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.
Cytoglobin (Cygb) was investigated for its capacity to function as a NO dioxygenase (NOD) in vitro and in hepatocytes. Ascorbate and cytochrome b5 were found to support a high NOD activity. Cygb-NOD activity shows respective Km values for ascorbate, cytochrome b5, NO, and O2 of 0.25 mm, 0.3 μm, 40 nm, and ∼20 μm and achieves a kcat of 0.5 s−1. Ascorbate and cytochrome b5 reduce the oxidized Cygb-NOD intermediate with apparent second order rate constants of 1000 m−1 s−1 and 3 × 106 m−1 s−1, respectively. In rat hepatocytes engineered to express human Cygb, Cygb-NOD activity shows a similar kcat of 1.2 s−1, a Km(NO) of 40 nm, and a kcat/Km(NO) (kNOD) value of 3 × 107 m−1 s−1, demonstrating the efficiency of catalysis. NO inhibits the activity at [NO]/[O2] ratios >1:500 and limits catalytic turnover. The activity is competitively inhibited by CO, is slowly inactivated by cyanide, and is distinct from the microsomal NOD activity. Cygb-NOD provides protection to the NO-sensitive aconitase. The results define the NOD function of Cygb and demonstrate roles for ascorbate and cytochrome b5 as reductants.  相似文献   

2.
KIF3AB is an N-terminal processive kinesin-2 family member best known for its role in intraflagellar transport. There has been significant interest in KIF3AB in defining the key principles that underlie the processivity of KIF3AB in comparison with homodimeric processive kinesins. To define the ATPase mechanism and coordination of KIF3A and KIF3B stepping, a presteady-state kinetic analysis was pursued. For these studies, a truncated murine KIF3AB was generated. The results presented show that microtubule association was fast at 5.7 μm−1 s−1, followed by rate-limiting ADP release at 12.8 s−1. ATP binding at 7.5 μm−1 s−1 was followed by an ATP-promoted isomerization at 84 s−1 to form the intermediate poised for ATP hydrolysis, which then occurred at 33 s−1. ATP hydrolysis was required for dissociation of the microtubule·KIF3AB complex, which was observed at 22 s−1. The dissociation step showed an apparent affinity for ATP that was very weak (K½,ATP at 133 μm). Moreover, the linear fit of the initial ATP concentration dependence of the dissociation kinetics revealed an apparent second-order rate constant at 0.09 μm−1 s−1, which is inconsistent with fast ATP binding at 7.5 μm−1 s−1 and a Kd,ATP at 6.1 μm. These results suggest that ATP binding per se cannot account for the apparent weak K½,ATP at 133 μm. The steady-state ATPase Km,ATP, as well as the dissociation kinetics, reveal an unusual property of KIF3AB that is not yet well understood and also suggests that the mechanochemistry of KIF3AB is tuned somewhat differently from homodimeric processive kinesins.  相似文献   

3.
Regulation of sulfate transport in filamentous fungi   总被引:2,自引:0,他引:2       下载免费PDF全文
Inorganic sulfate enters the mycelia of Aspergillus nidulans, Penicillium chrysogenum, and Penicillium notatum by a temperature-, energy-, pH-, ionic strength-, and concentration-dependent transport system (“permease”). Transport is unidirectional. In the presence of excess external sulfate, ATP sulfurylase-negative mutants will accumulate inorganic sulfate intracellularly to a level of about 0.04 m. The intracellular sulfate can be retained against a concentration gradient. Retention is not energy-dependent, nor is there any exchange between intracellular (accumulated) and extracellular sulfate. The sulfate permease is under metabolic control. Sulfur starvation of high methionine-grown mycelia results in about a 1000-fold increase in the specific sulfate transport activity at low external sulfate concentrations. l-Methionine is a metabolic repressor of the sulfate permease, while intracellular sulfate and possibly l-cysteine (or a derivative of l-cysteine) are feedback inhibitors. Sulfate transport follows hyperbolic saturation kinetics with a Michaelis constant (Km) value of 6 × 10−5 to 10−4m and a Vmax (for maximally sulfurstarved mycelia) of about 5 micromoles per gram per minute. Refeeding sulfur-starved mycelia with sulfate or cysteine results in about a 10-fold decrease in the Vmax value with no marked change in the Km. Azide and dinitrophenol also reduce the Vmax.  相似文献   

4.
1. The formation of adenosine 5′-phosphate, guanosine 5′-phosphate and inosine 5′-phosphate from [8-14C]adenine, [8-14C]guanine and [8-14C]hypoxanthine respectively in the presence of 5-phosphoribosyl pyrophosphate and an extract from Ehrlich ascites-tumour cells was assayed by a method involving liquid-scintillation counting of the radioactive nucleotides on diethylaminoethylcellulose paper. The results obtained with guanine were confirmed by a spectrophotometric assay which was also used to assay the conversion of 6-mercaptopurine and 5-phosphoribosyl pyrophosphate into 6-thioinosine 5′-phosphate in the presence of 6-mercaptopurine phosphoribosyltransferase from these cells. 2. At pH 7·8 and 25° the Michaelis constants for adenine, guanine and hypoxanthine were 0·9 μm, 2·9 μm and 11·0 μm in the assay with radioactive purines; the Michaelis constant for guanine in the spectrophotometric assay was 2·6 μm. At pH 7·9 the Michaelis constant for 6-mercaptopurine was 10·9 μm. 3. 25 μm-6-Mercaptopurine did not inhibit adenine phosphoribosyltransferase. 6-Mercaptopurine is a competitive inhibitor of guanine phosphoribosyltransferase (Ki 4·7 μm) and hypoxanthine phosphoribosyltransferase (Ki 8·3 μm). Hypoxanthine is a competitive inhibitor of guanine phosphoribosyltransferase (Ki 3·4 μm). 4. Differences in kinetic parameters and in the distribution of phosphoribosyltransferase activities after electrophoresis in starch gel indicate that different enzymes are involved in the conversion of adenine, guanine and hypoxanthine into their nucleotides. 5. From the low values of Ki for 6-mercaptopurine, and from published evidence that ascites-tumour cells require supplies of purines from the host tissues, it is likely that inhibition of hypoxanthine and guanine phosphoribosyltransferases by free 6-mercaptopurine is involved in the biological activity of this drug.  相似文献   

5.
1. The kinetic properties of the soluble and particulate hexokinases from rat heart have been investigated. 2. For both forms of the enzyme, the Km for glucose was 45μm and the Km for ATP 0·5mm. Glucose 6-phosphate was a non-competitive inhibitor with respect to glucose (Ki 0·16mm for the soluble and 0·33mm for the particulate enzyme) and a mixed inhibitor with respect to ATP (Ki 80μm for the soluble and 40μm for the particulate enzyme). ADP and AMP were competitive inhibitors with respect to ATP (Ki for ADP was 0·68mm for the soluble and 0·60mm for the particulate enzyme; Ki for AMP was 0·37mm for the soluble and 0·16mm for the particulate enzyme). Pi reversed glucose 6-phosphate inhibition with both forms at 10mm but not at 2mm, with glucose 6-phosphate concentrations of 0·3mm or less for the soluble and 1mm or less for the particulate enzyme. 3. The total activity of hexokinase in normal hearts and in hearts from alloxan-diabetic rats was 21·5μmoles of glucose phosphorylated/min./g. dry wt. of ventricle at 25°. The temperature coefficient Q10 between 22° and 38·5° was 1·93; the ratio of the soluble to the particulate enzyme was 3:7. 4. The kinetic data have been used to predict rates of glucose phosphorylation in the perfused heart at saturating concentrations of glucose from measured concentrations of ATP, glucose 6-phosphate, ADP and AMP. These have been compared with the rates of glucose phosphorylation measured with precision in a small-volume recirculation perfusion apparatus, which is described. The correlation between predicted and measured rates was highly significant and their ratio was 1·07. 5. These findings are consistent with the control of glucose phosphorylation in the perfused heart by glucose 6-phosphate concentration, subject to certain assumptions that are discussed in detail.  相似文献   

6.
1. Phosphomevalonate kinase and 5-pyrophosphomevalonate decarboxylase have been purified from the freeze-dried latex serum of the commercial rubber tree Hevea brasiliensis. 2. The phosphomevalonate kinase was acid- and heat-labile and required the presence of a thiol to maintain activity. 3. The 5-pyrophosphomevalonate decarboxylase was relatively acid-stable and more heat-stable than the phosphokinase. 4. Maximum activity of the phosphokinase was achieved at pH 7.2 with 0.2mm-5-phosphomevalonate (Km 0.042mm), 2.0mm-ATP (Km 0.19mm) and 8mm-Mg2+ at 40°C. The apparent activation energy was 14.8kcal/mol. 5. Maximum activity of 5-pyrophosphomevalonate decarboxylase was achieved at pH5.5–6.5 with 0.1mm-5-pyrophosphomevalonate (Km 0.004mm), 1.5mm-ATP (Km 0.12mm) and 2mm-Mg2+. The apparent activation energy was 13.7kcal/mol. The enzyme was somewhat sensitive to inhibition by its products, isopentenyl pyrophosphate and ADP.  相似文献   

7.
Terminal Oxidases of Chlorella pyrenoidosa   总被引:2,自引:2,他引:0  
In studies of the kinetics of oxygen uptake by glucose-stimulated Chlorella pyrenoidosa, two terminal oxidases could be distinguished. The cytochrome oxidase of Chlorella has a Km (O2) of 2.1 ± 0.3 μm, while the second oxidase has a Km (O2) of 6.7 ± 0.5 μm, and a maximum capacity about one-quarter of that of the cytochrome system. The identity of the second oxidase is unknown, but it is not inhibited by carbon monoxide, 1 mm cyanide, 0.1 mm thiocyanate, or 1 mm 8-hydroxyquinoline. In fresh cultures, the second oxidase accounts for at most 35% of the total oxygen uptake.  相似文献   

8.
1. Aminoacyl-transfer-RNA synthetase activity in extracts prepared from tobacco leaf was increased 3–5-fold when sodium thioglycollate (30mm) and magnesium chloride (16mm) were included in the extraction medium. Omitting sucrose (0·45m) from the extraction medium did not alter the activity. 2. Activity was a linear function of enzyme concentration up to 1 disk (30mg. fresh wt.)/ml. and was not affected by dialysis at any concentration. 3. Activity increased about 13-fold above control values when a mixture of 21 amino acids and amides (1mm) was added to the reaction mixture. 4. Under the conditions used in the standard assay for aminoacyl-transfer-RNA synthetase activity Km (ATP) was 0·65mm and Km (l-amino acids) was 70μm. 5. Activity above the control value was found with all amino acids and amides tested except alanine, arginine, glutamic acid, glutamine and hydroxyproline. Activity was highest with leucine, isoleucine, valine, cysteine and histidine. Total activity with a mixture of 21 amino acids and amides was 20% lower than the total activity of the enzymes assayed separately.  相似文献   

9.
We investigated 3,3′,5-tri-iodo-l-thyronine transport by human erythrocytes and by `ghosts'' prepared from these cells. Uptake of tri-iodothyronine by erythrocytes at 37°C was time-dependent with a maximum reached after 60min. Tracer analysis after incubation for 1min revealed only one saturable binding site, with Km 128±19nm (mean±s.e.m.; n=7) and Vmax. 4.6±0.7pmol of tri-iodothyronine/min per 6×107 cells. After 10min incubation Km 100±16nm (n=10) was found with Vmax. 7.7±1.2pmol of tri-iodothyronine/10min per 6×107 cells. At 0°C the uptake system is still active, with Km 132±26nm and Vmax. 1.8±0.3pmol of tri-iodothyronine/10min per 6×107 cells. The Vmax. with intact cells is 5-fold greater than the Vmax. with membranes derived from the same amount of cells when uptake studies are performed in media with similar free tri-iodothyronine concentrations. This indicates that at least 80% of tri-iodothyronine taken up by the intact erythrocytes enters the cell. This saturable uptake system can be inhibited by X-ray-contrast agents in a dose-dependent fashion. (±)-Propranolol, but not atenolol, has the same effect, indicating that the membrane-stabilizing properties of (±)-propranolol are involved. Furthermore, there is no inhibition by ouabain or vanadate, which indicates that tri-iodothyronine uptake is not dependent on the activity of Na++K+-dependent adenosine triphosphatase. We have prepared erythrocyte `ghosts'', resealed after 2.5min with 0mm-, 2mm- or 4mm-ATP inside. Inclusion of ATP and integrity of the membrane of the erythrocyte `ghosts'' were verified on the basis of an ATP-concentration-dependent functioning of the Ca2+ pump. No difference was found in the uptake of tri-iodothyronine by erythrocyte `ghosts'' with or without ATP included, indicating that uptake of tri-iodothyronine is not ATP-dependent. The following conclusions are drawn. (1) Tri-iodothyronine enters human erythrocytes. (2) There is only one saturable uptake system present for tri-iodothyronine, which is neither energy (i.e. ATP)-dependent nor influenced by the absence of an Na+ gradient across the plasma membrane. This mode of uptake of tri-iodothyronine by human erythrocytes is in sharp contrast with that of rat hepatocytes, which uptake system is energy-dependent and ouabain-sensitive [Krenning, Docter, Bernard, Visser & Hennemann (1978) FEBS Lett. 91, 113–116; Krenning, Docter, Bernard, Visser & Hennemann (1980) FEBS Lett. 119, 279–282]. (3) X-ray-contrast agents inhibit tri-iodothyronine uptake by erythrocytes in a similar fashion to that by which they inhibit the uptake of tri-iodothyronine by rat hepatocytes [Krenning, Docter, Bernard, Visser & Hennemann (1982) FEBS Lett. 140, 229–233].  相似文献   

10.
Mycobacterium tuberculosis (Mtb) and Rhodococcus jostii RHA1 have similar cholesterol catabolic pathways. This pathway contributes to the pathogenicity of Mtb. The hsaAB cholesterol catabolic genes have been predicted to encode the oxygenase and reductase, respectively, of a flavin-dependent mono-oxygenase that hydroxylates 3-hydroxy-9,10-seconandrost-1,3,5(10)-triene-9,17-dione (3-HSA) to a catechol. An hsaA deletion mutant of RHA1 did not grow on cholesterol but transformed the latter to 3-HSA and related metabolites in which each of the two keto groups was reduced: 3,9-dihydroxy-9,10-seconandrost-1,3,5(10)-triene-17-one (3,9-DHSA) and 3,17-dihydroxy-9,10-seconandrost-1,3,5(10)-triene-9-one (3,17-DHSA). Purified 3-hydroxy-9,10-seconandrost-1,3,5(10)-triene-9,17-dione 4-hydroxylase (HsaAB) from Mtb had higher specificity for 3-HSA than for 3,17-DHSA (apparent kcat/Km = 1000 ± 100 m−1 s−1 versus 700 ± 100 m−1 s−1). However, 3,9-DHSA was a poorer substrate than 3-hydroxybiphenyl (apparent kcat/Km = 80 ± 40 m−1 s−1). In the presence of 3-HSA the Kmapp for O2 was 100 ± 10 μm. The crystal structure of HsaA to 2.5-Å resolution revealed that the enzyme has the same fold, flavin-binding site, and catalytic residues as p-hydroxyphenyl acetate hydroxylase. However, HsaA has a much larger phenol-binding site, consistent with the enzyme''s substrate specificity. In addition, a second crystal form of HsaA revealed that a C-terminal flap (Val367–Val394) could adopt two conformations differing by a rigid body rotation of 25° around Arg366. This rotation appears to gate the likely flavin entrance to the active site. In docking studies with 3-HSA and flavin, the closed conformation provided a rationale for the enzyme''s substrate specificity. Overall, the structural and functional data establish the physiological role of HsaAB and provide a basis to further investigate an important class of monooxygenases as well as the bacterial catabolism of steroids.  相似文献   

11.
Accumulation of d-leucine, d-allo-isoleucine, and d-valine was observed in the growth medium of a lactic acid bacterium, Lactobacillus otakiensis JCM 15040, and the racemase responsible was purified from the cells and identified. The N-terminal amino acid sequence of the purified enzyme was GKLDKASKLI, which is consistent with that of a putative γ-aminobutyrate aminotransferase from Lactobacillus buchneri. The putative γ-aminobutyrate aminotransferase gene from L. buchneri JCM 1115 was expressed in recombinant Escherichia coli and then purified to homogeneity. The enzyme catalyzed the racemization of a broad spectrum of nonpolar amino acids. In particular, it catalyzed at high rates the epimerization of l-isoleucine to d-allo-isoleucine and d-allo-isoleucine to l-isoleucine. In contrast, the enzyme showed no γ-aminobutyrate aminotransferase activity. The relative molecular masses of the subunit and native enzyme were estimated to be about 49 kDa and 200 kDa, respectively, indicating that the enzyme was composed of four subunits of equal molecular masses. The Km and Vmax values of the enzyme for l-isoleucine were 5.00 mM and 153 μmol·min−1·mg−1, respectively, and those for d-allo-isoleucine were 13.2 mM and 286 μmol·min−1·mg−1, respectively. Hydroxylamine and other inhibitors of pyridoxal 5′-phosphate-dependent enzymes completely blocked the enzyme activity, indicating the enzyme requires pyridoxal 5′-phosphate as a coenzyme. This is the first evidence of an amino acid racemase that specifically catalyzes racemization of nonpolar amino acids at the C-2 position.  相似文献   

12.
5-Oxoprolinase has been found to be widely distributed in higher plants. This enzyme catalyzes the ATP-dependent hydrolysis of 5-oxo-l-proline (l-pyrollidone carboxylate, l-pyroglutamate) to glutamate. The enzyme has been purified almost 60 fold from wheat germ (Triticum aestivum L). This enzyme requires a divalent cation, either Mn2+ or Mg2+, and a combination of both appears to be the most effective. There is also an absolute requirement for a monovalent cation best fulfilled by either NH4+ or K+. The Km for ATP is 0.4 mm and for 5-oxo-l-proline is 14 μm. A small amount of activity is observed when other purine nucleotides such as ITP and GTP replace ATP. The substitution of the pyrimidine nucleotides CTP and UTP for ATP yield almost completely inactive preparations. The enzyme appears to have an active sulfhydryl group since there is an increase in activity in the presence of dithioerythritol. Preincubation with reagents such as N-ethylmaleimide or iodoacetamide lead to complete inactivation. The presence of this enzyme leads to the speculation of the possible presence of a γ-glutamyl cycle in higher plants.  相似文献   

13.
Both cyclic AMP phosphodiesterase and cyclic GMP phosphodiesterase were recovered mainly from the supernatant fractions of guinea-pig pancreas, but a higher proportion of the activity of the former was associated with the pellet fractions. The activities in the supernatant were not separated by gel filtration, but were clearly separated by subsequent chromatography on an anion-exchange resin. The activities of cyclic AMP phosphodiesterase and cyclic GMP phosphodiesterase had high-affinity (Km 6.5±1.1μm and 31.9±3.9μm respectively) and low-affinity (Km 0.56±0.05mm and 0.32±0.03mm respectively) components. The activity of neither enzyme was affected by the pancreatic secretogens, cholecystokinin-pancreozymin, secretin and carbachol. Removal of ions by gel filtration resulted in a marked reduction in cyclic nucleotide phosphodiesterase activity, which could be restored by addition of Mg2+. Mn2+ (3mm) was as effective as Mg2+ (3mm) in the case of cyclic AMP phosphodiesterase, but was less than half as effective in the case of cyclic GMP phosphodiesterase. The metal-ion chelators, EDTA and EGTA, also decreased activity. Ca2+ (1mm) did not affect the activity of cyclic nucleotide phosphodiesterase when the concentration of Mg2+ was 3mm. At concentrations of Mg2+ between 0.1 and 1mm, 1mm-Ca2+ was activatory, and at concentrations of Mg2+ below 0.1mm, 1mm-Ca2+ was inhibitory. These results are discussed in terms of the possible significance of cyclic nucleotide phosphodiesterase in the physiological control of cyclic nucleotide concentrations during stimulus–secretion coupling.  相似文献   

14.
Glucuronokinase from Lilium longiflorum pollen was purified 30- to 40- fold on a blue dextran-Sepharose column. Substrate analogs were tested for inhibitory effects, and nucleotide substrate specificity of the enzyme was determined. Nine nucleotides were tested, and all were inhibitory when the substrate was ATP. ADP was competitive with ATP and had a Ki value of 0.23 mm. None of the other nucleotide triphosphates could effectively substitute for ATP as a nucleotide substrate. Ten mm dATP and ITP reacted only 3% as rapidly as 10 mm ATP, while the rates for 10 mm GTP, CTP, UTP, and TTP were less than 1%. The glucuronic acid analogs, methyl α-glucuronoside, methyl β-glucuronoside, β-glucuronic acid-1-phosphate, and 4-O-methylglucuronic acid were tested as possible enzyme inhibitors. The three methyl derivatives showed little or no inhibition. The β-glucuronic acid-1-phosphate was inhibitory, with 50% inhibition obtained at 1 to 3 mm depending on the concentration of the glucuronic acid. It is concluded that the glucuronic acid-binding site on the enzyme is highly selective.  相似文献   

15.

Background and Aims

There is a need to evaluate the salt tolerance of plant species that can be cultivated as crops under saline conditions. Crambe maritima is a coastal plant, usually occurring on the driftline, with potential use as a vegetable crop. The aim of this experiment was to determine the growth response of Crambe maritima to various levels of airborne and soil-borne salinity and the ecophysiological mechanisms underlying these responses.

Methods

In the greenhouse, plants were exposed to salt spray (400 mm NaCl) as well as to various levels of root-zone salinity (RZS) of 0, 50, 100, 200 and 300 mm NaCl during 40 d. The salt tolerance of Crambe maritima was assessed by the relative growth rate (RGR) and its components. To study possible salinity effects on the tissue and cellular level, the leaf succulence, tissue Na+ concentrations, Na+ : K+ ratio, net K+/Na+ selectivity, N, P, K+, Ca2+, Mg2+, proline, soluble sugar concentrations, osmotic potential, total phenolics and antioxidant capacity were measured.

Key Results

Salt spray did not affect the RGR of Crambe maritima. However, leaf thickness and leaf succulence increased with salt spray. Root zone salinities up to 100 mm NaCl did not affect growth. However, at 200 mm NaCl RZS the RGR was reduced by 41 % compared with the control and by 56 % at 300 mm NaCl RZS. The reduced RGR with increasing RZS was largely due to the reduced specific leaf area, which was caused by increased leaf succulence as well as by increased leaf dry matter content. No changes in unit leaf rate were observed but increased RZS resulted in increased Na+ and proline concentrations, reduced K+, Ca2+ and Mg2+ concentrations, lower osmotic potential and increased antioxidant capacity. Proline concentrations of the leaves correlated strongly (r = 0·95) with RZS concentrations and not with plant growth.

Conclusions

Based on its growth response, Crambe maritima can be classified as a salt spray tolerant plant that is sensitive to root zone salinities exceeding 100 mm NaCl.  相似文献   

16.
The nicotinamide adenine dinucleotide-specific glutamate dehydrogenase (l-glutamate:NAD+ oxidoreductase, EC 1.4.1.2) of Chlorella sorokiniana was purified 1,000-fold to electrophoretic homogeneity. The native enzyme was shown to have a molecular weight of 180,000 and to be composed of four identical subunits with a molecular weight of 45,000. The N-terminal amino acid was determined to be lysine. The pH optima for the aminating and deaminating reactions were approximately 8 and 9, respectively. The Km values for α-ketoglutarate, NADH, NH4+, NAD+, and l-glutamate were 2 mm, 0.15 mm, 40 mm, 0.15 mm, and 60 mm, respectively. Whereas the Km for α-ketoglutarate and l-glutamate increased 10-fold, 1 pH unit above or below the pH optima for the aminating or deaminating reactions, respectively, the Km values for NADH and NAD+ were independent of change in pH from 7 to 9.6. By initial velocity, product inhibition, and equilibrium substrate exchange studies, the kinetic mechanism of enzyme was shown to be consistent with a bi uni uni uni ping-pong addition sequence. Although this kinetic mechanism differs from that reported for any other glutamate dehydrogenase, the chemical mechanism still appears to involve the formation of a Schiff base between α-ketoglutarate and an ε-amino group of a lysine residue in the enzyme. The physical, chemical, and kinetic properties of this enzyme differ greatly from those reported for the NH4+-inducible glutamate dehydrogenase in this organism.  相似文献   

17.
V-ATPase (VoV1) converts the chemical free energy of ATP into an ion-motive force across the cell membrane via mechanical rotation. This energy conversion requires proper interactions between the rotor and stator in VoV1 for tight coupling among chemical reaction, torque generation, and ion transport. We developed an Escherichia coli expression system for Enterococcus hirae VoV1 (EhVoV1) and established a single-molecule rotation assay to measure the torque generated. Recombinant and native EhVoV1 exhibited almost identical dependence of ATP hydrolysis activity on sodium ion and ATP concentrations, indicating their functional equivalence. In a single-molecule rotation assay with a low load probe at high ATP concentration, EhVoV1 only showed the “clear” state without apparent backward steps, whereas EhV1 showed two states, “clear” and “unclear.” Furthermore, EhVoV1 showed slower rotation than EhV1 without the three distinct pauses separated by 120° that were observed in EhV1. When using a large probe, EhVoV1 showed faster rotation than EhV1, and the torque of EhVoV1 estimated from the continuous rotation was nearly double that of EhV1. On the other hand, stepping torque of EhV1 in the clear state was comparable with that of EhVoV1. These results indicate that rotor-stator interactions of the Vo moiety and/or sodium ion transport limit the rotation driven by the V1 moiety, and the rotor-stator interactions in EhVoV1 are stabilized by two peripheral stalks to generate a larger torque than that of isolated EhV1. However, the torque value was substantially lower than that of other rotary ATPases, implying the low energy conversion efficiency of EhVoV1.  相似文献   

18.
1. The aerobic transport of d-glucose and d-galactose in rabbit kidney tissue at 25° was studied. 2. In slices forming glucose from added substrates an accumulation of glucose against its concentration gradient was found. The apparent ratio of intracellular ([S]i) and extracellular ([S]o) glucose concentrations was increased by 0·4mm-phlorrhizin and 0·3mm-ouabain. 3. Slices and isolated renal tubules actively accumulated glucose from the saline; the apparent [S]i/[S]o fell below 1·0 only at [S]o higher than 0·5mm. 4. The rate of glucose oxidation by slices was characterized by the following parameters: Km 1·16mm; Vmax. 4·5μmoles/g. wet wt./hr. 5. The active accumulation of glucose from the saline was decreased by 0·1mm-2,4-dinitrophenol, 0·4mm-phlorrhizin and by the absence of external Na+. 6. The kinetic parameters of galactose entry into the cells were: Km 1·5mm; Vmax 10μmoles/g. wet wt./hr. 7. The efflux kinetics from slices indicated two intracellular compartments for d-galactose. The galactose efflux was greatly diminished at 0°, was inhibited by 0·4mm-phlorrhizin, but was insensitive to ouabain. 8. The following mechanism of glucose and galactose transport in renal tubular cells is suggested: (a) at the tubular membrane, these sugars are actively transported into the cells by a metabolically- and Na+-dependent phlorrhizin-sensitive mechanism; (b) at the basal cell membrane, these sugars are transported in accordance with their concentration gradient by a phlorrhizin-sensitive Na+-independent facilitated diffusion. The steady-state intracellular sugar concentration is determined by the kinetic parameters of active entry, passive outflow and intracellular utilization.  相似文献   

19.
Vessal M  Hassid WZ 《Plant physiology》1973,51(6):1055-1060
d-Glucosamine-6-P N-acetyltransferase (EC 2.3.1.4) from mung bean seeds (Phaseolus aureus) was purified 313-fold by protamine sulfate and isoelectric precipitation, ammonium sulfate and acetone fractionation, and CM Sephadex column chromatography. The partially purified enzyme was highly specific for d-glucosamine-6-P. Neither d-glucosamine nor d-galactosamine could replace this substrate. The partially purified enzyme preparation was inhibited up to 50% by 2 × 10−2m EDTA, indicating the requirement of a divalent cation. Among divalent metal ions tested, Mg2+ was required for maximum activity of the enzyme. Mn2+ and Zn2+ were inhibitory, while Co2+ had no effect on the enzyme activity. The pH optimum of the enzyme in sodium acetate and sodium citrate buffers was found to be 5.2. The effect of Mg2+ on the enzyme in sodium acetate and sodium citrate buffers was particularly noticeable in the range of optimum pH. Km values of 15.1 × 10−4m and 7.1 × 10−4m were obtained for d-glucosamine-6-P and acetyl CoA, respectively. The enzyme was completely inhibited by 1 × 10−4mp-hydroxymercuribenzoate, and this inhibition was partially reversed by l-cysteine; indicating the presence of sulfhydryl groups at or near the active site of the enzyme.  相似文献   

20.
1. Rat tissue homogenates convert dl-1-aminopropan-2-ol into aminoacetone. Liver homogenates have relatively high aminopropanol-dehydrogenase activity compared with kidney, heart, spleen and muscle preparations. 2. Maximum activity of liver homogenates is exhibited at pH9·8. The Km for aminopropanol is approx. 15mm, calculated for a single enantiomorph, and the maximum activity is approx. 9mμmoles of aminoacetone formed/mg. wet wt. of liver/hr.at 37°. Aminoacetone is also formed from l-threonine, but less rapidly. An unidentified amino ketone is formed from dl-4-amino-3-hydroxybutyrate, the Km for which is approx. 200mm at pH9·8. 3. Aminopropanol-dehydrogenase activity in homogenates is inhibited non-competitively by dl-3-hydroxybutyrate, the Ki being approx. 200mm. EDTA and other chelating agents are weakly inhibitory, and whereas potassium chloride activates slightly at low concentrations, inhibition occurs at 50–100mm. 4. It is concluded that aminopropanol-dehydrogenase is located in mitochondria, and in contrast with l-threonine dehydrogenase can be readily solubilized from mitochondrial preparations by ultrasonic treatment. 5. Soluble extracts of disintegrated mitochondria exhibit maximum aminopropanol-dehydrogenase activity at pH9·1 At this pH, Km values for the amino alcohol and NAD+ are approx. 200 and 1·3mm respectively. Under optimum conditions the maximum velocity is approx. 70mμmoles of aminoacetone formed/mg. of protein/hr. at 37°. Chelating agents and thiol reagents appear to have little effect on enzyme activity, but potassium chloride inhibits at all concentrations tested up to 80mm. dl-3-Hydroxybutyrate is only slightly inhibitory. 6. Dehydrogenase activities for l-threonine and dl-4-amino-3-hydroxybutyrate appear to be distinct from that for aminopropanol. 7. Intraperitoneal injection of aminopropanol into rats leads to excretion of aminoacetone in the urine. Aminoacetone excretion proportional to the amount of the amino alcohol administered, is complete within 24hr., but represents less than 0·1% of the dose given. 8. The possible metabolic role of amino alcohol dehydrogenases is discussed.  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