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.
Kinetic studies on the RNase T1-catalyzed transesterification of 12 dinucleoside monophosphates, Np1N2 (N1 = A, C, and U; N2 = A, C, G, and U) at pH 5, 25 °C, and 0.2 m ionic strength, revealed that the catalytic efficiency (kcatKm) for GpN substrates (H. L. Osterman, and F. G. Walz, Jr., 1978, Biochemistry, 17, 4142) was ~106-fold greater than corresponding ApNs and at least 108-fold greater than corresponding CpNs and UpNs. The catalytic activity with ApN substrates survives phenol extraction which indicates (along with other criteria) that it is intrinsic to RNase T1 and is not due to trace contamination by other nucleases. Circumstantial evidence is presented which suggests that homologous GpN and ApN substrates bind productively at different sites on the enzyme. The results of steady-state kinetic studies of RNase T1 with IpNs (N = C and U) were compared with those for GpNs and indicated that the primary effect of the guanine 2-NH2 group is to enhance substrate binding at the primary recognition site by ~2.6 kcal/mol. Values of (kcatKm) showed the order NpC > NpU (N = A, G, and I) which evidences the existence of a subsite for the leaving nucleoside group that prefers cytidine: interactions at this subsite are reflected in kcat rather than Km.  相似文献   

2.
The effect of pH on the kinetic parameters for the chloroperoxidase-catalyzed N-demethylation of N,N-dimethylaniline supported by ethyl hydroperoxide was investigated from pH 3.0 to 7.0. Chloroperoxidase was found to be stable throughout the pH range studied. Initial rate conditions were determined throughout the pH range. The Vmax for the demethylation reaction exhibited a pH optimum at approximately 4.5. The Km for N,N-dimethylaniline increased with decreasing pH, while the Km for ethyl hydroperoxide varied in a manner paralleling Vmax. Comparison of the VmaxKm values for N,N-dimethylaniline and ethyl hydroperoxide indicated that the interaction of N,N-dimethylaniline with chloroperoxidase compound I was rate-limiting below pH 4.5, while compound I formation was rate-limiting above pH 4.5. The log of the VmaxKm for ethyl hydroperoxide was independent of pH, indicating that chloroperoxidase compound I formation is not affected by ionizations in this pH range. The plot of the log of the VmaxKm for N,N-dimethylaniline versus pH indicated an ionization on compound I with a pK of approximately 6.8. The plot of the log of the Vmax versus pH indicated an ionization on the compound I-N,N-dimethylaniline complex, with a pK of approximately 3.1. The results show that chloroperoxidase can demethylate both the protonated and neutral forms of N,N-dimethylaniline (pK approximately 5.0), suggesting that hydrophobic binding of the arylamine substrate is more important in catalysis than ionic bonding of the amine moiety. For optimal catalysis, a residue in the chloroperoxidase compound I-N,N-dimethylaniline complex with a pK of approximately 3.1 must be deprotonated, while a residue in compound I with a pK of approximately 6.8 must be protonated.  相似文献   

3.
The action of ATP and its analogs as well as the effects of alkali ions were studied in their action on the ouabain receptor. One single ouabain receptor with a dissociation constant (KD) of 13 nM was found in the presence of (Mg2+ + Pi) and (Na+ + Mg2+ + ATP). pH changes below pH 7.4 did not affect the ouabain receptor. Ouabain binding required Mg2+, where a curved line in the Scatchard plot appeared. The affinity of the receptor for ouabain was decreased by K+ and its congeners, by Na+ in the presence of (Mg2+ + Pi), and by ATP analogs (ADP-C-P, ATP-OCH3). Ca2+ antagonized the action of K+ on ouabain binding. It was concluded that the ouabain receptor exists in a low affinity (Rα) and a high affinity conformational state (Rβ). The equilibrium between both states is influenced by ligands of (Na+ + K+)-ATPase. With 3 mM Mg2+ a mixture between both conformational states is assumed to exist (curved line in the Scatchard plot).  相似文献   

4.
Phosphate uptake by yeast at pH 7.2 is mediated by two mechanisms, one of which has a Km of 30 μM and is independent of sodium, and a sodium-dependent mechanism with a Km of 0.6 μM, both Km values with respect to monovalent phosphate. The sodium-dependent mechanism has two sites with affinity for Na+, with affinity constants of 0.04 and 29 mM. Also lithium enhances phosphate uptake; the affinity constants for lithium are 0.3 and 36 mM. Other alkali ions do not stimulate phosphate uptake at pH 7.2. Rubidium has no effect on the stimulation of phosphate uptake by sodium.Phosphate and arsenate enhance sodium uptake at pH 7.2. The Km of this stimulation with regard to monovalent orthophosphate is about equal to that of the sodium-dependent phosphate uptake.The properties of the cation binding sites of the phosphate uptake mechanism and those of the phosphate-dependent cation transport mechanism have been compared. The existence of a separate sodium-phosphate cotransport system is proposed.  相似文献   

5.
Adenosine triphosphate-dependent glutathione transport was characterized using inside-out vesicles made from human erythrocytes. Kinetic analysis of the glutathione disulfide (GSSG) transport showed a biphasic Line-weaver-Burk plot as a function of GSSG concentration suggesting the operation of two different processes. One phase had a high affinity for GSSG and a low transport velocity. Most active at acidic pH and at 25°C, this transport activity was easily lost during the storage of vesicles at 4°C. The Km for Mg-ATP was 0.63 mM; guanosine triphosphate (GTP) substituted for ATP gave a 340% stimulation of transport activity. Neither dithiothreitol nor thiol reagents affected this transport process. The other phase had a low affinity for GSSG and a high transport velocity. Most active at pH 7.2 and 37°C, this transport activity was stable during storage of vesicles at 4°C for several days. The Km for Mg-ATP was 1.25 mM; GTP substituted with no change in activity. Dithiothreitol increased the V but did not alter the Km, and thiol reagents inhibited the transport. These findings suggest that there are two independent transfer processes for GSSG in human erythrocytes.  相似文献   

6.
Precise oxygen equilibrium curves for carp hemoglobin were determined at 15 °C in bis-Tris buffer, and in phosphate buffer in the presence and absence of P6-inositol, and at various temperatures in phosphate buffer. Parameters of the Koshland, Némethy and Filmer (1966) (KNF) simple, sequential models (square and tetrahedral) were estimated by non-linear least-square fit of the experimental data to Hill plots. Non-, negative and positive co-operativity can be fitted by the KNF models. Considering equilibrium arguments, K2ABKBB, the KNF parameter governing the co-operativity of the system, predicts a symmetry conserved mode of action in regions of high, positive co-operativity, and a symmetry non-conserved mode of action in regions of low, non- or negative co-operativity. The simple, sequential, square KNF model fits better the Hill plots than does the simple, sequential, tetrahedral KNF model. From the effect of temperature on carp hemoglobin in phosphate buffer, the heats and entropies of the subunit interaction parameter, K2ABKBB, and of the oxygenation parameters, KBBKxBKtAB and K32BBKxBKtAB, for the square and tetrahedral models, respectively, were calculated and show the square model to account well for previously published data on the carp hemoglobin molecule. This study indicates that the KNF model, in its simplest form, is capable of explaining many of the functional properties of cooperative systems, as opposed to the Monod, Wyman and Changeux (1965) model which seems only to be a special case of the KNF model in regions of high, positive co-operativity.  相似文献   

7.
Lussier P. E., Podesta R. B. and Mettrick D. F. 1982. Hymenolepis diminuta: the non-saturable component of methionine uptake. International Journal for Parasitoiogy12: 265–270. The concentration dependence of in vitro unidirectional methionine influx by Hymenolepis diminuta was analysed by the relation: J = (JmCb)(Kt + Cb) + Kd(Cb), where Jm is the maximum uptake rate, Kt is the the apparent affinity constant and Cb is the medium substrate concentration. The linear component was separated using an asymptotic least squares curve fitting procedure and the resulting constant, Kd, is thought to be an apparent permeability coefficient. Kd may be a reflection of a simple diffusive component, a second mediated component or a combination of a passive and mediated influx. The low Q10 value of the Kd's for methionine uptake (Q10 = 1.31) indicated that this component is probably a reflection of diffusion within the membrane. However, the decrease in the Kd component in the presence of leucine and glycine, implies that there is also a small, second, mediated component in addition to the diffusive component. Kd derived from the asymptotic portion of the concentration-flux relation was compared with the residual flux of methionine after near complete inhibition of the mediated component with leucine and glycine. The Kd component was found to be pH-sensitive, increasing as the pH decreased and was not affected by external sodium. Results indicate that the mediated component of methionine influx was accelerated by increasing external Na+ and H+ concentrations.  相似文献   

8.
Log-phase cells of Neurospora crassa, grown in standard minimal medium, possess an energy-dependent transport system for inorganic phosphate, with a K12 (at pH 5.8) of 0.123 mM and a Jmax of 1.64 mmoles/l cell water per min. Like the PO43? transport system in yeast, the Neurospora system is stimulated by high intracellular K+. In addition, it is inhibited by high extracellular salt concentrations, an effect which may be related to the known depolarization of the Neurospora plasma membrane at high salt concentrations.The most striking property of the system is its strong dependence upon the extracellular pH. From pH 4.0 to pH 7.3, the Jmax remains essentially constant but the K12 increases nearly 400-fold, from 0.01 to 3.62 mM. The increase cannot be accounted for by a single system with a preference for H2PO4? (which would show only a 3-fold increase in apparent K12 over this pH range) nor by two systems with different affinities and pH optima (which would display nonlinear double-reciprocal plots at intermediate pH values). It can be explained, however, by a model in which OH? or H+ is assumed to act as a modifier of the transport system, altering its affinity for substrate.  相似文献   

9.
Hydrolysis of benzyloxycarbonyl-GlyGlyPhe by nitro(Tyr 248)carboxypeptidase A over the pH range 4.88–8.04 has been examined. The nitroenzyme retains appreciable activity near pH 6.5, and the limiting value of Km is scarcely affected. The peptidase activity has a pH dependence characterized by the following parameters: pKE1 of 6.37 ± 0.19 and pKE2 of 6.60 ± 0.17 in kcatKm, and apparent pK of 5.59 ± 0.06 in Kcat. A spectroscopic pK of 6.75 ± 0.01, attributable to the nitro-Tyr 248 residue, has been determined. This correlates with the base-limb pKE2 in the kcatKm profile, which appears to be shifted from a higher value, pKE2 of 9.0, for the native enzyme. The single (acid-limb) pK which characterizes the kcat profile of the native enzyme is also found to be perturbed to a lesser extent by nitration. A kinetically competent reverse protonation mechanism, based on chemical modification and crystallographic evidence for the enzyme, is described.  相似文献   

10.
Studies of the localization of the Na+-dependent sugar transport in monolayers of LLC PK1 cells show that the uptake of a methyl α-d-glucoside, a nonmetabolizable sugar which shares the glucose-galactose transport system, occurs mainly from the apical side of the monolayer. Kinetics of [3H]phlorizin binding to monolayers of LLC PK1 cells were also measured. These studies demonstrate the presence of two distinct classes of receptor sites. The class comprising high affinity binding sites had a dissociation constant (Kd) of 1.2 μM and a concentration of high affinity receptors of 0.30 μmol binding sites per g DNA. The other class involving low affinity sites had a Kd of 240 μM with the number of binding sites equal to 12 μmol/g DNA. Phlorizin binding at high affinity binding sites is a Na+-dependent process. Binding at the low affinity sites on the contrary is Na+-independent. The mode of action of Na+ on the high affinity binding sites was to increase the dissociation constant without modifying the number of binding sites. The Na+ dependence and the matching of Kd for high affinity binding sites with the Ki of phlorizin for the inhibition of methyl α-d-glucoside strongly suggest that the high affinity phlorizin binding site is, or is part of the methyl α-d-glucoside transport system. Binding studies from either side of the monolayer also show that the binding of phlorizin at the Na+ dependent high affinity binding sites occurs mainly from the apical rather than the basolateral side. The specific location of the Na+-dependent sugar transport system in the apical membrane of LLC PK1 cells is, therefore, another expression of the functional polarization of epithelial cells that is retained under tissue culture condition. In addition, since this sugar transport almost disappears after the cells are brought into suspension, it can be used as a marker to study the development of the apical membrane in this cell line.  相似文献   

11.
12.
Previous communications from this laboratory have indicated that there exists a thiamine-binding protein in the soluble fraction of Saccharomyces cerevisiae which may be implicated to participate in the transport system of thiamine in vivo.In the present paper it is demonstrated that both activities of the soluble thiamine-binding protein and thiamine transport in S. cerevisiae are greatest in the early-log phase of the growth and decline sharply with cell growth. The soluble thiamine-binding protein isolated from yeast cells by conventional methods containing osmotic shock treatment appeared to be a glycoprotein with a molecular weight of 140 000 by sodium dodecyl sulfate polyacrylamide gel electrophoresis. The apparent Kd of the binding for thiamine was 29 nM which is about six fold lower than the apparent Km (0.18 μM) of thiamine transport. The optimal pH for the binding was 5.5, and the binding was inhibited reversibly by 8 M urea but irreversibly by 8 M urea containing 1% 2-mercaptoethanol. Several thiamine derivatives and the analogs such as pyrithiamine and oxythiamine inhibited to similar extent both the binding of thiamine and transport in S. cerevisiae, whereas thiamine phosphates, 2-methyl-4-amino-5-hydroxymethylpyrimidine and O-benzoylthiamine disulfide did not show similarities in the effect on the binding and transport in vivo. Furthermore, it was demonstrated by gel filtration of sonic extract from the cells that a thiamine transport mutant of S. cerevisiae (PT-R2) contains the soluble binding protein in a comparable amounts to that in the parent strain, suggesting that another protein component is required for the actual translocation of thiamine in the yeast cell membrane. On the other hand, the membrane fraction prepared from S. cerevisiae showed a thiamine-binding activity with apparent Kd of 0.17μM at optimal pH 5.0 which is almost the same with the apparent Km for the thiamine transport system. The membrane-bound thiamine-binding activity was not only repressible by exogenous thiamine in the growth medium, but as well as thiamine transport it was markedly inhibited by both pyrithiamine and O-benzoylthiamine disulfide. In addition, it was found that membrane fraction prepared frtom PT-R2 has the thiamine-binding activity of only 3% of that from the parent strain of S. cerevisiae.These results strongly suggest that membrane-bound thiamine-binding protein may be directly involved in the transport of thiamine in S. cerevisiae.  相似文献   

13.
Reversible flbrinogen polymer formation was examined at pH 6.6 and Γ/2 0.3. The equilibrium fraction of fibrinogen present as polymer, (Pmf)e, was determined by gel filtration for fibrinogen concentrations, FO, from 48 to 166 μm. Using FO in molarity, the experimental relation is ln [FO(Pmf)e] = 3.53 ln[FO(1 ? (Pmf)e)] + 23.73. This relation and attendant confidence limits are examined assuming, during filtration, that the original polymer population is either stable or selected polymer species dissociate to monomer. The possibility that all polymers are open is excluded since the calculated microscopic association constant would then increase with FO. Acceptable models are based on the assumptions that polymers are open, with association constant Ka, until restricted by closure, with association constant Kr, at an integral degree of polymerization, n. Values are selected on the basis that interaction parameters are independent of FO and that the required molar decrease in free energy is a minimum. Assuming polymer stability, the experimental relation at 273 °K gives n = 4, KrKa = 1.2 m, and Ka = 736 m?1. Temperature dependence gives ΔH= ?16.9 kcal/mol and ΔSOa = ?48.8 e.u. KrKa indicates a relation between changes in entropy. The probability is >0.90 that KrKa ? 56 m, which indicates a greater loss of degrees of freedom on closure than on association. Conclusions are not altered by the assumption that only the closed polymer species is stable. As ionic strength is decreased at pH 6.6, Ka increases. The clotting time of an otherwise constant system decreases as system Pmf is increased.  相似文献   

14.
Luit Slooten  Adriaan Nuyten 《BBA》1984,766(1):88-97
(1) Rates of ATP synthesis and ADP-arsenate synthesis catalyzed by Rhodospirillum rubrum chromatophores were determined with the firefly luciferase method and by a coupled enzyme assay involving hexokinase and glucose-6-phosphate dehydrogenase. (2) Vm for ADP-arsenate synthesis was about 2-times lower than Vm for ATP-synthesis. With saturating [ADP], K(Asi) was about 20% higher than K(Pi). With saturating [anion], K(ADP) was during arsenylation about 20% lower than during phosphorylation. (3) Plots of 1v vs. 1[substrate] were non-linear at low concentrations of the fixed substrate. The non-linearity was such as to suggest a positive cooperativity between sites binding the variable substrate, resulting in an increased VmKm ratio. High concentrations of the fixed substrate cause a similar increase in VmKm, but abolish the cooperativity of the sites binding the variable substrate. (4) Low concentrations of inorganic arsenate (Asi) stimulate ATP synthesis supported by low concentrations of Pi and ADP about 2-fold. (5) At high ADP concentrations, the apparent Ki of Asi for inhibition of ATP-synthesis was 2–3-times higher than the apparent Km of Asi for arsenylation; the apparent Ki of Pi for inhibition of ADP-arsenate synthesis was about 40% lower than the apparent Km of Pi for ATP synthesis. (6) The results are discussed in terms of a model in which Pi and Asi compete for binding to a catalytic as well as an allosteric site. The interaction between these sites is modulated by the ADP concentration. At high ADP concentrations, interaction between these sites occurs only when they are occupied with different species of anion.  相似文献   

15.
(1) Vanadate (VO3?) fully inhibits the ATP-dependent uncoupled Ca efflux (Ca pump) in dialyzed squid axons. (2) Vanadate inhibits with high affinity. The mean apparent affinity (K12) obtained was 7 μM. (3) Inhibition by vanadate is dependent on Cao. External Ca lead to a release of the inhibitory effect. (K12 ≈ 3 mM). This antagonic effect can be reverted by increasing the vanadate concentration. Internal K+ increases the affinity of the intracellular vanadate binding site. External K+ has no effect on the inhibition. (4) Vanadate has no effect on the Nao-dependent Ca efflux component (forward Na-Ca exchange) in the absence of ATP. In axons containing ATP vanadate modified this component.  相似文献   

16.
(1) A quantitative study has been made of the binding of ouabain to the (Na+ + K+)-ATPase in homogenates prepared from brain tissue of the hawk moth, Manduca sexta. The results have been compared to those obtained in bovine brain microsomes. (2) The insect brain (Na+ + K+)-ATPase will bind ouabain either in the presence of Mg2+ and Pi, (‘Mg2+, Pi’ conditions) or in the presence of Na+, Mg2+, and an adenine nucleotide (‘nucleotide’ conditions) as is the case for the bovine brain (Na+ + K+)-ATPase. The binding conditions did not alter the total number of receptor sites measured at high ouabain concentrations in either tissue. (3) Potassium ion decreases the affinity (increases the KD) of ouabain to the M. sexta brain (Na+ + K+)-ATPase under both binding conditions. However, ouabain binding is more sensitive to K+ inhibition under the nucleotide conditions. In bovine brain ouabain binding is equally sensitive to K+ inhibition under the both conditions. (4) The enzyme-ouabain complex has a rate of dissociation that is 10-fold faster in the M. sexta preparation than in the bovine brain preparation. Because of this, the M. sexta (Na+ + K+)-ATPase has a higher KD for ouabain binding and is less sensitive to inhibition by ouabain than the bovine brain enzyme. (5) This data supports the hypothesis that two different conformational states of the M. sexta (Na+ + K+)-ATPase can bind ouabain.  相似文献   

17.
The transport of sucrose by selected mutant and wild-type cells of Streptococcus mutans was studied using washed cocci harvested at appropriate phases of growth, incubated in the presence of fluoride and appropriately labelled substrates. The rapid sucrose uptake observed cannot be ascribed to possible extracellular formation of hexoses from sucrose and their subsequent transport, formation of intracellular glycogen-like polysaccharide, or binding of sucrose or extracellular glucans to the cocci. Rather, there are at least three discrete transport systems for sucrose, two of which are phosphoenolpyruvate-dependent phosphotransferases with relatively low apparent Km values and the other a non-phosphotransferase (non-PTS) third transport system (termed TTS) with a relatively high apparent Km. For strain 6715-13 mutant 33, the Km values are 6.25·10?5 M, 2.4·10?4 M, and 3.0·10?3 M, respectively; for strain NCTC-10449, the Km values are 7.1·10?5 M, 2.5·10?4 M and 3.3·10?3 M, respectively. The two lower Km systems could not be demonstrated in mid-log phase glucose-adapted cocci, a condition known to repress sucrose-specific phosphotransferase activity, but under these conditions the highest Km system persists. Also, a mutant devoid of sucrose-specific phosphotransferase activity fails to evidence the two high affinity (low apparent Km) systems, but still has the lowest affinity (highest Km) system. There was essentially no uptake at 4°C indicating these processes are energy dependent. The third transport system, whose nature is unknown, appears to function under conditions of sucrose abundance and rapid growth which are known to repress phosphoenolpyruvate-dependent sucrose-specific phosphotransferase activity in S. mutans. These multiple transport systems seem well-adapted to S. mutans which is faced with fluctuating supplies of sucrose in its natural habitat on the surfaces of teeth.  相似文献   

18.
Cooperativity in ligand binding: a new graphic analysis.   总被引:16,自引:0,他引:16  
When analyzing binding of ligands to macromolecules, the existence of site-site interactions complicates a straightforward interpretation of the binding parameters obtained through classical analytical methods, such as the Scatchard plot. For describing site-site interactions, we propose a new parameter, the average affinity of the receptor sites, K, calculated as (BF)/(Ro?B). Plotting K as a function of fractional occupancy (BRo), reveals that: (1) at very low occupancy a limiting high K is obtained (Ke) (“empty sites” conformation); (2) when the fraction of sites filled increases above a certain threshold, K begins to fall due to increasing site-site interactions until (3) a limiting low K (Kf) is obtained (“filled sites” conformation). This method has been successfully applied to the negative cooperativity of insulin receptors.  相似文献   

19.
Initial velocity measurements were made on the ribonuclease A catalyzed hydrolysis of P-5′-Urd-2′:3′-P in the pH range 4.0–8.0 at 25 °C in 0.1 m Tris-acetate/0.1 m KCl. The pH dependence of the Michaelis constant, Km, the turnover number ks, and ksKm for P-5′-Urd-2′:3′-P were similar to those reported for Urd-2′:3′-P (5). When P-5′-Urd-2,3-P and Urd-2′:3′-P were compared under similar conditions the average difference in ks and Km indicated that these parameters were 5-fold and 23-fold lower, respectively, for P-5′-Urd-2′:3′-P. The slight difference in the pH dependence of ksKm for these two substrates can be interpreted in terms of a specific interaction of the enzyme at the 5′ position of P-5′-Urd-2′:3′-P, which permits a less exclusive dependence on the ionized state of the free enzyme in binding this substrate. The nature of the interaction of the substrate 5′-phosphomonoester group with the enzyme is discussed in terms of possible interactions with Lys-41 and His-119.  相似文献   

20.
The binding characteristics of the β-adrenergic antagonist, [3H]dihydroalprenolol, to hamster white adipocyte membranes were studied. This binding occurred at two classes of sites, one having high affinity (Kd = 1.6±1.3 nM) but low capacity (32±17 fmol/mg membrane protein) and one having low affinity but high binding capacity. While the binding at the high-affinity sites was competitively and stereoselectively displaced by both β-antagonists and β-agonists, competition at the low-affinity sites occurred only with β-antagonists and was non-stereoselective. Thus, the β-agonist (?)-isoproterenol was further used to define nonspecific binding. Under these conditions, saturation studies showed a single class of high-affinity (Kd = 1.6±0.5 nM) binding sites with a binding capacity of 53 ± 13 fmol/mg membrane protein (corresponding to 4000 ± 980 sites per cell), and independent kinetic analysis provided a Kd value of 1.9 nM. Competition experiments showed that these binding sites had the characteristics of a β1-receptor subtype, yielding Kd values in good agreement with the Kact and the Ki values found for agonist-stimulation and for antagonist-inhibition of adenylate cyclase in membranes and of cyclic AMP accumulation and lipolysis in intact cells. Furthermore, the ability of β-agonists to compete with this binding was severely depressed by p[NH]ppG. These results thus support the contention that the specific [3H]dihydroalprenolol binding sites defined as the binding displaceable by (?)-isoproterenol represent the physiologically relevant β-adrenergic receptors of hamster white adipocytes. Finally, studies of the lipolytic response of these cells to (?)-norepinephrine showed that the inhibitory effect of the α2-component of this catecholamine was apparent only when the effects of endogenous adenosine were suppressed, a result which argues against an important regulatory role for the α2-receptors in the adrenergic control of lipolysis in hamster white adipocytes.  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