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.
The physical and chemical characterization of horse serum butyrylcholinesterase has been extended. The results show that the enzyme is a glycoprotein containing about 20% carbohydrate by weight. Mannose, glucosamine, galactose, and sialic acid are the sugar residues found. The extinction coefficient of butyrylcholinesterase, E1cm1% at 280 nm, was found to be 15.2 ± 0.3 by dry weight determination. The molecular weight of the protein in dilute phosphate buffer was determined to be (31.7 ± 1.2) × 104 by high speed equilibrium sedimentation with a redetermined partial specific volume of 0.723 ± 0.003 ml/g. Subunit molecular weights for the dissociated protein were found to be (7.9 ± 0.4) × 104 and (8.1 ± 0.1) × 104, respectively, in guanidine hydrochloride and in a solution at pH 11.8. The subunit molecular weight was also estimated to be (8.8 ± 0.2) × 104 by analytical sodium dodecyl sulfate-gel electrophoresis. This apparently higher subunit molecular weight from dodecyl sulfate gels is expected for glycoproteins containing significant amounts of carbohydrate. No free sulfhydryl group was detected, even though there are six half-cystines in each subunit. Therefore, it seems likely that there are three pairs of disulfide bonds per subunit. The available data indicate that native butyrylcholinesterase is a tetrameric glycoprotein consisting of subunits of equal molecular weight.  相似文献   

2.
Bacillus macerans enzyme (BME)-derived high molecular weight dextrins, which are by-products in the course of the industrial production of cylodextrins, were isolated and their chemical structures were characterized.Dextrin I was obtained in a yield of about 24% from BME-hydrolyzate (a mixture of dextrin and cylodextrins, 50% each) of potato starch by fractionation with an ultrafiltrator having a membrane of cut-off molecular weight 2.0 × 104. Dextrin II was obtained in a yield of about 15% from BME-hydrolyzate (a mixture of dextrins and cyclodextrins, 70 : 30) of Dextrin I by the same method.Dextrin I and II consisted of dextrin having molecular weights over 20 × 106 and dextrins having molecular weights 4 × 103−1 × 105 in the ratio of 80 : 12 and 66: 15, respectively.The results of hydrolysis by β-amylase and methylation analysis indicated that the average, exterior and interior chain lenghts of the dextrins having molecular weights over 20 × 106 and 4 × 103−1 × 105 from Dextrin I were 16.5, 8.2 and 7.3, and 11.5, 6.9 and 3.6, respectively, than those from Dextrin II were 13.6, 4.7 and 9.9, and 10.4, 5.1 and 4.3, respectively.  相似文献   

3.
NADP-malic enzyme (EC 1.1.1.40), which is involved in the photosynthetic C4 pathway, was isolated from maize leaf and purified to apparent homogeneity as judged by polyacrylamide gel electrophoresis. At the final step, chromatography on Blue-Sepharose, the enzyme had been purified approximately 80-fold from the initial crude extract and its specific activity was 101 μmol malate decarboxylated/mg protein/min at pH 8.4. The enzyme protein had a sedimentation coefficient (s20,w) of 9.7 and molecular weight of 2.27 × 105 in sucrose density gradient centrifugation, and molecular weight of 2.26 × 105 calculated from sedimentation equilibrium analysis. The molecular weight of the monomeric form was determined to be 6.3 × 104 by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. In the pyruvate carboxylation reaction, HCO3? proved to be the active molecular species involved. With all other substrates at saturating concentration, the following kinetic constants were obtained: Km (malate), 0.4 mm; Km (NADP), 17.6 μm; Km (Mg2+), 0.11 mm. The maize leaf malic enzyme was absolutely specific for NADP. The Arrhenius plot obtained from enzyme activity measurements was linear in a temperature range of 13 to 48 °C, and the activation energy was calculated to be 9500 cal/mol.  相似文献   

4.
Purified, water-soluble polysaccharide of Phellodendron amurense Ruprecht (P-WSPS) was obtained in highly homogeneous state. The partial specific volume and intrinsic viscosity were, respectively, 0.56 and 11.68 dL/g. The S value and molecular weight depended strongly on the P-WSPS concentration, but extrapolation to infinite dilution gave Sw,200 = 10.67 × 10?13 s?1 and M0 = 625 × 103 from velocity-sedimentation and equilibrium-sedimentation experiments, respectively, results that showed the highly branched structure of P-WSPS. The molecular weight calculated from the D0 value (1.035 × 10?7) and S0 value agreed with that from the equilibrium-sedimentation experiment to within ~±9%, showing that the diffusion method is useful for rapid estimation of molecular weight.  相似文献   

5.
This work was aimed at evaluating the gill carbonic anhydrase (CA) activity of the estuarine crab Chasmagnathus granulata exposed in vivo to cadmium, at different salinities. The in vivo effect of the specific inhibitor acetazolamide (AZ) was also assayed. Besides, the inhibition of CA activity by different heavy metals (cadmium, copper, zinc) and AZ were evaluated under in vitro conditions. For the in vivo assays, adult males were acclimated to salinities of 2.5 or 30‰. The corresponding 96-h LC50 of cadmium was 2.69 mg l−1 at 2.5‰, and >50 mg l−1 at 30‰. Cadmium only caused a significant lower CA activity than control at 2.5‰. EC50 for CA inhibition was estimated to be 1.59 mg l−1 at 2.5‰. Statistical differences in Na+ hemolymphatic levels (P<0.05) were only detected at 2.5‰, between 0 and 1.25 mg l−1 of cadmium, but no statistical differences were observed for Cl levels at any assayed salinity. As CA inhibition registered at 2.5‰ was followed by only changes in Na+ concentration, it is likely that cadmium exposure could differentially affect ions permeability, among others factors. The concentrations that inhibited in vitro 50% of enzymatic activity (IC50) were 2.15×10−5, 1.62×10−5, 3.75×10−6 and 4.4×10−10 M for cadmium, copper, zinc and AZ, respectively. The comparison with IC50 values of other aquatic species, indicates a higher CA sensitivity for C. granulata to pollutants.  相似文献   

6.
Protein methylase II (S-adenosylmethionine:protein—carboxyl methyltrans-ferase), which modifies free carboxyl residues of protein, was purified from both rat and human blood, and properties of the enzymes were studied. The pH optima for the reaction were dependent on the substrate proteins used; pH 7.0 was found with endogenous substrate, 6.1 with plasma, 6.5 with γ-globulin, and 6.0 with fibrinogen. The molecular weight of the enzymes from both rat and human erythrocytes were identical (25,000 daltons) determined by Sephadex G-75 chromatography. Partially purified enzyme from rat erythrocytes showed three peaks on electrofocusing column at pH 4.9, 5.5 and 6.0. The Km values of the enzymes from rat and human erythrocytes showed 3.1 × 10?6m and 1.92 × 10?6m at pH 6.0, 1.96 × 10?6m and 1.78 × 10?6m at pH 7.2, respectively, for S-adenosyl-l-methionine. It is also found that S-adenosyl-l-homocysteine is a competitive inhibitor for protein methylase II with Ki value of 1.6 × 10?6m.  相似文献   

7.
The molecular weight and polypeptide chain stoichiometry of the native pyruvate dehydrogenase multienzyme complex from Escherichia coli were determined by independent techniques. The translational diffusion coefficient (Do20,w) of the complex was measured by laser light intensity fluctuation spectroscopy and found to be 0.90 (±0.02) × 10?11m2/s. When this was combined in the Svedberg equation with the measured sedimentation coefficient (so20,w = 60.2 (±0.4) S) and partial specific volume (v? = 0.735 (±0.01) ml/g), the molecular weight of the intact native complex was calculated to be 6.1 (±0.3) × 106. The polypeptide chain stoichiometry (pyruvate decarboxylase: lipoate acetyltransferase: lipoamide dehydrogenase) of the same sample of pyruvate dehydrogenase complex was measured by the radioamidination technique of Bates et al. (1975) and found to be 1.56:1.0:0.78.From this stoichiometry and the published polypeptide chain molecular weights estimated by sodium dodecyl sulphate/polyacrylamide gel electrophoresis, a minimum chemical molecular weight of 283,000 was calculated. This structure must therefore be repeated approximately 22 times to make up the native complex, a number which is in good agreement with the expected repeat of 24 times if the lipoate acetyltransferase core component has octahedral symmetry. It is consistent with what appears in the electron microscope to be trimer-clustering of the lipoate acetyltransferase chains at the corners of a cube. It rules out any structure based on 16 lipoate acetyltransferase chains comprising the enzyme core.The preparation of pyruvate dehydrogenase complex was polydisperse: in addition to the major component, two minor components with sedimentation coefficients (so20,w) of 90.3 (±0.9) S and 19.8 (±0.3) S were observed. Together they comprised about 17% of the total protein in the enzyme sample. Both were in slowly reversible equilibrium with the major 60.2 S component but appeared to be enzymically active in the whole complex reaction. The faster-sedimenting species is probably a dimer of the complex, whereas the slower-sedimenting species has the properties of an incomplete aggregate of the component enzymes of the complex based on a trimer of the lipoate acetyltransferase chain.  相似文献   

8.
Receptors for luteinizing hormone/human chorionic gonadotropin (LH/hCG) have been identified in porcine, rabbit, rat, and human myometrium. To determine the estrous cycle and pregnancy related changes in the receptor capacity and affinity, radioreceptor assays were performed with membrane homogenates of porcine uterine tissues. Cycling gilts were divided into four experimental groups: I (n=6), day 1–2; II (n=5), day 6–7; III (n=5), day 11–12; and IV (n=6), day 18–20 of the estrous cycle. Pregnant pigs were divided into three experimental groups: I (n=5), day 35–40; II (n=5), day 65–70; and III (n=4), day 95–105 of pregnancy. The concentrations [femtomoles/mg protein (fmol/mg protein)] and affinities of unoccupied LH/hCG binding sites were characterized in all samples of myometrium. Receptor concentrations were highest (P<0.01) in groups II and III (19.3±2.5 and 35.8±2.1 fmol/mg protein, respectively), and was lowest in groups I and IV (5.3±1.4 and 7.5±0.7 fmol/mg protein, respectively). Receptor affinity constants (Ka) were consistent (P>0.05) throughout the estrous cycle [I, (5.1±1.5)×109; II, (3.0±0.8)×109; III, (3.2±0.9)×109; IV, 5.5±0.7×109 lm−1]. Plasma hormone concentrations of progesterone, estrogen and LH were typical of values noted at these times. During pregnancy, receptor concentrations were greatest (P<0.05) in group II (85.4±18.5 fmol/mg protein). In groups I and III receptor numbers were 10.8±2.3 and 26.7±6.6 fmol/mg protein, respectively. The Ka in group I was 10 times greater (P<0.05) than Ka in groups II and III, (I, 3.1±0.9×1010 lm−1; II, 3.4±0.3×109 lm−1; III, 3.3±1.1×109 lm−1). Plasma hormone concentrations typically found during pregnancy were noted. The function of these LH/hCG binding sites remains unknown; however, changes in receptor capacity during the estrous cycle and pregnancy support a role for modulation of the receptor by hormonal factors.  相似文献   

9.
A series of pullulan fractions with molecular weights in the range 5 × 103 to 8 × 105 were prepared. The weight-average molecular weight (Mw) of all the samples was determined by sedimentation equilibrium. The hydrodynamic properties of pullulan in aqueous solution were investigated by viscometry and ultracentrifugation. The experimental results indicate that pullulan molecules in water are fairly stable and behave as expanded random coils when Mw is above 2 × 104. The molecular weight distributions of the fractions were measured by gel filtration. The ratio Mw/Mn was close to 1·1, except for a sample with the highest Mw.It is concluded that the pullulan fractions prepared by the present work are well characterized and have a narrow molecular weight distribution. They may be useful as standard samples for studies of water-soluble polymers.  相似文献   

10.
Low molecular weight heparin of low-anticoagulant activity and high molecular weight heparin of correspondingly high activity were prepared by chromatography on protamine-Sepharose; preparations subjected to limited N-desulfation (5–10% free amino groups) by solvolysis were labeled with 5-dimethylaminonaphthalene-1-sulfonyl chloride (dansyl chloride) or rhodamine B isothiocyanate (RITC). The fluorescent heparins retained approximately 50% of the original anticoagulant activities. Dansyl-heparin on binding to antithrombin III (ATIII) exhibited a 2.5-fold enhancement of dansyl fluorescence intensity. This effect could be prevented by excess unlabeled heparin. A 7900 molecular weight dansyl-heparin preparation bound to ATIII with a stoichiometry of close to 2:1 and with an apparent association constant for binding (Ka) of 4.9 × 105, m?1, whereas a 21,600 molecular weight fraction bound at 0.7:1 with the protein and with an apparent Ka = 7.9 × 105, m?1. When ATIII reacted with a mixture of low molecular weight dansyl-heparin and low molecular weight RITC-heparin, there was enhancement of RITC fluorescence emission when excited at the dansyl excitation maximum; this effect was not observed when either of the labeled heparin species was prepared from high molecular weight material. The results are consistent with the proposal that a single molecule of high molecular weight, high-activity heparin occupies two sites when it binds to ATIII, whereas low molecular weight, low-activity heparin binds to the two sites separately.  相似文献   

11.
The finding that molt-inhibiting hormone (MIH) regulates vitellogenesis in the hepatopancreas of mature Callinectes sapidus females, raised the need for the characterization of its mode of action. Using classical radioligand binding assays, we located specific, saturable, and non-cooperative binding sites for MIH in the Y-organs of juveniles (J-YO) and in the hepatopancreas of vitellogenic adult females. MIH binding to the hepatopancreas membranes had an affinity 77 times lower than that of juvenile YO membranes (KD values: 3.22 × 10-8 and 4.19 × 10-10 M/mg protein, respectively). The number of maximum binding sites (BMAX) was approximately two times higher in the hepatopancreas than in the YO (BMAX values: 9.24 × 10-9 and 4.8 × 10-9 M/mg protein, respectively). Furthermore, MIH binding site number in the hepatopancreas was dependent on ovarian stage and was twice as high at stage 3 than at stages 2 and 1. SDS-PAGE separation of [125I] MIH or [125I] crustacean hyperglycemic hormone (CHH) crosslinked to the specific binding sites in the membranes of the J-YO and hepatopancreas suggests a molecular weight of ~51 kDa for a MIH receptor in both tissues and a molecular weight of ~61 kDa for a CHH receptor in the hepatopancreas. The use of an in vitro incubation of hepatopancreas fragments suggests that MIH probably utilizes cAMP as a second messenger in this tissue, as cAMP levels increased in response to MIH. Additionally, 8-Bromo-cAMP mimicked the effects of MIH on vitellogenin (VtG) mRNA and heterogeneous nuclear (hn) VtG RNA levels. The results imply that the functions of MIH in the regulation of molt and vitellogenesis are mediated through tissue specific receptors with different kinetics and signal transduction. MIH ability to regulate vitellogenesis is associated with the appearance of MIH specific membrane binding sites in the hepatopancreas upon pubertal/final molt.  相似文献   

12.
The physico-chemical properties of the DNA released from bacteriophage G (active on Bacillus megatherium) are described. Phage G, an unusually large bacteriophage, has a nucleic acid content of 4 to 6 × 108 daltons.Sedimentation velocity analysis at low angular speed and examination by electron microscopy, indicate that a single DNA molecule, sedimenting with s20, w0 = 125 ± 1.5 S and at least 200 ± 20 μm long, is released upon thermal or osmotic shock. Melting temperature data and chromatographic analysis indicate a mean base composition of 70% A + T. CsCl and Cs2SO4 buoyant density data, circular dichroism spectra and sensitivity to specific nucleases indicate that phage G DNA is similar to the DNAs from T-even phages and is more glucosylated than phage T6 DNA. Direct glucose determination indicates a 185% molar ratio of glucose to cytosine. Linear density extrapolated from literature data and contour length measurement yield a lower limit for the molecular weight of phage G DNA of 4.9 × 108. Comparison of this value with the s20,w0 measured with the analytical ultracentrifuge seems to confirm the validity of the empirical relationship proposed by Freifelder (1970), between s20, w0 and molecular weight, over a larger range than that previously known. A possible systematic error in defect in length determination, however, prevents a discrimination between this and other empirical formulae proposed by various authors, which predict a molecular weight that is 20 to 25% higher.  相似文献   

13.
Crystals of alkaline phosphatase (EC 3.1.3.1; Mr 94,000) grown at pH 9.5 from 2.25 m-(NH4)2SO4 with 5 × 10?5 m-Zn and 10?2 m-Mg present were analyzed by X-ray diffraction at pH 7.5 in 2.66 m-(NH4)2SO4 with 10?2 m-Zn and 10?2 m-Mg present. The crystals are orthorhombic with a = 195.5 A?, b = 168.3 A?and c = 76.33 A?, and the space group is I222. X-ray phases were determined by the multiple isomorphous replacement and anomalous dispersion method using K2PtCl4, KAu(CN)2 and K2OsO4 derivatives. The electron density maps and analysis of metal binding sites reveal one molecule per asymmetric unit with an internal, non-crystallographic, 2-fold rotation axis relating the subunits. Each subunit contains a major αβ domain with a seven-stranded β-sheet flanked by helices. The sheets are roughly coplanar but the general direction of the strands in each is at 20 ° to the rotation axis and thus 40 ° from each other. The helical content of the αβ domain is approximately 27% of the 459 residues in the monomer and the β content is approximately 7%. The chains in a smaller domain are more convoluted and less easily characterized than in the αβ domain. In both there is extensive monomer-monomer contact.Removal of the zinc and magnesium from the parent crystal produces a stable apoenzyme crystal and addition of cobalt at 10?2 m or cadmium at 10?2 or 5 × 10?2 m reveals seven metal binding sites per dimer. The active centers are 32 Å apart and each is shown by anomalous dispersion data to contain two metal binding sites, A and B. The cadmium derivative refinement determined the A-B separation to be 4.9 Å. Comparison of the parent and apo structures by means of difference maps reveals the double metal site with Zn at A and probably Mg at B. A prominent, partially resolved peak centered 7 Å away is interpreted as a stabilization of the backbone in this position by the metal ion co-ordination to a side-chain. Several negative peaks within 10 Å of the metals indicate local differences between apo and native structures but no significant differences are seen in the other parts of the molecule. At 5 × 10?2 m-Cd two metal sites (D and D′) are found 25.5 Å from the active center, on the surface of the minor domain. They are related to each other by the molecular 2-fold axis with a D-D′ distance of 25 Å. The seventh Cd site, E, is 20 Å from the active center, on the major domain, near a crystalline contact region, and devoid of any molecular symmetry mate.The apparent dissociation constants for cadmium at the A, B and D sites (and A′, B′, D′) are 3 × 10?3 m, 1.5 × 10?1 m and 1.3 × 10?2 m, respectively. Thus in these conditions cadmium is seen to distribute between A and B sites when the combined stoichiometry is two metal ions per dimer.  相似文献   

14.
The effects of cadmium ions or cadmium-metallothionein on the activities of acyl-CoA:1acyl-sn-glycerol 3-phosphoric acid or 1-acyl-sn-glycero 3-phosphocholine acyltransferase of rat liver microsomes have been studied, in vitro. Cadmium ions were found to cause a noncompetitive type inhibition of these two acyltransferases. The Ki values were calculated, and found to be smallest (1.7 × 10?5m) for palmitoyl-CoA and greatest (1.0 × 10?4m) for linoleoyl-CoA, among the several fatty acyl-CoA's tested on the 1-acyl-sn-glycerol 3-phosphoric acid acyltransferases. With the 1-acyl-sn-glycero 3-phosphocholine acyltransferase, the Ki values were found to be smallest for the plamitoyl-CoA acyltransferase (3.8 × 10?5m) and largest for thearachidonoyl-CoA acyltransferase (1.1 × 10?4m). In contrast, mouse liver cadmium-metallothionein, including 4 mol of cadmium and 2 mol of zinc in one molecule of metallothionein, was not found to be inhibitory or rather stimulative on the above two acyltransferases at the same concentration of cadmium tested in the cadmium ion inhibitor experiments. The above results demonstrate that there is a strong and irreversible inhibition by cadmium ions on acyl-CoA acyltransferases, but that when cadmium acts on the enzyme in the form of a cadmium-metallothionein complex, the inhibition effect does not occur. These findings may reflect differing degrees of toxicity of these two types of cadmium compounds in mammalian tissues.  相似文献   

15.
We optimized the conditions for isolation of extracellular catalases from Penicillium piceum F-648 and P. piceum A3 by means of volume chromatography with cadmium hydroxide gel. Our study showed that 55–57 mg wet gel are sufficient for the maximum sorption of catalase from 1 ml of culture fluid. This gel was formed in 1 ml 70 mM Cd(NO3)2 after addition of NaOH (Cd(NO3)2/NaOH molar ratio 1: 2.2). The eluting solution contained 50 mM NaH2PO4(pH 7.0), 5.0 mM dithiothreitol, and 0.3% sodium cholate and was potent in desorbing catalase from the gel. Subsequent ultrafiltration of the eluate on the membrane with a retention limit of 50 kDa allowed us to concentrate and purify the sample from low-molecular-weight protein impurities. NH4Cl (1.0 M) containing 0.3% sodium cholate was used to wash the sample from low-molecular-weight aromatic metabolites. Purified catalases included 33–34% antiparallel β-structures and 9%-spirals. Under optimal conditions in the medium of 10 mM phosphate buffered saline (pH 7.0) at 30°C, catalases from P. piceum F-648 were characterized by the following parameters: K M, 158.8 mM; catalytic constant, 2.83 × 106 s?1; enzyme inactivation rate constant in H2O2 decomposition, 3.5 × 10?2 s?1; and constant of the interaction between catalase complex I and second molecule of H2O2, 1.8 × 107M?1 s?1.  相似文献   

16.
In the preceding paper are described the isolation and physical characterization of seven narrowly disperse fractions of calf thymus DNA in the molecular weight range 0.3 to 1.3 × 106 daltons. Herein, we have determined by light scattering the molecular weights and root mean square radii of these fractions in a solvent comprising 0.2 M NaCl, 2 mM EDTA, 2m MNa-PO4, pH 7. Measurements were made in a modified Wippler—Scheibling photometer to a 20° lower limit of scattering angle on solutions rendered virtually dust-free by procedures described. The optical aniso tropics of the DNA fractions were measured permitting the experimental molecular weights and root mean square radii to be corrected to their true values. From these values, with appropriate polydispersity corrections, we calculate a Kratky—Porod persistence length, a, of 54.0 ± 5.6 nm which is invariant over the molecular weight range examined. From the sedimentation coefficients (preceding paper) and the theory of Yamakawa and Fujii, we calculate a to be 66 nm, a value found to apply equally well to several DNA samples of various origins whose sedimentation rates are known in the molecular weight range from about 4 × 104 to 108 daltons. Similarly, from the intrinsic viscosities and the theory of Yamakawa and Fujii, we calculate a to be 59 nm, which again adequately applies to a number of DNA samples whose viscosities have been measured by other workers in the molecular weight range 3 × 105 to 108 daltons. The Flory—Mandelkern parameter, β, was found to vary with molecular weight in the manner predicted by the theory of Yamakawa and Fujii. The average value of a from the three sets of measurements is 60 ± 6 nm, which we believe applies to double-stranded DNA molecules, independent of chain length, over the whole range of molecular weights for which reliable data exist.  相似文献   

17.
The incorporation of [3H]UTP into RNA by isolated polytene salivary gland nuclei of Chironomus thummi was investigated under different incubation conditions; the labeled RNA fractions were characterized by electrophoresis. The results suggested that at two characteristic ionic conditions most of the RNA synthesized was the product of RNA polymerase I or RNA polymerase II as distinguished by their differential sensitivities to α-amanitin. Electrophoretical analysis of the RNA synthesized under conditions favouring polymerase I showed that this RNA population consisted mainly of four distinct molecular weight fractions within a range between 2.8 × 104 and 2.5 × 106. Under conditions favouring polymerase II two fractions were detected: one with a broad molecular weight distribution around 0.4 × 106 containing considerable amounts of poly(A)-bearing RNA molecules, and a second with a peak at a molecular weight of 2.8 × 104.  相似文献   

18.
The coupling factor protein isolated previously in pure form with a molecular weight of 11–12 × 103 (K.-S. You and Y. Hatefi, 1976, Biochim. Biophys. Acta423, 398–412) has been shown to restore ATP-induced NAD reduction by succinate, transhydrogenation from NADH to NADP, and ATP-33Pi exchange to submitochondrial particles rendered deficient by extraction with 1 m NH4OH. The factor also stimulated the oxidative phosphorylation activity of the extracted particles 2.5- to >3-fold. The stimulatory effect of the factor was inhibited by mercurials, Cd2+, phenylarsine oxide, and diamide, indicating that it contains an essential dithiol. Dithiothreitol and dihydrolipoate did not replace the protein factor in stimulating the deficient particles. The purified dithiol-containing protein was precipitated and inhibited by antibody raised against coupling factor B. Since this antibody also inhibits coupling factor F2, it is concluded that the active principle of coupling factors B and F2 is the purified dithiol-containing protein of molecular weight 11–12 × 103 referred to above.  相似文献   

19.
20.
The extraction temperature had a significant impact on the concentration of polysaccharides derived from solid-liquid extraction of Spirulina. The polysaccharide concentration was significantly higher when the extraction was performed at 90°C than when it was performed at 80, 70, and 50°C. This result is related to the diffusion coefficients of the polysaccharides, which increased from 1.07 × 10?12 at 50°C to 3.02 × 10?12 m2/sec at 90°C. Using the Arrhenius equation, the pre-exponential factor (D 0 ) and the activation energy (E a ) for Spirulina polysaccharide extraction were calculated as 7.958 × 10?9 m2/sec and 24.0 kJ/mol, respectively. Among the methods used for the separation of Spirulina polysaccharides, cetyltrimethylammonium bromide (CTAB, method I) and organic solvent (ethanol, in methods II and III) provided similar yields of polysaccharides. However, the separation of polysaccharides using an ultrafiltration (UF) process (method III) and ethanol precipitation was superior to separation via CTAB or vacuum rotary evaporation (method II). The use of a membrane with a molecular weight cut-off (MWCO) of 30 kDa and an area of 0.01 m2 at a feed pressure of 103 kPa with a mean permeate flux of 39.3 L/m2/h and a retention rate of 95% was optimal for the UF process. The addition of two volumes (v/v) of ethanol, which gave a total polysaccharide content of approximately 4% dry weight, was found to be most suitable for polysaccharide precipitation. The results of a Sepharose 6B column separation showed that the molecular weights of the polysaccharides in fractions I and II were 212 and 12.6 kDa, 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