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.
High activity of taurine:alpha-ketoglutarate aminotransferase was found exclusively in cell-free extracts of Achromobacter superficialis and A. polymorph. The former was chosen for characterization of the enzymatic reaction. The enzyme activity was enhanced by addition of beta-alanine to the growth medium. The product from alpha-ketoglutarate was identified as l-glutamate. Another product has been isolated, purified, and identified as sulfoacetaldehyde (2-oxoethanesulfonate), a deamination product from taurine, by comparison between the 2,4-dinitrophenylhydrazones of the synthetic and enzymatic products on the basis of studies by paper chromatography, by visible, infrared, and nuclear magnetic resonance spectrophotometries, and by elemental analysis. This enzymatic transamination was found to proceed stoichiometrically and reversibly as follows: NH(2).CH(2).CH(2).SO(3)H + HOOC.CH(2).CH(2).CO.COOH right harpoon over left harpoon OHC.CH(2).SO(3)H + HOOC.CH(2).CH(2).CH(NH(2)).COOH.  相似文献   

2.
The reliability of dynamic measurement methods of k(l)a in fermentors using a step oxygen concentration change in the feed gas was tested. The tests were performed both for the original variant using the nitrogen right harpoon over left harpoon air exchange and the newly presented variant using the oxygen-enriched air (27 vol % O(2)) --> air exchange. The testing consisted in comparing k(l)a values determined from these methods with values determined from the steady-state Na(2)SO(3) feeding method and the dynamic pressure method, the reliability of which was proven earlier. The measurements were done in water (coalescent batch) and in 0.5M Na(2)SO(4) solution with and without the addition of 1 wt % carboxymethylcellulose (noncoalescent batches). It was found that in noncoalescent liquids the methods tested give extremely low k(l)a values (as low as 15% of the correct value). The methods are defective in principle irrespective of the gases used for exchange.  相似文献   

3.
p-Fluorophenylalanine (FPA) causes a 100-fold increase in alkaline phosphatase in Escherichia coli B, strain PR1 at 30 C in minimal medium that contains excess inorganic phosphate (1.92 x 10(-3)m). Little increase in alkaline phosphatase synthesis occurs under these conditions at 22 C. [This strain is known to have a mutation in a regulator gene (R(2)) that, in the absence of FPA, permits derepression of alkaline phosphatase synthesis at 37 C, but not at 30 C or below.] In contrast, E. coli B3 (the strain from which E. coli B strain PR1 was derived) is not derepressed at 30 C by FPA. (14)C-FPA is incorporated into bacterial proteins. Temperature-shift experiments (30 Cright harpoon over left harpoon22 C) in the presence of FPA are consistent with the following mechanism. FPA is incorporated into the genetically altered R(2) protein at 30 and 22 C. This further alteration due to the incorporation of analogue makes the R(2) protein inactive at 30 C, but active at 22 C.  相似文献   

4.
Single-walled aluminum nitride nanotubes (AlNNTs) are introduced as an electronic sensor for detection of sulfur dioxide (SO(2)) molecules based on density functional theory calculations. The proposed sensor benefits from several advantages including high sensitivity: HOMO-LUMO energy gap of the AlNNT is appreciably sensitive toward the presence of SO(2) so that it decreases from 4.11?eV in the pristine tube to 1.01?eV in the SO(2)-adsorbed form, pristine application: this nanotube can detect the SO(2) molecule in its pristine type without manipulating its structure through doping, chemical functionalization, making defect, etc., short recovery time: the adsorption energy of SO(2) molecule is not so large to hinder the recovery of AlNNTs and therefore the sensor will possess short recovery times, and good selectivity: the tube can selectively detect the SO(2) molecule in the presence of several molecules such as H(2)O, CO, NH(3), HCOH, CO(2), N(2), and H(2).  相似文献   

5.
The anion influx was measured in order to study the interaction among organic phosphates, magnesium, haemoglobin and the N-terminal of the cytoplasmic domain of band 3 protein in human, chicken and trout erythrocytes. The rate constant for SO(4)(2-) influx in human and trout erythrocytes increased significantly when it was measured with an increased concentration of intracellular Mg(2+). The SO(4)(2-) influx was also measured in human erythrocyte ghosts in the presence and absence of Mg(2+). The smaller activation provoked by Mg(2+) in ghosts could be caused by the presence of a small quantity of haemoglobin which remained inside. The SO(4)(2-) uptake in chicken erythrocytes in the presence and in absence of Mg(2+) was characterized by very similar rate constants. The results suggest that the small increase in intracellular Mg(2+) in the erythrocytes involves an increase in the formation of Mg(2+)-ATP and Mg(2+)-2,3 BPG complexes reducing the affinity of the organic phosphates for Hb. This new situation may influence the functions of the anion transporter with consequent variations of SO(4)(2-) influx throughout the erythrocyte membrane in human and in trout erythrocytes, whereas in chicken RBCs this function cannot occur and, in fact, no increase in sulphate influx was noticeable. The measurement of Hb/O(2) affinity by the use of alternating fixed and variable concentrations of organic phosphates and Mg(2+), confirms the interactions between these elements and their effect on the mechanism of the affinity. When we measured the sulphate influx in the presence of DIDS we found some differences in the three types of cells.  相似文献   

6.
The simian immunodeficiency virus fusion peptide constitutes a 12-residue N-terminal segment of the gp32 protein that is involved in the fusion between the viral and cellular membranes, facilitating the penetration of the virus in the host cell. Simian immunodeficiency virus fusion peptide is a hydrophobic peptide that in Me(2)SO forms aggregates that contain beta-sheet pleated structures. When added to aqueous media the peptide forms large colloidal aggregates. In the presence of lipidic membranes, however, the peptide interacts with the membranes and causes small changes of the membrane electrostatic potential as shown by fluorescein phosphatidylethanolamine fluorescence. Thioflavin T fluorescence and Fourier transformed infrared spectroscopy measurements reveal that the interaction of the peptide with the membrane bilayer results in complete disassembly of the aggregates originating from an Me(2)SO stock solution. Above a lipid/peptide ratio of about 5, the membrane disaggregation and water precipitation processes become dependent on the absolute peptide concentration rather than on the lipid/peptide ratio. A schematic mechanism is proposed, which sheds light on how peptide-peptide interactions can be favored with respect to peptide-lipid interactions at various lipid/peptide ratios. These studies are augmented by the use of the fluorescent dye 1-(3-sulfonatopropyl)-4-[beta[2-(di-n-octylamino)-6-naphthyl]vinyl ] pyridinium betaine that shows the interaction of the peptide with the membranes has a clear effect on the magnitude of the so-called dipole potential that arises from dipolar groups located on the lipid molecules and oriented water molecules at the membrane-water interface. It is shown that the variation of the membrane dipole potential affects the extent of the membrane fusion caused by the peptide and implicates the dipolar properties of membranes in their fusion.  相似文献   

7.
8.
For many proteins, compact conformations are known to accumulate in advance of the rate-limiting step in folding. To understand the nature and significance of these early conformational events, we employed ultrarapid mixing methods to fully characterize the kinetics of folding of the 57-residue B1 domain of protein G. Continuous-flow fluorescence measurements exhibit a major exponential phase on the submillisecond time scale (600-700 micros), which is followed by a slower phase with a denaturant-dependent time constant (2-30 ms) observable by conventional stopped-flow measurements. The combined kinetic traces quantitatively account for the total change in Trp 43 fluorescence upon folding, including the previously unresolved 'burst phase' signal. The denaturant dependence of the two rate constants and their relative amplitudes are fully consistent with a three-state mechanism, U right harpoon over left harpoon I right harpoon over left harpoon N, where I is a productive intermediate with native-like fluorescence properties. The relatively slow rate and exponential time course of the initial folding phase indicates that a substantial free energy barrier is encountered during chain condensation, resulting in a partially organized ensemble of states distinct from the initial unfolded conformations.  相似文献   

9.
Polyethylene glycol (PEG) is extensively employed for protein purification by fractional precipitation. Efficiency of precipitation is highest when the solution pH is near the isoelectric point of the target protein. At pH values far from the isoelectric point of the target protein, proteins develop a net positive or negative charge and are not more resistant to precipitation. We have found that divalent cations (Ba2+, Sr2+, and Ca2+) or divalent anions (SO4(2-)) significantly change the pattern of PEG precipitation when the ion is chosen so as to counteract the expected net charge on the target protein. At moderate (5-50 mM) concentrations of Ba2+, negatively charged proteins can be precipitated from solution at pH values as high as 10 with efficiency unchanged from precipitation at pH values near their isoelectric point values. The mechanism of PEG precipitation of protein at these high pH values appears to be unchanged from the mechanism operative at the protein isoelectric point. Precipitation is rapid and the capacity for protein precipitation is high. There is no detectable coprecipitation of small molecules (AMP, ATP, and NADH) or soluble proteins (carbonic anhydrase) induced when large quantities of protein are precipitated by this method. The purification of bovine carbonic anhydrase from erythrocyte lysate is more efficient at pH 10 in the presence of Ba2+ than is conventional PEG precipitation carried out at the isoelectric point of carbonic anhydrase. Application of these observations should broaden the utility of protein purification by fractional precipitation with PEG.  相似文献   

10.
The kinetics of protein-fluorescence change when rabbit skeletal myosin subfragment 1 is mixed with ATP or adenosine 5'-(3-thiotriphosphate) in the presence of Mg(2+) are incompatible with a simple bimolecular association process. A substrate-induced conformation change with DeltaG(0)<-24kJ.mol(-1) (i.e. DeltaG(0) could be more negative) at pH8 and 21 degrees C is proposed as the additional step in the binding of ATP. The postulated binding mechanism is M+ATPright harpoon over left harpoonM.ATPright harpoon over left harpoonM*.ATP, where the association constant for the first step, K(1), is 4.5x10(3)m(-1) at I 0.14m and the rate of isomerization is 400s(-1). In the presence of Mg(2+), ADP binds in a similar fashion to ATP, the rate of the conformation change also being 400s(-1), but with DeltaG(0) for that process being -14kJ.mol(-1). The effect of increasing ionic strength is to decrease K(1), the kinetics of the conformation change being essentially unaltered. Alternative schemes involving a two-step binding process for ATP to subfragment 1 are possible. These are not excluded by the experimental results, although they are perhaps less likely because they imply uncharacteristically slow bimolecular association rate constants.  相似文献   

11.
The urea-induced equilibrium unfolding of the alpha-subunit of tryptophan synthase (alphaTS) from Escherichia coli can be described by a four-state model, N right harpoon over left harpoon I1 right harpoon over left harpoon I2 right harpoon over left harpoon U, involving two highly populated intermediates, I1 and I2 [Gualfetti, P. J., Bilsel, O., and Matthews, C. R. (1999) Protein Sci. 8, 1623-1635]. To extend the physical characterization of these stable forms, the apparent radius was measured by several techniques. Size-exclusion chromatography (SEC), analytical ultracentrifugation (UC), and dynamic light scattering (DLS) experiments yield an apparent Stokes radius, R(s), of approximately 24 A for the native state of alphaTS. The small-angle X-ray scattering (SAXS) experiment yields a radius of gyration, R(g), of 19.1 A, consistent with the value predicted from the X-ray structure and the Stokes radius. As the equilibrium is shifted to favor I1 at approximately 3.2 M and I2 at 5.0 M urea, SEC and UC show that R(s) increases from approximately 38 to approximately 52 A. Measurements of the radius by DLS and SAXS between 2 and 4.5 M urea were complicated by the self-association of the I1 species at the relatively high concentrations required by those techniques. Above 6 M urea, SEC and UC reveal that R(s) increases linearly with increasing urea concentration to approximately 54 A at 8 M urea. The measurements of R(s) by DLS and R(g) by SAXS are sufficiently imprecise that both values appear to be identical for the I2 and U states and, considering the errors, are in good agreement with the results from SEC and UC. Thermodynamic parameters extracted from the SEC data for the N right harpoon over left harpoon I1 and I1 right harpoon over left harpoon I2 transitions agree with those from the optical data, showing that this technique accurately monitors a part of the equilibrium model. The lack of sensitivity to the I2 right harpoon over left harpoon U transition, beyond a simple swelling of both species with increasing urea concentration, implies that the Stokes radii for the I2 and U states are not distinguishable. Surprisingly, the hydrophobic core known to stabilize I2 at 5.0 M urea [Saab-Rincón, G., Gualfetti, P. J., and Matthews, C. R. (1996) Biochemistry 35, 1988-1994] develops without a significant contraction of the polypeptide, i.e., beyond that experienced by the unfolded form at decreasing urea concentrations. Kratky plots of the SAXS data, however, reveal that I2, similar to N and I1, has a globular structure while U has a more random coil-like form. By contrast, the formation of substantial secondary structure and the burial of aromatic side chains in I1 and, eventually, N are accompanied by substantial decreases in their Stokes radii and, presumably, the size of their respective conformational ensembles.  相似文献   

12.
Band 3 protein was reconstituted with lipid vesicles consisting of 94:6 (molar ratio) egg phosphatidylcholine-bovine heart phosphatidylserine in a 2500:1 phospholipid:protein molar ratio by means of a Triton X-100/beads method. The SO2-4 permeability of the resulting vesicles was measured using an influx assay procedure in which the vesicles were sampled and subsequently eluted over Sephadex columns at appropriate time intervals. The accuracy of the assay was greatly increased by using an internal standard in order to correct for vesicle recovery. In agreement with previous work, it could be demonstrated that incorporation of band 3 in the vesicles caused an increase in SO2-4 permeability, which could be (partially) inhibited by high concentrations of DIDS or a competitive anion such as thiocyanate. However, the magnitude of the increased SO2-4 permeability was highly variable, even when vesicles were reconstituted using band 3 isolated from one batch of ghosts. In addition, the SO2-4 influx curves showed complex kinetics. These results are related to the existence of vesicle heterogeneity with respect to protein content and vesicle size as revealed by stractan density gradient centrifugation and freeze-fracture electron microscopy. Band 3 incorporation also increased the L-glucose permeability of the vesicles which could also be inhibited by DIDS. Glycophorin, which has no known transport function, reconstituted with lipid vesicles consisting of 94:6 (molar ratio) egg phosphatidylcholine-bovine heart phosphatidylserine in a 400:1 phospholipid:protein molar ration increased the bilayer permeability towards SO2-4 as well as towards L-glucose. Surprisingly, the SO2-4 permeability in the vesicles could also be inhibited by DIDS and thiocyanate. It is concluded that the use of DIDS and a competitive anion, thiocyanate, in order to prove that band 3 is functionally reconstituted, is highly questionable. The increased SO2-4 and L-glucose permeability of band 3-lipid as well as glycophorin-lipid vesicles and the inhibitory action of DIDS are discussed in the light of the presence of defects at the lipid/protein interface and protein aggregation, which may induce the formation of pores. Since the band 3-lipid vesicles are more permeable for SO2-4 than for L-glucose, in contrast to the glycophorin-containing vesicles, it is suggested that some anion specificity of the increased bilayer permeability in the band 3-lipid vesicles is still preserved.  相似文献   

13.
The syntheses are reported for the first time of alpha-L-IdopA2SO(3)-(1-->3)-beta-D-GalpNAc4SO(3)-(1-->4)-alpha-L-IdopA2SO(3)-(1-->OMe), its disulfated analogue alpha-L-IdopA2SO(3)-(1-->3)-beta-D-GalpNAc-(1-->4)-alpha-L-IdopA2SO(3)-(1-->OMe), and of beta-D-GalpNAc4SO(3)-(1-->4)-alpha-L-IdopA2SO(3)-(1-->3)-beta-D-GalpNAc4SO(3)-(1-->OMe), which represent structural fragments of dermatan sulfate, unavailable directly by chemical or enzymatic degradation of the glycosaminoglycan polymer. These molecules were readily obtained from a pair of key disaccharide intermediates, in which the relative difference of stability of the D-GalNAc 4-hydroxy protecting groups (acetate or pivalate) toward saponification conditions allowed access to various sulfoforms from a common precursor. For the preparation of these blocks, the 4-O-pivaloyl-D-galacto moiety was readily obtained through a one-pot stereospecific intramolecular nucleophilic displacement on an easily available 3-O-pivaloyl-D-gluco precursor, and the L-IdoA moiety through selective radical oxidation at C-6 of a L-ido 4,6-diol derivative with oxoammonium salts.  相似文献   

14.
Sodium houttuyfonate analogs (SHAs), CH(3)-(CH(2))(n)-CO-CH(2)-CH(OH)SO(3)Na, (n=6-14) were synthesized and their molecular interactions with renin and angiotensin I converting enzyme (ACE) studied using fluorescence quenching techniques. Unlike renin, inhibition of ACE activity was not directly proportional to the aliphatic chain length of SHAs. Ability of SHAs to inhibit enzyme activities and quench protein fluorescence was greater with renin than with ACE. The presence of an ACE substrate (angiotensin I) did not reduce quenching ability of SHAs, suggesting that enzyme-inhibitor interactions did not involve the active site or the substrate was displaced by inhibitor molecules. The results showed that renin is a more sensitive target than ACE for the potential antihypertensive ability of SHAs.  相似文献   

15.
Uptake of SO(4) (2-) into brush-border membrane vesicles isolated from rat kindey cortex by a Ca(2+)-precipitation method was investigated by using a rapid-filtration technique. Uptake of SO(4) (2-) by the vesicles was osmotically sensitive and represented transport into an intra-vesicular space. Transport of SO(4) (2-) by brush-border membranes was stimulated in the presence of Na(+), compared with the presence of K(+) or other univalent cations. A typical ;overshoot' phenomenon was observed in the presence of an NaCl gradient (100mm-Na(+) outside/zero mm-Na(+) inside). Radioactive-SO(4) (2-) exchange was faster in the presence of Na(+) than in the presence of K(+). Addition of gramicidin-D, an ionophore for univalent cations, decreased the Na(+)-gradient-driven SO(4) (2-) uptake. SO(4) (2-) uptake was only saturable in the presence of Na(+). Counter-transport of Na(+)-dependent SO(4) (2-) transport was shown with MoO(4) (2-) and S(2)O(3) (2-), but not with PO(4) (2-). Changing the electrical potential difference across the vesicle membrane by establishing different diffusion potentials (anion replacement; K(+) gradient+/-valinomycin) was not able to alter Na(+)-dependent SO(4) (2-) uptake. The experiments indicate the presence of an electroneutral Na(+)/SO(4) (2-)-co-transport system in brush-border membrane vesicles isolated from rat kidney cortex.  相似文献   

16.
Anion dependence of (Ca2+ + K+)-stimulated Mg2+-dependent transport ATPase and its phosphorylated intermediate have been characterized in both "intact" and "broken" vesicles from endoplasmic reticulum of rat pancreatic acinar cells using adenosine 5'-[gamma-32P] triphosphate ([gamma-32P]ATP). In intact vesicles (Ca2+ + K+)-Mg2+-ATPase activity was higher in the presence of Cl- or Br- as compared to NO3-, SCN-, cyclamate-, SO4(2-) or SO3(2-). Incorporation of 32P from [gamma-32P]ATP into the 100-kDa intermediate of this Ca2+ATPase was also higher in the presence of Cl-, Br-, NO3- or SCN- as compared to cyclamate-, SO4(2-) or SO3(2-). When the membrane permeability barrier to anions was abolished by breaking vesicle membrane with the detergent Triton X-100 (0.015%) (Ca2+ + K+)-Mg2+ATPase activity in the presence of weakly permeant anions, such as SO4(2-) and cyclamate-, increased to the level obtained with Cl-. However, 32P incorporation into 100-kDa protein was still higher in the presence of Cl- as compared to cyclamate-, indicating a direct effect of Cl- on the Ca2+ATPase molecule. The anion transport blocker 4,4-diisothiocyanostilbene-2,2-disulfonate (DIDS) inhibited (Ca2+ + K+)-Mg2+ATPase activity to about 10% of the Cl- stimulation level, irrespective of the sort of anions present in both intact and broken vesicles. This indicates a direct effect of DIDS on (Ca2+ + K+)-Mg2+ATPase. K+ ionophore valinomycin influenced (Ca2+ + K+)-Mg2+ATPase activity according to the actual K+ gradient: Ko+ greater than Ki+ caused inhibition, Ko+ less than Ki+ caused stimulation. From these results we conclude that Ca2+ transport into endoplasmic reticulum is coupled to ion movements which must occur to maintain electroneutrality.  相似文献   

17.
"Band 3," an integral membrane protein of red blood cells, plays a relevant role in anionic transport. The C- and N-terminal portions of band 3 are cytoplasmatics, and the last is the link site for different glycolitic enzymes, such as glyceraldehyde-3-phosphate dehydrogenase, aldolase, phosphofructokinase, and hemoglobin. All or some of these interactions on the CDB3 protein could allow a subtle modulation of anion flux. The interaction among HbA, Mg(2+), and membrane proteins has been sufficiently investigated, but not the effect of Mg(2+) on pathological hemoglobin in relation to the influx of the SO(4)(2-). The aim of this study was to evaluate the involvement of hemoglobin S in sulfate transport. This has been measured with native and increased concentrations of Mg(2+), using normal erythrocytes containing HbA, sickle red cells containing HbS, or ghosts obtained from both erythrocytes and normal erythrocytes ghosts with HbS added. The magnitude of the SO(4)(2-) rate constant measured in normal red blood cells increased markedly when measured in the presence of varied Mg(2+) concentrations. The results show that a low increase of intracellular Mg(2+) concentrations exercises a different HbA modulation on band 3 protein and consequently higher anion transport activity. The same experiments carried out in sickle red cells showed that the SO(4)(2-) rate constant measured in the presence of native concentrations of Mg(2+) was normal, compared to normal red cells, and was not affected by any increase of intracellular Mg(2+). Our suppositions with regard to the importance exercised by the hemoglobin and the Mg(2+) on the SO(4)(2-) influx were confirmed by comparison of the data obtained through measuring SO(4)(2-) influx with native and increased concentrations of Mg(2+) in both normal and sickle red cell ghosts. Both revealed the same sensitivity to Mg(2+) due to withdrawal of hemoglobins. The incorporation of HbS in normal as well as in sickle red cell ghosts reduced the Mg(2+) response to sulfate influx in both the reconstituted ghosts. Our research demonstrated that the different effects exercised on the rate constants of SO(4)(2-) influx in normal (HbA) and sickle red cells (HbS) by the increased intracellular Mg(2+) could be ascribed to the physical-chemical influence exercised either on the hemoglobins or on the intracellular contents of erythrocytes.  相似文献   

18.
Stopped-flow spectrophotometry and e.p.r. spectroscopy were used to study the kinetics of reduction by dithionite of the oxidized Fe protein of nitrogenase from Klebsiella pneumoniae (Kp2ox.) in the presence of MgADP at 23 degrees C at pH 7.4. The active reductant, SO2.-, produced by the predissociation of S2O4(2-) in equilibrium 2SO2.-, reacts with Kp2ox. (MgADP)2, with k4 = 3.0 X 10(6) +/- 0.4 X 10(6) M-1 X s-1. The inhibition of this reaction by the Mo-Fe protein (Kp1) has enabled the rate of dissociation of Kp2ox. (MgADP)2 from Kp1+ (the Kp2-binding site on Kp1) to be measured (k-3 = 6.4 +/- 0.8 s-1). Comparison with the steady-state rate of substrate reduction shows that the dissociation (k-3) of the complex Kp2ox. (MgADP)2-Kp1+, which is formed after MgATP-induced electron transfer from Kp2 to Kp1+, is the rate-limiting step in the catalytic cycle for substrate reduction.  相似文献   

19.
1. Sugar-cane leaf pyruvate,P(i) dikinase was prepared free of enzymes that would interfere with studies on the stoicheiometry and mechanism of the reaction it catalyses. The reaction was unequivocally shown to involve the conversion of equimolar amounts of pyruvate, ATP and P(i) into phosphoenolpyruvate, AMP and PP(i). 2. The purified enzyme was stable at pH8.3 only if stored at about 20 degrees in the presence of Mg(2+) and a thiol-reducing reagent, care being taken to prevent the oxidation of the thiol. 3. The apparent Michaelis constants for phosphoenolpyruvate and PP(i) were 0.11mm and 0.04mm respectively and that for AMP was less than 4mum. 4. At pH8.3 the initial velocity of the reaction was about 6 times as fast in the direction towards phosphoenolpyruvate synthesis as in the reverse direction. 5. With the exception of ATP, all the products of the reaction in both directions were inhibitory. 6. The phosphate groups of PP(i) were derived from P(i) and from the terminal phosphate of ATP. 7. Isotope-exchange studies indicated that the reaction proceeds in the following steps:Enzyme+ATP+P(i) right harpoon over left harpoon Enzyme-P+AMP+PP(i)Enzyme-P+pyruvate right harpoon over left harpoon Enzyme+phosphoenolpyruvate  相似文献   

20.
When a 100,000 x g supernatant from rat intestinal mucosa was incubated with 4,4'-bis([3H]methylsulfonyl)-2,2',5,5'-tetrachlorobiphenyl, [(CT3SO2)2TCB] a (CT3SO2)2TCB-protein complex was formed. The (CT3SO2)2TCB-protein complex was isolated and purified using gel filtration and ion-exchange chromatography. The protein portion of this complex was characterized to be liver fatty acid binding protein (L-FABP) by SDS-polyacrylamide gel electrophoresis and immunoblot analysis. No cross reactivity was observed in the immunoblot analysis between the purified protein and anti-heart or anti-intestinal fatty acid binding protein. (CT3SO2)2TCB was extractable from L-FABP and therefore not covalently bound to L-FABP.  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