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.
Dielectric dispersion measurements were made on aqueous solutions of a triple-helical polysaccharide schizophyllan over a wide concentration range 10-50 wt % at -45 to +30 degrees C. In the solution state, three different water structures with the different relaxation times tau were found, namely, bound water (taul), structured water (taus), and loosely structured water (tauls) in addition to free water (tauP). Structured water is less mobile and loosely structured water is nearly as mobile as free water, but bound water with taul is much less mobile, thus taul > taus > tauls greater, similar tauP. The order-disorder transition accompanies the conversion between structured water and loosely structured water. However, the species with taus remains even in the disordered state and constitutes part of bound water in the entire temperature range. In the frozen state, in addition to bulk water formed by partial melting, two mobile species existed, which were assigned to liquidlike bound water and found to be a continuation of bound water in the solution state. These relaxation time data are discussed in connection with the entropy levels of the four structures deduced from heat capacity data (cf. Yoshiba, K.; et al. Biomacromolecules 2003, 4, 1348-1356).  相似文献   

2.
Deuterium oxide solutions of schizophyllan, a triple-helical polysaccharide, undergoing an order-disorder transition centered at 17 degrees C, were studied by optical rotation (OR) and heat capacity (C(p)) to elucidate the molecular mechanism of the transition and water structure in the solution and frozen states. The ordered structure at low temperature consisted of the side chains and water in the vicinity forming an ordered hydrogen-bonded network surrounding the helix core and was disordered at higher temperature. In the solution state appeared clearly defined transition curves in both the OR and C(p) data. The results for three samples of different molecular weights were analyzed theoretically, treating this transition as a typical linear cooperative transition from the ordered to disordered states and explained quantitatively if the molecular weight polydispersity of the sample was considered. The excess heat capacity C(EX)(p) defined as the C(p) minus the contributions from schizophyllan and D(2)O was estimated. In the frozen state it increased with raising temperature above 150 K until the mixture melted. This was compared with the dielectric increment observed in this temperature range and ascribed to unfreezable water. From the heat capacity and dielectric data, unfreezable water is mobile but more ordered than free water. In the solution state, the excess heat capacity originates from the interactions of D(2)O molecules as bound water and structured water, and so forth. Thus the schizophyllan triple helix molds water into various structures of differing orders in solution and in the solid state.  相似文献   

3.
Schizophyllan exists in aqueous solution as a triple helix, which is intact at room temperature. Its aqueous solution forms some ordered structure at low temperatures but undergoes a sharp transition to a disordered structure as the temperature is raised. The transition temperature Tc is about 7 and 18°C for H2O and D2O solutions, respectively. This transition was followed by time-domain reflectometry to investigate dynamic aspects of the transition. In addition to a major peak around 10 GHz, the dielectric dispersion curve of a 20 wt % schizophyllan in D2O exhibited a small peak around 100 MHz below Tc and around 10 MHz above Tc. The major peak is due to bulk water, whereas the 100 MHz peak is assigned to “bound” or “structured” water, and that around 10 MHz to side-chain glucose residues. However, unlike usual bound water reported for biopolymer solutions, this “structured” water disappears abruptly when the temperature becomes close to Tc without accompanying a conformational transition of the main chain. The above assignment is consistent with the structure of the ordered phase derived from previous static data that it consists of side-chain glucose residues along with nearby water molecules surrounding the helix core that are interacting with each other loosely through hydrogen bonds, and spreads radially only a layer of one or two water molecules but a long distance along the helix axis. © 1995 John Wiley & Sons, Inc.  相似文献   

4.
A comparison of 17O and 2H NMR relaxation rates of water in lysozyme solutions as a function of concentration, pH/pD, and magnetic field suggests that only 17O monitors directly the hydration of lysozyme in solution. NMR measurements are for the first time extended to 11.75 T. Lysozyme hydration data are analyzed in terms of an anisotropic, dual-motion model with fast exchange of water between the "bound" and "free" states. The analysis yields 180 mol "bound" water/mol lysozyme and two correlation times of 7.4 ns ("slow") and 29 ps ("fast") for the bound water population at 27 degrees C and pH 5.1, in the absence of salt, assuming anisotropic motions of water with an order parameter value for bound water of 0.12. Under these conditions, the value of the slow correlation time of bound water (7.4 ns) is consistent with the value of 8 ns obtained by frequency-domain fluorescence techniques for the correlation time associated with the lysozyme tumbling motion in solutions without salt. In the presence of 0.1 M NaCl the hydration number increases to 290 mol/mol lysozyme at pD 4.5 and 21 degrees C. The associated correlation times at 21 degrees C in the presence of 0.1 M NaCl are 4.7 ns and 15.5 ps, respectively. The value of the slow correlation time of 4.7 ns is consistent with the calculated value (4.9 ns) for the lysozyme monomer tumbling in solution. The systematic deviations of the relaxation rates, estimated with the single-exponential approximation, from the theoretical, multiexponential nuclear (I' + 1/2) spin relaxation are evaluated at various frequencies for 17O (I = 5/2) with the first-order, linear approximation (25). All NMR relaxation data for hydrated lysozymes are affected by protein activity and are sensitive both to the ionization of protein side chains and to the state of protein aggregation.  相似文献   

5.
Differential scanning calorimetry (DSC) and nuclear magnetic resonance (NMR) spectroscopy are applied to characterize the nonfreezable water molecules in fully hydrated D2O/sphingomyelin at temperatures below 0 degrees C. Upon cooling, DSC thermogram displays two thermal transitions peaked at -11 and -34 degrees C. The high-temperature exothermic transition corresponds to the freezing of the bulk D2O, and the low-temperature transition, which has not previously been reported, can be ascribed to the freezing of the phosphocholine headgroup in the lipid bilayer. The dynamics of nonfreezable water are also studied by 2H NMR T1 (spin-lattice relaxation time) and T2e (spin-spin relaxation time obtained by two pulse echo) measurements at 30.7 MHz and at temperatures down to -110 degrees C. The temperature dependence of the T1 relaxation time is characterized by a distinct minimum value of 2.1 +/- 0.1 ms at -30 degrees C. T2e is discontinuous at temperature around -70 degrees C, indicating another freezing-like event for the bound water at this temperature. Analysis of the relaxation data suggest that nonfreezable water undergoes both fast and slow motions at characteristic NMR time scales. The slow motions are affected when the lipid headgroup freezes.  相似文献   

6.
Properties of the aqueous dispersions of n-octadecylphosphocholine are examined by differential scanning calorimetry, fluorescence depolarization, light scattering, 31P-NMR, pig pancreatic phospholipase A2 binding, and X-ray diffraction. On heating, these dispersions exhibit a sharp lamellar to micelle transition at 20.5°C. The lamellar phase consists of frozen (gel-state) alkyl chains which do not bind phospholipase A2. The kinetics of the transition are asymmetric: the micelle to lamellar transition is very slow and the lamellar to micelle transition is fast. It is suggested that the lamellar phase is a frozen chain bilayer in which the chains interdigitate.  相似文献   

7.
UHF-dielectrometry method is based on the following facts: i) there is dispersion (i.e. dependence on frequency) of the dielectric permeability epsilon; ii) bound and free water have remarkable different epsilon, mobility and dispersion regions; iii) conformational changes in a macromolecule lead to redistribution of free and bound water and to change of the amount of free water molecules. Choosing the working frequency in the region of dispersion of free water molecules (9.2 GHz) we can detect conformational changes in proteins using free water as a marker. In this work the temperature dependencies of dielectric parameters of albumin and fibrinogen solutions were obtained in the temperature interval 5-40 degrees C. In contrast to dependencies for poor solvent, temperature dependencies of dielectric parameters for protein solutions are of non-monotonous character; they have a number of peculiarities in the temperature ranges of 8-10, 22-24 and 34-36 degrees C. At these temperatures redistribution of free and bound water in protein-water system occurs due to structural changes in protein molecules. In this work the mechanism of temperature changes of spatial organisation of protein molecules was proposed. Perhaps, this mechanism is responsible for maintenance of thermal stability of the functionally active conformation of native proteins.  相似文献   

8.
A prodrug (Fig. 1(IV)) is synthesized consisting of the beta-blocker bupranolol which is covalently linked to 1, 3-dipalmitoyl-2-succinyl-glycerol. The resulting lipid-like prodrug is amphipathic and surface active. It disperses readily in H2O above 30 degrees C forming a smectic lamellar phase. This prodrug bears one positive charge at neutral pH and hence the swelling behaviour of dispersions in H2O is similar to that of charged phospholipids: the dispersions show continuous swelling with increasing water content and consequently in the excess H2O region of the phase diagram the thermodynamically most stable structure is the unilamellar vesicle. This includes oligomeric vesicles which may be defined as unilamellar vesicles containing smaller, also unilamellar vesicles entrapped in their internal aqueous compartment. The prodrug dispersions in H2O are polydisperse with vesicle sizes ranging from 0.1 micron to several micron. Sonication of these dispersions produce small unilamellar vesicles of an average size and size distribution similar to sonicated egg phosphatidylcholine dispersions. Unsonicated dispersions of the prodrug in H2O undergo reversibly sharp order-disorder transitions at 32 degrees C with an enthalpy change of delta H = 10 kcal/mol. In sonicated aqueous dispersions this phase transition is asymmetric and significantly broadened indicating that the cooperativity is markedly reduced. The peak temperature and enthalpy change of this broad transition are reduced compared to the transition observed with unsonicated dispersions. The temperature dependence of the electron spin resonance (ESR) hyperfine splitting and order parameter also reflects the order-disorder transition. From ESR spin labeling it is concluded that in sonicated dispersions the prodrug molecule is more mobile and its anisotropy of motion is reduced compared to unsonicated dispersions. This result indicates that the molecular packing in the highly curved bilayers of small unilamellar prodrug vesicles is significantly perturbed compared to bilayers of unsonicated dispersions.  相似文献   

9.
The kinetics of the lamellar (L alpha)-inverse hexagonal (HII) phase transition in diacylphosphatidylethanolamine (PE)--water systems were probed with time-resolved X-ray diffraction. Transition kinetics in the fast time regime (approximately 100 ms) were studied by initiating large temperature jumps (up to 30 degrees C) with a 50-ms electrical current pulse passed through a lipid-salt water dispersion, resulting in ohmic heating of the sample. Diffraction with a time resolution to 10 ms was acquired at the National Synchrotron Light Source. The time constant for the phase transition for 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE) was on the order of 100 ms for the largest temperature jumps recorded. Faster transition behavior was found for a 1,2-dielaidoyl-sn-glycero-3-PE mixture. The HII lattice parameters for both systems were seen to swell from an initial value commensurate with the lamellar lattice to the final equilibrium value. The rate of swelling was seen to be independent of the magnitude of the temperature jump. For small temperature jumps (less than 10 degrees C), the phase transition kinetics slow dramatically, and transition studies can readily be performed on a conventional rotating anode X-ray source. At 4 degrees C, a DOPE sample was observed to slowly convert to the hexagonal phase over the course of a week, with the decay in the lamellar intensity fitting a power law behavior over four decades of time. This power law behavior is shown to have interesting consequences to the determination of the phase transition temperature of lipid-water dispersions by conventional methods such as calorimetry.  相似文献   

10.
The formation of gel network structures during isothermal heating of whey protein aqueous dispersions was probed by mechanical spectroscopy. It was anticipated that the pathway of the sol-to-gel transition of whey protein dispersions is quite different from that of ordinary cross-linking polymers (e.g., percolation-type transition), since aqueous solutions of native whey proteins have been shown to be highly structured even before gelation, in our previous study. At 20 degrees C, aqueous dispersions of beta-lactoglobulin, the major whey protein, and those of whey protein isolate (WPI), a mixture of whey proteins, exhibited solid-like mechanical spectra, i.e., the predominant storage modulus G' over the loss modulus G", in a certain range of the frequency omega (1-100 rad/s), regardless of the presence or absence of added NaCl. The existence of the added salt was, however, a critical factor for determining transitions in mechanical spectra during gelation at 70 degrees C. beta-Lactoglobulin dispersions in 0.1 mol/dm(3) NaCl maintained the solid-like nature during the entire gelation process and, after passing through the gelation point, satisfied parallel power laws (G' approximately G" approximately omega(n)) that have been proposed for a critical gel (i.e., the gel at the gelation point) that possesses a self-similar or fractal network structure. In contrast, beta-lactoglobulin dispersions without added salt exhibited a transition from solid-like [G'(omega) > G"(omega)] to liquid-like [G'(omega) < G"(omega)] mechanical spectra before gelation, but no parallel power law behavior was recognized at the gelation point. During extended heating time (aging), beta-lactoglobulin gels with 0.1 mol/dm(3) NaCl showed deviations from the parallel power laws, while spectra of gels without added NaCl approached the parallel power laws, suggesting that post-gelation reactions also significantly affect gel network structures. A percolation-type sol-to-gel transition was found only for WPI dispersions without added salt.  相似文献   

11.
J Andrasko 《Biophysical journal》1975,15(12):1235-1243
The dependence of the spin-lattice relaxation time in the rotating frame (T1rho) on radio frequency (RF) field strength and temperature has been studied for agarose gels in order to investigate molecular motion. The results indicate the presence of slow motions with a correlation time of ca. 5-10(-6) s at room temperature. This interaction is responsible for the short spin-spin relaxation times (T2) for water protons in agarose gels and is ascribed to firmly bound water. The fraction of bound water is estimated to about 0.003 for a 7.3% agarose gel. The motion of the more mobile protons in agarose-water systems can not be characterized by single correlation time. This fraction is presumably composed of water in different motional states and some of the agarose hydroxyl protons. Higher mobilities are the most common.  相似文献   

12.
Positron lifetimes have been determined in phospholipid dispersions. In fluid phosphatidylcholines, a lifetime of 3.3 ns is found, and a lifetime of 2.8 ns is found for frozen phosphatidylcholines. In dispersions where fluid and frozen phases coexist due to lateral phase separation, an intermediate lifetime is found.  相似文献   

13.
Water proton spin-lattice relaxation is studied in dilute solutions of bovine serum albumin as a function of magnetic field strength, oxygen concentration, and solvent deuteration. In contrast to previous studies conducted at high protein concentrations, the observed relaxation dispersion is accurately Lorentzian with an effective correlation time of 41 +/- 3 ns when measured at low proton and low protein concentrations to minimize protein aggregation. Elimination of oxygen flattens the relaxation dispersion profile above the rotational inflection frequency, nearly eliminating the high field tail previously attributed to a distribution of exchange times for either whole water molecules or individual protons at the protein-water interface. The small high-field dispersion that remains is attributed to motion of the bound water molecules on the protein or to internal protein motions on a time scale of order one ns. Measurements as a function of isotope composition permit separation of intramolecular and intermolecular relaxation contributions. The magnitude of the intramolecular proton-proton relaxation rate constant is interpreted in terms of 25 +/- 4 water molecules that are bound rigidly to the protein for a time long compared with the rotational correlation time of 42 ns. This number of bound water molecules neglects the possibility of local motions of the water in the binding site; inclusion of these effects may increase the number of bound water molecules by 50%.  相似文献   

14.
The influences of 1-octanol and 1-decanol on aqueous multilamellar dispersions of 1-hexadecanoyl(octadecanoyl)-2-[2H31]hexadecanoyl-sn-glycero -3-phosphorylcholine (PC-d31)/cholesterol (3:1) have been examined using 2H-NMR. The gel to liquid crystalline phase transition of the PC-d31/cholesterol dispersion is modulated by the addition of 1-alkanol, which reduces the onset temperature and increases the width of the transition. 1-Octanol has a greater effect on the transition onset and completion temperatures than does 1-decanol, as determined from analysis of the temperature-dependent 2H-NMR spectra. 2H-NMR C-2H bond order parameters as a function of phospholipid acyl chain position at 60 degrees C, where all dispersions are fully liquid crystalline, have been calculated from the depaked spectra. 1-Decanol reduces the phospholipid order by only 2%. This can be attributed to the lower effective cholesterol concentration in the 1-alkanol/PC-d31/cholesterol dispersions. 1-Octanol, however, reduces the phospholipid order by 10% at 60 degrees C. Correlations between the effects of 1-octanol and 1-decanol on phospholipid order parameters and phospholipid/cholesterol phase transitions are discussed.  相似文献   

15.
For phospholipid membranes with zwitterionic head groups, the dipole can be considered as a specific label for tracing the changes in the dynamic behaviour of this region of the bilayer in its various phases. Measurements of the dielectric properties of fully hydrated 1,2-dipalmitoyl-sn-glycero-3-phosphocholine bilayers in the frequency range 1--50 MHz show a dispersion which is attributed to the motion of the phosphocholine dipoles in the plane of the bilayers. When the temperature is varied, both the permittivity and loss factor increase sharply at the pretransition (35 degrees C) and the main transition (42 degress C). The relaxation time and amplitude were also determined for this dispersion and these further reflect the structural changes occurring with temperature. The relaxation times varied between 4 ns at 30 degrees C and 2.3 ns at 50 degrees C. Due to steric hindrances a restriction in the angle of head group rotation occurs at lower temperatures but is greatly reduced above the main transition.  相似文献   

16.
Dilauroylphosphatidylethanolamine dispersion forms a crystalline phase at physiological pH and temperature and in the presence of excess water. This phenomenon was observed and studied by differential scanning calorimetry, scanning densitometry and X-ray diffraction. The crystalline phase is stable at pH 5.5-9.5 and below 40 degrees C. The crystalline phase formed at pH 5.5 and pH 9.5 index according to orthorhombic cells with a = 9.41, b = 8.15, c = 46.0 and a = 9.33, b = 8.05, c = 45.8 (A), respectively. Around 43 degrees C, the crystalline phase is transformed into a multilayer liquid crystal phase. Cooling from 44 degrees C results in the disappearance of the original transition at 43 degrees C and the appearance of a second transition at around 30 degrees C. Below 30 degrees C the lipid forms a gel phase. This gel phase is metastable at pH 5.5 and a crystalline phase may be recovered from it by dispersing or aging methods. Suspensions of dilauroylphosphatidylethanolamine show similar phase transition behaviour at pH 5.5 and pH 9.5, although the transitions are somewhat broader at the higher pH. The thermotropic phase behaviour of dilauroylphosphatidylethanolamine dispersions may be governed by changes in hydration.  相似文献   

17.
P Mazur  K W Cole 《Cryobiology》1989,26(1):1-29
The cause of slow freezing injury and the basis of the protection by solutes like glycerol are subjects of debate. During slow freezing, cells are sequestered in unfrozen channels between ice crystals that grow by removing pure water from the channels. As a consequence, the solute concentration in the channels rises and the volume of liquid in the channels progressively decreases. The rise in solute concentration, in turn, causes the cells to progressively shrink osmotically. Until recently cryobiologists have ascribed slow freezing injury to either the rise in solute (electrolyte) concentrations in the channels or to the consequent cell shrinkage, rather than to the decrease in the of the channels. Although ordinarily reciprocally coupled, it is possible to separate the composition of the channels from their size, or more precisely from the magnitude of the unfrozen fraction, by suspending cells in NaCl/cryoprotectant solutions in which the mole ratio of the two is held constant, but the molality of the NaCl is allowed to vary below and above isotonic. When human red cells are frozen in such solutions to temperatures that produce given NaCl concentrations (ms), but varying unfrozen fractions (U), survival at low U is found to be strongly dependent on U but independent of ms. At higher values of U, survival becomes inversely dependent on both ms and U. Although cell volume during freezing is independent of the NaCl tonicity in the solution, the cells in the several solutions differ in volume both prior to the onset of freezing and after the completion of thawing. We have now examined and compared the effect of returning the thawed cells to isotonic solutions and isotonic volume or nearly so, and find that there is little change in survival after exposure to low U, but that survival after exposure to high U values exhibits substantially increased sensitivity to ms, a sensitivity that is probably a manifestation of posthypertonic hemolysis. Low values of U were in general attained by the use of solutions with low tonicities of NaCl, and as a consequence cells frozen to low U values had larger volumes prior to freezing than cells frozen to higher U values. The significance of this confounding is discussed.  相似文献   

18.
Among classes of sugar transporters, there exists a comparatively new family of transporters named SWEET transporters (semi-SWEETS in bacteria) that are uniport transmembrane proteins. It is hypothesized that sugar is transported from the extracellular side (via outward-open state) to intracellular side (inward-open state) through intermediate occluded state (both extracellular and intracellular gates closed). In our study, extensive unbiased all-atom molecular dynamics simulations were carried out with the outward-open and inward-open conformations to study this transition mechanism. We find that after 100?ns, the outward-open structure without sugar bound starts changing to the occluded form leading to closure of extracellular gates stabilized by electrostatic and hydrophobic interactions. Further simulations (up to 7?μs) have led to a transition toward the inward-open form and suggest that there exists more than one intermediate occluded conformation. We have also performed 5-μs simulations on the glucose-docked structure to identify the putative substrate-bound translocation pathway. Glucose binds to semi-SWEET with strong hydrogen bonds to Asn66 and Trp50. Comparative simulations of substrate bound, and unbound forms suggested that glucose, the putative substrate, facilitates relatively rapid conformational transitions. For the first time, we captured the release of glucose to the cytosol, in this family of transporters. We find that prior to release of glucose, the glucose forms interactions with polar residues near the intracellular gate which may facilitate its release. The distance between the residues Asn31 and Gly34 of the other protomer was found to play a decisive role in the transport of glucose to the cytoplasmic side.  相似文献   

19.
Heat capacity measurements were made on aqueous solutions of a triple-helical polysaccharide schizophyllan by precision adiabatic calorimetry over a wide range of concentrations 30.45-90.93 wt % at temperatures between 5 and 315 K. The heat capacity curves obtained were divided into four groups depending on the weight fraction of schizophyllan w regions I-IV. In region I, triple-helices with the sheath of bound water, structured water, and loosely structured water forming layers around the helix core are embedded in free water. In region II, there is no free water, and loosely structured water decreases until it vanishes, but structured water stays constant with increasing w. In region III, bound water remains unaffected, but structured water decreases with increasing w by overlapping each other. Finally, in region IV, only schizophyllan and bound water exist, the latter decreasing upon increasing w. The maximum thickness of each layer is 0.18(3) nm for bound water, 0.13(4) nm for structured water, and 0.23(6) nm for loosely structured water, and these layers of water are at the enthalpy levels of 53%, 93.7%, and nearly 100%, respectively, between ice (0%) and free water (100%).  相似文献   

20.
The rotational diffusion behavior of phosphorus metabolites present in calf lens cortical and nuclear homogenates was investigated by the NMR technique of 31P off-resonance rotating frame spin-lattice relaxation as a means of assessing the occurrence and extent of phosphorus metabolite-lens protein interactions. 31P NMR spectra of calf lens homogenates were obtained at 10 and 18 degrees C (below and above the cold cataract phase transition temperature, respectively) at 7.05 T. Effective rotational correlation times (tau 0,eff) for the major phosphorus metabolites present in cortical and nuclear bovine calf lens homogenates were derived from nonlinear least-squares analysis of R vs omega e (spectral intensity ratio vs precessional frequency about the effective field) data with the assumption of isotropic reorientational motion. Intramolecular dipole-dipole (1H-31P, 31P-31P), chemical shift anisotropy (CSA), and solvent (water) translational intermolecular dipole-dipole (1H-31P) relaxation contributions were assumed in the analyses. In those cases where the limiting value of the spectral intensity ratio failed to reach unity at large offset frequency, a modified formalism incorporating chemical exchange mediated saturation transfer between two sites was used. Values of tau 0,eff for phosphorus metabolites present in the cortex varied from a low of ca. 2 ns [L-alpha-glycero-phosphocholine (GPC)] to a high of 12 ns (alpha-ATP) at 10 degrees C, whereas at 18 degrees C the range was from ca. 1 to 9 ns. For the nucleus the tau 0,eff values ranged from ca. 3 ns (GPC) to 41 ns (Pi) at 10 degrees C; at 18 degrees C the corresponding values ranged from 4 to 39 ns. For PME (phosphomonoester; in lens the predominant metabolite is L-alpha-glycerol phosphate) at 18 degrees C evidence was obtained for binding and subsequent exchange with solid like protein domains. The diversity in tau 0,eff values for lenticular phosphorus metabolites is suggestive of differential binding to more slowly tumbling macromolecular species, most likely lens crystallin proteins. Corresponding measurement of tau 0,eff values for the mobile protein fraction present in calf lens cortical and nuclear homogenates at 10 and 18 degrees C, by 13C off-resonance rotating frame spin-lattice relaxation, provided average macromolecular correlation times that were assumed to represent the bound metabolite state. A fast-exchange model (on the T1 time scale), between free and bound forms, was employed in the analysis of the metabolite R vs omega e curves to yield the  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