. Transport coefficients calculated from different models are compared with the experimental data.  相似文献   

5.
Proton magnetic resonance data and conformational calculations of a series of model compounds containing a NH-CαH group substituted as in peptides have been used to generate a proton–proton coupling constant–dihedral angle relation for the peptide unit. For those substances used in which the dihedral angle about the N-Cα bond is not fixed, the angle distribution was calculated from conformational theory. Using eight examples in which the number of theoretical assumptions were least, the best values of the coefficients A, B, and C in the expression J(θ) = Acos2θ + B cosθ + Csin2θ were found by a least-squares procedure to be 7.9, ?1.55, and 1.35, respectively. This relation gives reasonable values for the dihedral angles ? in cyclic oligopeptide structures for which the availability of both NMR data and other structural information allow comparison. When applied to N-acetylamino acid N-methylamides having side chains extending beyond Cβ, however, agreement with the calculated conformational distribution was found for Leu, Met, and Trp, but observed values of J were larger than expected for Val, He, Phe, and Tyr, These disagreements are considered to be the result of interactions not yet taken into account in the usual conformational calculations.  相似文献   

6.
1. Lago Maggiore is a deep subalpine lake that has regained its previous oligotrophic state during the last decade after going through a process of cultural eutrophication and subsequent restoration. Autotrophic picoplankton (APP) have been studied in Lago Maggiore since 1992, with the primary production of the fraction being measured in 1994 and 1995. In 1998, we began to study the population structure and photosynthetic characteristics of APP after the restoration of the lake. We also compared the seasonal dynamics of APP and nanoplankton biomass and production. In this paper, we show the trend of annual and interannual dynamics of APP from 1993 to 1998. 2. Since 1993 and 1994, APP have almost tripled in numbers (mean values: 1993=44 × 103 cell mL–1 1998=123 × 103 cell mL–1) and in production. On average, APP produced 16 gC m–2 year–1 in 1994 and 41.2 gC m–2 year–1 in 1998, accounting for 10 and 20%, respectively, of total phototrophic production. 3. Although nanoplankton dominated in terms of biomass and production in the first 6 months of the year, APP were able to compete successfully with them in periods of P limitation. In September 1998, APP carbon fixation actually surpassed that of nanoplankton, reaching 13 mgC m–3 h–1 at 3 m. In a comparison of the daily chlorophyll‐specific photosynthesis rates of nanoplankton and APP, the latter proved to be more efficient in fixing carbon, particularly in summer and autumn. 4. While the spatial distribution of APP abundance and production followed the isotherms of 18 and 20 °C, and was strictly related to water stratification and thermocline appearance, the APP efficiency [mgC (mg chl)–1Em–2] had its maximum at the bottom of the euphotic zone, with irradiance in the range 0.5–18 μE m–2 s–1. 5. The existence of a significant regression between APP abundance and carbon fixation (P < 0.001, r2=0.92, d.f.=17; APPprod=10.6 × 103 cell mL–1 + 15) enabled us to infer APP production from cell numbers, and supports the view that picocyanobacteria production is closely dependent on its standing crop. 6. Lago Maggiore is a good example of how APP in a stabilised oligotrophic system can, under favourable conditions, increase in numbers and effectively exploit the few available resources. We can conclude that the APP population reacts to the reduction in P‐levels, but only after several years of stable low nutrient conditions.  相似文献   

7.
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   

8.
The cellular prion protein (PrPC) is a Cu2+ binding protein connected to the outer cell membrane. The molecular features of the Cu2+ binding sites have been investigated and characterized by spectroscopic experiments on PrPC-derived peptides and the recombinant human full-length PrPC (hPrP-[23-231]). The hPrP-[23-231] was loaded with 63Cu under slightly acidic (pH 6.0) or neutral conditions. The PrPC/Cu2+-complexes were investigated by extended X-ray absorption fine structure (EXAFS), electron paramagnetic resonance (EPR), and electron nuclear double resonance (ENDOR). For comparison, peptides from the copper-binding octarepeat domain were investigated in different environments. Molecular mechanics computations were used to select sterically possible peptide/Cu2+ structures. The simulated EPR, ENDOR, and EXAFS spectra of these structures were compared with our experimental data. For a stoichiometry of two octarepeats per copper the resulting model has a square planar four nitrogen Cu2+ coordination. Two nitrogens belong to imidazole rings of histidine residues. Further ligands are two deprotonated backbone amide nitrogens of the adjacent glycine residues and an axial oxygen of a water molecule. Our complex model differs significantly from those previously obtained for shorter peptides. Sequence context, buffer conditions and stoichiometry of copper show marked influence on the configuration of copper binding to PrPC. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

9.
A configuration partition function, which incorporates concepts embodied in the amphipathic helix hypothesis, has been formulated for a polypeptide in the presence of zwitterionic phospholipid. An enhanced probability is assigned to helix formation in any region of the polypeptide chain where side chains bearing charges of opposite sign will be situated on the same side of the α-helix but displaced from one another by one turn. This situation will arise when residues i ? 4 (or i ? 3) and i bear charges of opposite sign and residue i ? 4 (or i ? 3) through i are in a helical state. Illustrative calculations are performed for polypeptide chains in which the generalized nonionic amino acid residue serving as host has Zimm-Bragg parameters of σ = 10?4, s = 1. These calculations define conditions under which two interacting charged pairs can cooperate in a synergistic helix augmentation even when the two pairs are separated by significantly more than four generalized nonionic amino acid residues. Furthermore, the two interacting charged pairs, as well as the intervening amino acid residues, may become helical as one unit. Significant augmentation in helicity is observed with plausible values for the enhanced probablity assigned to helix formation for an interacting pair. This model predicts correctly that glucagon and secretin, but not vasoactive intestinal peptide, undergo a coil-to-helix trnsition in the presence of zwitterionic phospholipid. This prediction is made with plausible values for the parameter used to express the helicity enhancement. The experimental observation with zwitterionic phospholipids is the direct opposite of that seen for these three peptides in the presence of anionic lipids and detergents. In anionic lipids the amount of induced helicity is in the following order: glucagon < secretin < vasoactive intestinal peptide. Results obtained with these three peptides demonstrate that the nature of the head group of the lipid is important for lipid–protein interaction and that the resulting conformational changes can be rationalized by matrix methods.  相似文献   

10.
Biphasic aqueous‐organic systems are important reaction systems for catalytic processes. This is especially true for biocatalysis where the range of accessible products can be significantly extended. In such systems, the aqueous phase is the reactive phase in which the biocatalyst is dissolved and the organic phase is nonreactive and acts as substrate reservoir and as in situ product extraction solvent. Here, the choice of the nonreactive phase is highly important for the overall performance of the system. In this contribution, a systematic approach to solvent selection for biphasic aqueous‐organic systems is presented with respect to partition coefficients. The model reaction is the stereoselective carbon‐carbon coupling of two 3,5‐dimethoxy‐benzaldehyde molecules to (R)‐3,3',5,5'‐tetramethoxy‐benzoin catalyzed by benzaldehyde lyase (EC 4.1.2.38) from Pseudomonas fluorescens. A systematic approach to solvent selection consisting of two steps is proposed: Firstly, the conductor‐like screening model for real solvents (COSMO‐RS) is used to facilitate a fast solvent screening. Since this is an ab initio approach it allows a pre‐screening without laborious experimental input. The proposed ranking of solvents, based on the ratio of partition coefficients at infinite dilution, is a sound basis for the successive steps. Secondly, a dynamic model is fitted to experimental data in order to obtain detailed and reliable results for mass transfer and partition coefficients. Therefore, the method makes efficient use of the experimental data and substantiates quantitative results with guided experiments.  相似文献   

11.
1. Growth and emergence of the acidification tolerant stonefly Leuctra nigra were studied in four coniferous forest streams with contrasting pH. Invertebrate samples were taken bimonthly during May 1997 to April 1998. Emerging L. nigra were captured in two of the streams during April to July 1998 using pyramidal emergence traps. In addition, a laboratory experiment was conducted to evaluate growth of L. nigra in response to variations in food quality (neutral or acid conditioned leaves) and water pH. 2. The invertebrate community in low pH streams had lower diversity and density compared with neutral streams. Abundance of L. nigra was also lower, despite the dominance of this stonefly in the invertebrate community in the acidic streams. 3. Secondary production of L. nigra was 2.04 and 3.07 g DW m–2 year–1 in the neutral streams compared with 1.02 and 1.27 g DW m–2 year–1 in the acidic streams. The P/B ratio differed only slightly among streams. 4. More adult L. nigra emerged in the neutral stream compared with the acidic stream, despite the density of the pre‐emerging fifth instar nymphs of L. nigra being similar in the two streams. Adult size (expressed as wing length) and female fitness were also significantly greater in the neutral stream. 5. Laboratory growth of L. nigra nymphs was significantly lower when fed acid conditioned alder leaves rather than neutral conditioned leaves. There was no difference in growth when nymphs were fed neutral conditioned leaves in acid and neutral water. 6. Our study shows multiple, sub‐lethal effects on life history of the acidification tolerant stonefly L. nigra in coniferous streams with low pH. The results suggest that these effects could be mediated primarily by low food quality. These findings add to the existing evidence that coniferous plantation forestry has a negative impact on stream ecosystems.  相似文献   

12.
Abstract

This paper concentrates on two very important conducting polymers poly(p-phenylene) and polypyrrole. Detailed atomistic molecular models have been developed with the help of ab initio and semi-empirical quantum mechanical calculations using the Cerius2 and WinMOPAC (version 6.0) programs.

Their optimised geometry had been calculated and compared with experimental X-ray diffraction data. The simulated and experimental vibrational spectra of biphenyl as well as isolated pyrrole monomers and oligomers from n = 1 and 2, where n is the number of structural repeat units used, have been computed using the ab initio 3–21G basis set. The results obtained are compared with experimental data for the case of biphenyl and for oligomers with n = 2 to 5 for both neutral benzenoid and quinonoid oligopyrroles, from semi-empirical predictions obtained by AM1 and PM3. The trends in the computed harmonic force fields, vibrational frequencies and intensities are monitored as a function of the chain length. The data are analyzed in conjunction with the trends in computed equilibrium geometries.  相似文献   

13.
Z-Dehydrophenylalanine (ΔzPhe) possessing four oligopeptides, Boc-(L -Ala-ΔzPhe-Aib)n-OCH3 (n = 1–4: Boc, t-butoxycarbonyl; Aib, α-aminoisobutyric acid), were synthesized, and their solution conformations were investigated by 1H-nmr, ir, uv, and CD spectroscopy and theoretical CD calculation. 1H-nmr (the solvent accessibility of NH groups) and ir studies indicated that all the NH groups except for those belonging to the N-terminal L -Ala-ΔzPhe moiety participate in intramolecular hydrogen bonding in chloroform. This suggests that the peptides n = 2–4 have a 4 → 1 hydrogen-bonding pattern characteristic of 310-helical structures. The uv spectra of all these peptides recorded in chloroform and in trimethyl phosphate showed an intense maximum around 276 nm assigned to the ΔzPhe chromophores. The corresponding CD spectra of the peptides n = 2–4 showed exciton couplets with a negative peak at longer wavelengths, whereas that of the peptide n = 1 showed only weak signals. Theoretical CD spectra were calculated for the peptides n = 2–4 of several helical conformations, on the basis of exciton chirality method. This calculation indicated that the three peptides form a helical conformation deviating from the perfect 310-helix that contains three residues per turn, and that their side chains of Δz Phe residues are arranged regularly along the helix. The center-to-center distance between the nearest phenyl pair(s) was estimated to be ~ 5.5 Å. The chemical shifts of the ΔzPhe side-chain protons (Hβ and aromatic H) for the peptides n = 2–4 indicated anisotropic shielding effect of neighboring phenyl group(s); the effect also supports a regular arrangement of the Δz Phe side chains along the helical axis. © 1993 John Wiley & Sons, Inc.  相似文献   

14.
Vpr, one of the accessory gene products encoded by HIV-1, is a 96-residue protein with a number of functions, including targeting of the viral pre-integration complex to the nucleus and inducing growth arrest of dividing cells. We have characterized by 2D NMR the solution conformations of bioactive synthetic peptide fragments of Vpr encompassing a pair of H(F/S)RIG sequence motifs (residues 71–75 and 78–82 of HIV-1 Vpr) that cause cell membrane permeabilization and death in yeast and mammalian cells. Due to limited solubility of the peptides in water, their structures were studied in aqueous trifluoroethanol. Peptide Vpr59–86 (residues 59–86 of Vpr) formed an α-helix encompassing residues 60–77, with a kink in the vicinity of residue 62. The first of the repeated sequence motifs (HFRIG) participated in the well-defined α-helical domain whereas the second (HSRIG) lay outside the helical domain and formed a reverse turn followed by a less ordered region. On the other hand, peptides Vpr71–82 and Vpr71–96, in which the sequence motifs were located at the N-terminus, were largely unstructured under similar conditions, as judged by their CαH chemical shifts. Thus, the HFRIG and HSRIG motifs adopt α-helical and turn structures, respectively, when preceded by a helical structure, but are largely unstructured in isolation. The implications of these findings for interpretation of the structure–function relationships of synthetic peptides containing these motifs are discussed. © 1998 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
Hydrogenases from Desulfovibrio are found to catalyze hydrogen uptake with low potential multiheme cytochromes, such as cytochrome c3, acting as acceptors. The production of Fe-only hydrogenase from Desulfovibrio vulgaris Hildenborough was improved with respect to the growth phase and media to determine the best large-scale bacteria growth conditions. The interaction and electron transfer from Fe-only hydrogenase to multiheme cytochrome has been studied in detail by both BIAcore and steady-state measurements. The electron transfer between [Fe] hydrogenase and cytochrome c3 appears to be a cooperative phenomenon (h = 1.37). This behavior could be related to the conductivity properties of multihemic cytochromes. An apparent dissociation constant was determined (2 × 10-7 M). The importance of the cooperativity for contrasting models proposed to describe the functional role of the hydrogenase/cytochrome c3 complex is discussed. Presently, the only determined structure is from [NiFe] hydrogenase and there are no obvious similarities between [NiFe] and [Fe] hydrogenase. Furthermore, no crystallographic data are available concerning [Fe] hydrogenase. The first results on crystallization and X-ray crystallography are reported. Proteins 33:590–600, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

16.
The properties of fluorescein are substantially altered upon conjugation to nucleic acids, affecting not only the molar absorptivities and fluorescence quantum yields but also the protolytic equilibrium constant and fluorescence lifetimes. Around neutral pH, the fluorescein moiety is present as both mono- and dianion, and the pKa relating them is increased from 6.43 for free fluorescein to about 6.90 for fluorescein attached to both single- and double-stranded oligonucleotides of at least 12 bases/base pairs. This difference reflects the local electrostatic potential around the nucleic acid, which is calculated to −28 mV. The molar absorptivities and spectral responses of the conjugated fluorescein protolytic species are also determined, from which the concentrations of fluorescein–oligonucleotide conjugates can be calculated by assuming: ε494 = 62000/[1 + 10−(pH−6.90)] + 12000/[1 + 10(pH−6.90)] (M−1 cm−1). The fluorescence quantum yield of the conjugates depends, in a complex way, on temperature, environment and oligonucleotide length, sequence and conformation, and must be determined for each experimental situation. © 1998 John Wiley & Sons, Inc. Biopoly 46: 445–453, 1998  相似文献   

17.
Abstract

Aqueous two-phase systems have been studied for almost a century to separate biomolecules in harmless conditions. Proteases produced by Aspergillus tamarii URM 4634 were extracted in polyethylene glycol (PEG)/phosphate aqueous two-phase system under discontinuous and continuous (perforated rotative discs column) process. On the discontinuous process, it was evaluated the effect of operational conditions (PEG molar mass and its concentration, phosphate concentration and pH) over the partition coefficient, activity yield and purification factor. Protease partitioned to PEG-phase with partition coefficients up to 55.73. The best process parameters were 17.5% of PEG, with molar mass 8000?g·mol?1, 15% of phosphate salt at pH 6, with 113.15% of activity yield and purification factor of 2.62. Under continuous extraction, hold up data showed that 57.1% of the discontinuous phase was available for protein extraction. Further, separation achieved 90.0% of efficiency. The yields surpassed 100% in almost all runs, and the best purification factor was 1.84, with both flows of 2?mL·min?1. Thus, the best operational conditions reached an activity yield of 95.3% and 90.0% of separation efficiency. Hence, aqueous two-phase system PEG/phosphate extraction is an efficient process for separation of proteases produced by Aspergillus tamarii URM 4634, under continuous extraction likewise under discontinuous process.  相似文献   

18.
The thermal triple helix–coil transition of covalently bridged collagenlike peptides with repeating sequences of (Ala-Gly-Pro)n, n = 5–15, was studied optically. The peptides were soluble in water/acetic acid (99:1) and were found to form triple-helical structures in this solvent system beginning with n = 8. The thermodynamic analysis of the transition equilibrium curves for n = 9–13 yielded the parameters ΔH°s = ?7.0 kJ per tripeptide unit, ΔS°s = ?23.1 J deg?1 mol?1 per tripeptide unit for the coil-to-helix transition, and the apparent nucleation parameter σ ? 5 × 10?2. It was suggested through double-jump temperature experiments that the rate-limiting step during refolding is not only influenced by the difficulties of nucleation, but also by cistrans isomerization of the Gly-Pro peptide bond.  相似文献   

19.
Conformational studies of enkephalins are hampered by their high flexibility which leads to mixtures of quasi-isoenergetic conformers in solution and makes NOEs very difficult to detect in NMR spectra. In order to improve the quality of the NMR data, Leu–enkephalin was synthesized with 15N-labelled uniformly on all amide nitrogens and examined in a viscous solvent medium at low temperature. HMQC NOESY spectra of the labelled Leu–enkephalin in a DMSOd6/H2O mixture at 275 K do show numerous NOEs, but these are not consistent with a single conformer and are only sufficient to describe the conformational state as a mixture of several conformers. Here a different approach to the structure–activity relationships of enkephalins is presented: it is possible to analyse the NMR data in terms of limiting canonical structures (i.e. β- and γ-turns) and finally to select only those consistent with the requirements of δ selective agonists and antagonists. This strategy results in the prediction of a family of conformers that may be useful in the design of new δ selective opioid peptides. © 1998 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
A new addition method is described in this study for calculating the partition coefficients of peptides. LogP and logD values of peptides are calculated by summing the contributions of the component amino acids. The final models are derived from a multivariate linear regression analysis of 219 peptides with known experimental data. The standard errors in a leave-one-out cross-validation are 0.23 and 0.24 log units for the logP and logD values, respectively. The predictive ability of the model is tested by an extra set of ten peptides, and the self-consistency of the model is further demonstrated by a new validation procedure called the evolution test. The parameters obtained in regression could be used as hydrophobicity scales for amino acids. The application of such hydrophobicity scales has also been discussed.  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