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.
SANS studies of concentrated protein solutions. I. Bovine serum albumin   总被引:1,自引:0,他引:1  
R Nossal  C J Glinka  S H Chen 《Biopolymers》1986,25(6):1157-1175
Small-angle neutron scattering (SANS) was used to examine concentrated bovine serum albumin solutions of up to 20% protein w/v. At higher protein concentrations, scattering data show distinct features that can be ascribed to strong intermolecular interactions. Differential scattering cross-sections are fitted to a theoretical model of interparticle potential consisting of a hard core plus an exponentially decaying “tail.” For moderate ionic strength (0.03M K Acetate, pH 5.9), the intermolecular interaction agrees with the double-layer repulsive part of the well-known DLVO (Derjaguin, Landau, Verwey, Overbeek) theory for interacting colloidal particles. We thus demonstrate that it is possible to determine size parameters and the surface charge of protein molecules in dense solutions. At high salt concentrations (≥0.2M NaCl) data can be fitted by the same potential model, although interpretation in terms of DLVO theory is not possible. Even in this case, however, “effective” molecular size and potential parameters can be determined.  相似文献   

2.
Salt induced dissociation of protamine, poly(L-lysine) and poly(L-arginine) from DNA was measured by relative light scattering at theta = 90 degrees and/or centrifugation. Dissociation of histones from DNA was studied using relative light scattering and intrinsic tyrosine fluorescence. Protamine was dissociated from DNA at 0.15 M MgCl2 (ionic strength mu = 0.45) or 0.53 M NaCl (mu = 0.53) based on light scattering data and at approximately 0.2 M MgCl2 (mu = 0.6) or 0.6 M NaCl based on centrifugation data. NaCl induced dissociation of poly(Lys) or poly(Arg) from natural DNAs measured by light scattering did not depend on the guanine plus cytosine content. To dissociate poly(Arg) from DNA higher ionic strength using NaCl, MgCl2, or CaCl2, similar ionic strength using NaClo4, and lower ionic strength using Na2SO4 was needed then to dissociated poly(Lys). Both the decrease in light scattering and the enhancement of tyrosine fluorescence of chromatin occurred between 0.5 and 1.5 M NaCl when histones were dissociated.  相似文献   

3.
We report dynamic light scattering measurements over a wide range of scattering vectors for fractionated samples of porcine submaxillary mucin (PSM) glycoproteins in two different solvents: 0.1M NaCl, and 6M GdnHCl. The relaxation spectrum has been successfully resolved into a slow mode corresponding to pure translational diffusion and a fast mode containing information on the relaxation times for intramolecular motion. Analysis of the slow mode permits a light scattering evaluation of the polydispersity of these high molecular weight mucin glycoprotein fractions. Determination of the longest intramolecular relaxation times tau 1 shows that these are much longer for the PSM fractions in 0.1M NaCl compared to 6M GdnHCl. These data are consistent with earlier studies showing that the chain conformation is the same in both solvents, but that in 0.1M NaCl, the PSM glycoprotein undergoes a self-association process that is end-to-end in nature. Since the tau 1 value is intimately related to the viscoelastic behavior of PSM solutions and gels, it is interesting to speculate that the end-to-end association process may be physiologically important.  相似文献   

4.
Sensitivity of wheat (Triticum aestivum L.) seedlings to salt stress was investigated as dependent on light conditions of plant growing. In two-week-old seedlings grown on salt-free medium, aboveground organs were detached from the roots and subjected to a brief stress at different concentrations of NaCl. The extent of salt stress effect expressed as a decrease in the rate of the photosynthetic release of oxygen and the relative content of water and chlorophyll in the leaves greatly depended on light conditions of growing. The plants grown at low light intensity were notable for a greater sensitivity to NaCl in the medium. Plant responses to salt stress were different at low and high salt concentrations. At low NaCl concentrations (0.05–0.10 M) in the solution, in plants grown at low light intensity, the rate of photosynthesis calculated per unit of chlorophyll increased. This effect was not observed in plants grown at the higher light intensity. At high NaCl concentrations (0.2–0.4 M) in the medium, the rate of photosynthesis rapidly decreased in all the types of treatment, with the effect being most pronounced in plants grown at low light intensity. The obtained results suggest a narrow range of NaCl concentrations with an optimum at 0.1 M positively affecting the wheat seedlings physiological state upon salt stress development depending on light conditions of plant growing.  相似文献   

5.
The major membrane polar lipid components in Halobacterium cutirubrum are the diphytanyl ether analogues of phosphatidylglycerol phosphate, glycolipid sulfate, phosphatidylglycerol and phosphatidylglycerol sulfate. Dispersions of total polar lipids in water formed large birefringent liposomes showing concentric lipid bilayers in the elctron microscope; they behaved as ideal osmometers in KCl or NaCl solutions in the concentration range 0.005–0.2 M. At concentrations above 0.2 M KCl the liposomes shrank to spherical particles which were much less birefringent, showed no distinct bilayer structures by electron microscopy, and no longer behaved as ideal osmometers. Dispersions of phosphatidylglycerol phosphate, phosphatidylglycerol or phosphatidylglycerol sulfate alone did not behave as osmometers at any concentration of KCl or NaCl, but glycolipid sulfate alone or mixed with phosphatidylglycerol phosphate or phosphatidylglycerol phosphate + phosphatidylglycerol sulfate showed ideal osmometer behavior in 0.005–0.2 M KCl or NaCl. The highly negatively charged total polar lipids of H. cutirubrum thus can form stable lipid bilayers only at low ionic concentrations (0.005–0.2 M), much lower than the salt concentration (4 M) of the growth medium, and the presence of glycolipid sulfate is essential. Stability of the membrane in 4 M salt appears to require direct participation of the protein components.  相似文献   

6.
Biotinylated and non-biotinylated copolymers of poly(ethylene oxide) (PEO) and poly(2-(diethylamino)ethyl methacrylate) (PDEAEMA) were synthesized by the atom transfer radical polymerization technique. The chemical compositions of the copolymers as determined by NMR are represented by PEO(113)PDEAEMA(70) and biotin-PEO(104)PDEAEMA(93), respectively. The aggregation behavior of these polymers in aqueous solutions at different pHs and ionic strengths was studied using a combination of potentiometric titration, dynamic light scattering, static light scattering, and transmission electron microscopy. Both PEO-b-PDEAEMA and biotin-PEO-b-PDEAEMA diblock copolymers form micelles at high pH with hydrodynamic radii (R(h)) of about 19 and 23 nm, respectively. At low pH, the copolymers are dispersed as unimers in solution with R(h) values of about 6-7 nm. However, at a physiological salt concentration (c(s)) of about 0.16 M NaCl and a pH of 7-8, the copolymers form large loosely packed Gaussian chains, which were not present at the low c(s) of 0.001 M NaCl. The critical micelle concentrations (cmc's) and the cytotoxicities of the copolymers were investigated to determine a suitable polymer concentration range for future biological applications. Both PEO-b-PDEAEMA and biotin-PEO-b-PDEAEMA diblock copolymers possess identical cmc values of about 0.0023 mg/g, while the cytotoxicity test indicated that the copolymers are not toxic up to 0.05 mg/g (>83% cell survival at this concentration).  相似文献   

7.
By means of the scanning differential calorimetry, x-ray diffractometry, and the dynamic light scattering, we have systematically studied the phase and packing properties of dipalmitoylphosphatidylcholine vesicles or multibilayers in the presence of ethanol. We have also determined the partial ternary phase diagram of such dipalmitoylphosphatidylcholine/water/ethanol mixtures. The directly measured variability of the structural bilayer parameters implies that ethanol binding to the phospholipid bilayers increases the lateral as well as the transverse repulsion between the lipid molecules. This enlarges the hydrocarbon tilt (by up to 23 degrees) and molecular area (by < or = 40%). Ethanol-phospholid association also broadens the interface and, thus, promotes lipid headgroup solvation. This results in excessive swelling (by 130%) of the phosphatidylcholine bilayers in aqueous ethanol solutions. Lateral bilayer expansion, moreover, provokes a successive interdigitation of the hydrocarbon chains in the systems with bulk ethanol concentrations of 0.4-1.2 M. The hydrocarbon packing density as well as the propensity for the formation of lamellar gel phases simultaneously increase. The pretransition temperature of phosphatidylcholine bilayers is more sensitive to the addition of alcohol (initial shift: delta Tp = 22 degrees C/mol) than the subtransition temperature (delta Ts reversible 5 degrees C/mol), whereas the chain-melting phase transition temperature is even less affected (delta Tm = 1.8 degrees C/mol). After an initial decrease of 3 degrees for the bulk ethanol concentrations below 1.2 M, the Tm value increases by 2.5 degrees above this limiting concentration. The gel-phase phosphatidylcholine membranes below Tm are fully interdigitated above this limiting concentration. The chain tilt on the fringe of full chain interdigitation is zero and increases with higher ethanol concentrations. Above Tm, some of the lipid molecules are solubilized by the bound ethanol molecules. More highly concentrated ethanol solutions (> 7 M) solubilize the phosphatidylcholine bilayers with fluid chains fully and result in the formation of mixed lipid-alcohol micelles.  相似文献   

8.
The radiation response of Chinese hamster cells (V79) exposed to a wide concentration range of Li2SO4, Na2SO4 or K2SO4 has been examined and compared with the radiation response of cells treated in an identical manner with LiCl, NaCl, or KCl solutions. At hypotonic salt concentrations, cells were radiosensitized by both the chloride and sulphate salts. At high salt concentrations, approximately greater than 0.9 M, a radioprotective effect was observed with both chloride and sulphate salts. At intermediate salt concentrations from about 0.2 to 0.9 M, the cells that were treated with the sulphate salt solutions were radioprotected; cells treated with chloride salt solutions were radiosensitized. The difference in radiation response was attributed to the difference in anions for the two types of salts used.  相似文献   

9.
V79 Chinese hamster cells were treated with hypertonic solutions of NaCl or KCl and irradiated rat various times before, during, or after exposure to the solution. In solutions of molarities between 0-2 and 0-5 M, the cellular radiosensitivity increases with the molarity of the bathing solution. At these molarities, the hypertonic solution need not be present during irradiation to sensitize cells. Furthermore, radiosensitivity of cells could be increased by exposing cells for longer times to the hypertonic solution before irradiation. At higher salt concentrations (at 1-5 to 1-8 M), significant radioprotection is observed. Survival curve data showed that this protection was characterized by an increase in DO and a decrease in n, while the survival curves of cells sensitized with 0-465 M NaCl or with lower concentrations exhibited mainly changes in DO. The 1-55 M NaCl solution must be present during radiation to give a protective effect. Prolonged exposure to the salt before irradiation reduced the amount of radioprotection afforded by the salt. The results are discussed in terms of the effects of ions on histones, cellular water structure and the cell-aging cycle.  相似文献   

10.
As glycerol was suggested as an osmotic agent in the salt tolerant Debaryomyces hansenii the concentrations of total, intracellular, and extracellular glycerol produced by this yeast was followed during growth in 4 mM, 0.68 M, and 2.7 M NaCl media. The total amount of glycerol was not directly proportional to biomass production but to the cultural salinity with maximum concentrations just prior to or at the beginning of the stationary phase. In all cultures the cells lost some glycerol to the media, at 2.7 M NaCl the extracellular glycerol even amounted maximally to 80% of the total. A distinct maximum of intracellular glycerol, related to dry weight or cell number, appeared during the log phase at all NaCl concentrations. As the intracellular calculated glycerol concentrations amounted to 0.2 M, 0.8 M, and 2.6 M in late log phase cells at 4mM, 0.68 M, and 2.7 M NaCl, respectively, whereas the corresponding analysed values for the glycerol concentrations of the media were 0.7 mM, 2.5 mM, and 3.0 mM, glycerol contributes to the osmotic balance of the cells. During the course of growth all cultures showed a decreasing heat production related to cell substance produced, most pronounced at 2.7 M NaCl. At 2.7 M NaCl the total heat production amounted to--1690 kJ per mole glucose consumed in contrast to--1200 and--1130 kJ at 4 mM and 0.68 M NaCl, respectively. The Ym-values were of an inverse order, being 129, 120, and 93 at 4 mM, 0.68 M, and 2.7 M NaCl respectively.  相似文献   

11.
The proteodermatan sulfate (PDS) of bovine skin is a low molecular weight proteoglycan with a molecular structure consisting of a protein chain and a sulfated polysaccharide chain covalently linked at the 4-serine of the protein. Static and dynamic laser light scattering methods have been used to determine the weight-average molecular weight, Mw, zeta-average radius of gyration, Rg zeta, and zeta-average translational diffusion coefficient, Dto, zeta, of bovine skin PDS. We have also characterized the two components of PDS, i.e., the protein core and the dermatan sulfate (DS) chain. (The latter contained an N-terminal-linked penta- or tetrapeptide.) Interpretation of the PDS data is complicated by the block copolymer nature of its structure. When appropriate corrections are made, our results indicate that Mw for PDS monomer is 62,000 when dissolved in 4M guanidine hydrochloride (GdnHCl), and increases to 610,000 in 0.15M NaCl. Mw for the core protein in 4M GdnHCl is 39,000, and this also increases substantially to 650,000 in 0.15M NaCl. In contrast, Mw for the DS chain is 24,000 in 0.15M NaCl, indicating that there is minimal self-association of DS in 0.15M NaCl. Thus we conclude that the self-association of PDS involves the protein core. Comparison of Rg zeta and Rh, the average hydrodynamic radius, suggests that trace amounts of aggregation persist for the PDS and its core protein even in 4M GdnHCl. This conclusion is supported by evaluation of the second moments of the dynamic light scattering correlation function. Comparisons of the observed Dto, zeta for PDS with predicted values using hydrodynamic theory are consistent with a "lollipop" conformation for the molecule.  相似文献   

12.
Formation of beta-amyloid plaques is a crucial feature of Alzheimer's disease. In the present work time resolved static light scattering was applied to investigate the size and shape of growing beta-amyloid aggregates preceding plaque formation. The beta-amyloid protein with 40 amino acid residues was used. Salt free buffer solutions and solutions with 0.15M NaCl at 37 degrees C served as the aggregation medium. The focus lay on the first 2h following initiation of the aggregation process which corresponds to the protofibril phase. Addition of the NaCl accelerated the aggregation process considerably. Scattering data from aggregation in saline solutions indicated formation of long fibers which suggest interpretation of data with the worm-like chain model. Two important results were revealed: (i) At the end of the time resolved recordings, the worm-like chain model provided a fully adequate picture for the growing aggregates. Chain stiffness is characterised in terms of the persistence length, which is close to 50 nm. The linear mass density of the growing fibers approached a value of two monomers per nm corresponding to single stranded fibers, which is in accordance with presently existing models for the aggregation of beta-amyloid. The fibers finally reached contour lengths of several thousand nanometers. (ii) The plateau values for the persistence length and linear mass density observed in the final regime are gradually approached from higher values. This observation is inconsistent with simple worm-like chains. Rather does it indicate existence of another species during the initial phase of the aggregation, in addition to monomers and fibers. Aside from further insight into fundamental aspects of beta-amyloid aggregation, time resolved static light scattering provides an appropriate tool for assay tests with drugs designed to interfere with the aggregation process.  相似文献   

13.
The critical micelle concentration (CMC) of four synthetic phosphatidylcholines (containing two hexanoyl, heptanoyl, octanoyl or nonanoyl residues respectively) in aqueous solutions have been determined by surface tension measurements. The dependence of the CMC on the chain length is discussed on the basis of the mass action model for micelle formation. For the three higher homologues a contribution of 1.08 kT per CH2 group to the standard free energy of micellisation is found. The change in this free energy in going from the dihexanoyl- to the diheptanoyllecithin is somewhat larger (1.2 kT per CH2 group).The influence of high concentrations (several moles per liter) of simple electrolytes on the CMC is interpreted as a salting-out of nonpolar solutes in water. Contrary to expectations the effects of NaCl and Lil on the CMC of dioctanoyllecithin are not additive.  相似文献   

14.
The vesicle-to-micelle transition of egg phosphatidylcholine LUVs induced by octylglucoside was studied in buffers with 0-4 M sodium chloride, sucrose or urea. We used both light scattering and fluorescent probes to follow the lipid-detergent complexes in these buffers. The vesicle-to-micelle transition process was fundamentally the same in each solute. However, the detergent-to-lipid ratio required for micelle formation shifted in ways that depended on the aqueous solute. The partitioning of octylglucoside between the vesicles and the aqueous phase was primarily determined by the change in its critical micelle concentration (cmc) induced by each solute. Specifically, the cmc decreased in high salt and sucrose buffers but increased in high concentrations of urea. Cmc for two additional nonionic detergents, decyl- and dodecyl-maltoside, and three zwittergents (3-12, 3-14 and 3-16) were determined as a function of concentration for each of the solutes. In all cases NaCl and sucrose decreased the solubility of the detergents, whereas urea increased their solubilities. The effects clearly depended on acyl chain length in urea-containing solutions, but this dependence was less clear with increasing NaCl and sucrose concentrations. The contributions of these solutes to solubility and to interfacial interactions in the bilayers, pure and mixed micelles are considered.  相似文献   

15.
Beta-lactamase from a moderately halophilic organism is expected to show salt-dependent stability. Here we examined the temperature-dependence of stability at different salt concentrations using circular dichroism (CD) and enzyme activity. NaCl showed opposing effects on melting temperature and reversibility of the thermal melting. Increasing NaCl concentration greatly increased the melting temperature from, e.g., 41 degrees C in the absence of NaCl to 61 degrees C in 3 M NaCl. Conversely, reversibility decreased from 92% to 0% in the corresponding NaCl solutions. When beta-lactamase was heated at different temperatures and NaCl concentrations, the activity recovery followed the reversibility, not the melting temperature. Heating beta-lactamase at 63 degrees C, slightly above the onset temperature of melting in 2 M NaCl and far above the melting in 0.2 M NaCl, showed a much greater recovery of activity in 0.2 M NaCl than in 2 M NaCl, again consistent with the reversibility of melting.  相似文献   

16.
Dynamic and static light scattering, CD, and optical melting experiments have been conducted on M13mp19 viral circular single-strand DNA as a function of NaCl concentration. Over the 10,000-fold range in concentration from 100 microM to 1.0 M NaCl, the melting curves and CD spectra indicate an increase in base stacking and stability of stacked regions with increased salt concentration. Analysis of dynamic light scattering measurements of the single-strand DNA solutions as a function of K2 from 1.56 to 20 X 10(10) cm-2 indicates the collected autocorrelation functions are biexponential, thus revealing the presence of two decaying dynamic components. These components are taken to correspond to (1) global translational motions of the molecular center of mass and (2) motions of the internal molecular subunits. From the evaluated relaxation rates of these components, diffusion coefficients D0 and Dplat are determined. The center of mass translational diffusion coefficient D0, varies in a nonmonotonic manner, by 10%, from 3.75 X 10(-8) to 3.39 X 10(-8) cm2/s over the NaCl concentration range from 100 microM to 1.0 M. Likewise, the radius of gyration RG, obtained from static light scattering experiments, varies by 15% from 699 to 830 A over the same NaCl range Dplat, the diffusion coefficient of the internal subunits, displays a different dependence on the NaCl concentration and decreases, by nearly 22% in a titratable fashion, from 12.46 X 10(-8) to 10.26 X 10(-8) cm2/s, when the salt is increased from 100 microM to 1.0 M. A semiquantitative interpretation of these results is provided by analysis of the light scattering data in terms of the circular Rouse-Zimm chain. Rouse-Zimm model parameters are estimated from the experimental results, assuming the circular chains are composed of a fixed number of Gaussian segments, N + 1 = 15. The rms displacement of the internal segments, b, is estimated to be the smallest (442 A) in 100 mM NaCl. Increases of b to 467 A in 100 microM and 524 A in 1.0 M NaCl are observed. Meanwhile, the hypothetical friction factor of the internal subunits, f, progressively increases as the NaCl concentration is raised. It is inferred from the evaluated Rouse-Zimm model parameters that both the static flexibility of the circular chain and diffusive displacements of the internal subunits decrease with increases in NaCl concentration from 100 mM to 1.0 M.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
Aggregated and disaggregated forms of gizzard myosin rod and its fragments in various concentrations of NaCl (0-0.30 M) at various pH (7.4-8.6) were distinguished from each other by their permeability through a Sepharose 4B column. The rod existed in three forms, namely: large aggregates impermeable to the column, small aggregates eluted at the void volume of the column and a disaggregated monomer which penetrated the column. The relative proportions of the three forms varied depending on the salt concentration and pH. The monomeric rod was detected in NaCl solutions above 0.20 M and its relative proportion at 0.25 M NaCl was larger than those of the small and large aggregates. The small aggregates of the rod were predominant at below 0.05 M NaCl and, upon decrease in pH from 8.6 to 7.4, these small aggregates in NaCl solutions between 0.10 M and 0.15 M were replaced by the large aggregates. Light meromyosin, which corresponded to the C-terminal two-thirds of the rod, existed exclusively as large aggregates in NaCl solutions below 0.15 M; increase of NaCl concentration to above 0.20 M resulted in the formation of its monomer, instead of the large aggregates. In contrast to the rod, no small aggregated form of the light meromyosin was detected. Truncated light meromyosin which had lost a small segment from either the C-terminal or N-terminal of light meromyosin was eluted only as a monomer in any NaCl concentration at any pH. It may be deduced from the above results that a small segment in the light meromyosin is requisite for the assembly of both rod and light meromyosin in the NaCl solutions below 0.15 M and that the relative proportion of small and large aggregates of the rod is determined in a pH-dependent manner by the subfragment 2 segment, the N-terminal third of the rod.  相似文献   

18.
The effects of calcium ions on the solution properties of porcine submaxillary mucin (PSM) have been investigated by static and dynamic light scattering. The weight average molecular weights of PSM fractions are unaffected by the addition of up to 0.5M CaCl2: these data are within experimental error of those for solutions in 0.1M NaCl. The distribution of relaxation frequencies derived from the dynamic data shows the existence of two distinct relaxation modes. The average relaxation times have been interpreted to yield the z-average translational diffusion coefficient and the longest intramolecular relaxation time tau1. A plot of tau1 vs the mean value of 1/Rh-3z is linear, and consistent with plots of such data recorded for PSM in 0.1m NaCl and 6M GdnHCl solutions. However, the tau values and the associated results for the mean value of R-1h-1z in 0.5M CaCl2 are smaller than those determined in 0.1M NaCl. This suggests that the conformation of PSM in CaCl2 solution is more contracted than those in the other two solvents. These results are consistent with the compact packaging of mucin in the secretary granules that have elevated Ca2+ levels.  相似文献   

19.
Solutions of Busycon canaliculatum have been studied by light scattering. In 0.05 M Trizma buffer +0.1 M NaCl at pH 7.0 at 14 degrees, the weight-average molecular weight is 8.9 X 10(6). In the presence of added CaCl2 (0.02 M), the molecular weight of the protein increases to 10.7 X 10(6), and the second virial coefficient is reduced. At pH 9.95, the molecular weights with and without 0.02 M CaCl2, are 3.7 X 10(6) and 1.3 X 10(6), respectively; and the effect of Ca++ in reducing the second virial coefficient is much greater than at pH 7.0. These results can be understood on the basis that at pH 7.0, ca++ increases the association of hemocyanin, by binding and intermolecular linkage through the carboxyl groups of protein side chains. At pH 9.95, amino groups are deprotonated and therefore also become available for Ca++ binding. The relative effect of Ca++ in enhancing the association of hemocyanin therefore becomes greater at the higher pH.  相似文献   

20.
As glycerol was suggested as an osmotic agent in the salt tolerantDebaryomyces hansenii the concentrations of total, intracellular, and extracellular glycerol produced by this yeast was followed during growth in 4 mM, 0.68 M, and 2.7 M NaCl media. The total amount of glycerol was not directly proportional to biomass production but to the cultural salinity with maximum concentrations just prior to or at the beginning of the stationary phase. In all cultures the cells lost some glycerol to the media, at 2.7 M NaCl the extracellular glycerol even amounted maximally to 80% of the total. A distinct maximum of intracellular glycerol, related to dry weight or cell number, appeared during the log phase at all NaCl concentrations. As the intracellular calculated glycerol concentrations amounted to 0.2 M, 0.8 M, and 2.6 M in late log phase cells at 4 mM, 0.68 M, and 2.7 M NaCl, respectively, whereas the corresponding analysed values for the glycerol concentrations of the media were 0.7 mM, 2.5 mM, and 3.0 mM, glycerol contributes to the osmotic balance of the cells.During the course of growth all cultures showed a decreasing heat production related to cell substance produced, most pronounced at 2.7 M NaCl. At 2.7 M NaCl the total heat production amounted to-1690 kJ per mole glucose consumed in contrast to-1200 and-1130 kJ at 4 mM and 0.68 M NaCl, respectively. TheY m -values were of an inverse order, being 129, 120, and 93 at 4 mM, 0.68 M, and 2.7 M NaCl, respectively.  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