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.
beta-Glucosidase is a key enzyme in the hydrolysis of cellulose to D-glucose. beta-Glucosidase was purified from cultures of Trichoderma reesei QM 9414 grown on wheat straw as carbon source. The enzyme hydrolyzed cellobiose and aryl beta-glucosides. The double-reciprocal plots of initial velocity vs. substrate concentration showed substrate inhibition with cellobiose and salicin. However, when p-nitrophenyl beta-D-glucopyranoside was the substrate no inhibition was observed. The corresponding kinetic parameters were: K = 1.09 +/- 0.2 mM and V = 2.09 +/- 0.52 mumol.min-1.mg-1 for salicin; K = 1.22 +/- 0.3 mM and V = 1.14 +/- 0.21 mumol.min-1.mg-1 for cellobiose; K = 0.19 +/- 0.02 mM and V = 29.67 +/- 3.25 mumol.min-1.mg-1 for p-nitrophenyl beta-D-glucopyranoside. Studies of inhibition by products and by alternative product supported an Ordered Uni Bi mechanism for the reaction catalyzed by beta-glucosidase on p-nitrophenyl beta-D-glucopyranoside as substrate. Alternative substrates as salicin and cellobiose, a substrate analog such as maltose and a product analog such as fructose were competitive inhibitors in the p-nitrophenyl beta-D-glucopyranoside hydrolysis.  相似文献   

2.
This study tested the hypothesis that an acute infusion of beta-hydroxybutyrate inhibits myocardial fatty acid uptake and oxidation in vivo. Anesthetized pigs were untreated (n = 6) or treated with an intravenous infusion of fat emulsion (n = 7) to elevate plasma free fatty acid levels. A third group received fat emulsion plus an intravenous infusion of beta-hydroxybutyrate (25 micromol.kg-1.min-1; n = 7) for 60 min. All animals received a continuous infusion of [3H]palmitate, and myocardial fatty acid oxidation was measured from the cardiac production of 3H2O. Plasma free fatty acid concentrations were elevated in the fat emulsion group (0.77 +/- 0.11 mM) compared with the untreated group (0.15 +/- 0.03 mM), which resulted in greater myocardial free fatty acid oxidation. In contrast, the group receiving beta-hydroxybutyrate in addition to fat emulsion had elevated beta-hydroxybutyrate concentration (0.87 +/- 0.11 vs. 0.04 +/- 0.01 mM), but suppressed fatty acid oxidation (0.053 +/- 0.013 micromol.g-1.min-1) (P < 0.05) compared with the fat emulsion group (0.116 +/- 0.029 micromol.g-1.min-1). There were no differences among the three groups in the tissue content for malonyl-CoA, acetyl-CoA, or free CoA or the activity of acetyl-CoA carboxylase; thus the inhibition of fatty acid oxidation by elevated beta-hydroxybutyrate did not appear to be due to malonyl-CoA inhibition of carnitine palmitoyl transferase-I or to an increase in the acetyl-CoA-to-free CoA ratio. In conclusion, fatty acid uptake and oxidation is blocked by an infusion of beta-hydroxybutyrate; this effect was not due to elevated myocardial malonyl-CoA content.  相似文献   

3.
Efflux of branched chain alpha-keto acids from preloaded rat heart mitochondria was slow at low external pH. Efflux was first order, and measured rate constants, kappa efflux, were 0.104 +/- 0.005 and 0.115 +/- 0.006 min-1 for alpha-ketoisovalerate and alpha-ketoisocaproate (KIC), respectively. Efflux was stimulated significantly by branched chain alpha-keto acids and related carboxylates such as alpha-ketocaproate and alpha-ketovalerate, but not by substrates for the pyruvate transporter. KIC was the preferred substrate, and the apparent exchange K0.5 for KIC was 0.14 +/- 0.10 mM. Exchange was 7-8-fold faster than efflux, and the maximal rate of exchange at saturating concentrations of alpha-ketoisovalerate and KIC appeared to be independent of the metabolite used. It is proposed that branched chain alpha-keto acids cross the inner mitochondrial membrane on a specific transporter. Transport occurs with a proton, i.e. by proton symport, and is sensitive to inhibition by cinnamic acid derivatives.  相似文献   

4.
Eubanks LM  Poulter CD 《Biochemistry》2003,42(4):1140-1149
1-Deoxy-d-xylulose 5-phosphate synthase (DXP synthase) catalyzes the thiamine diphosphate (TPP)-dependent condensation of pyruvate and d-glyceraldehyde 3-phosphate (GAP) to yield DXP in the first step of the methylerythritol phosphate pathway for isoprenoid biosynthesis. Steady-state kinetic constants for DXP synthase calculated from the initial velocities measured at varying concentrations of substrates were as follows: k(cat) = 1.9 +/- 0.1 s(-1), K(m)(GAP) = 0.068 +/- 0.001 mM, and K(m)(pyruvate) = 0.44 +/- 0.05 mM for pyruvate and GAP; k(cat) = 1.7 +/- 0.1 s(-1), K(m)(d-glyceraldehyde) = 33 +/- 3 mM, and K(m)(pyruvate) = 1.9 +/- 0.5 mM for d-glyceraldehyde and pyruvate. beta-Fluoropyruvate was investigated as a dead-end inhibitor for pyruvate. Double-reciprocal plots showed a competitive inhibition pattern with respect to pyruvate and noncompetitive inhibition with respect to GAP/d-glyceraldehyde. (14)CO(2) trapping experiments demonstrated that the binding of both substrates (pyruvate and GAP/d-glyceraldehyde) is required for the formation of a catalytically competent enzyme-substrate complex. These results are consistent with an ordered mechanism for DXP synthase where pyruvate binds before GAP/d-glyceraldehyde.  相似文献   

5.
Calcium content and distribution of the 25-day pregnant (PR) and post partum (PP) rabbit uterus was studied by atomic absorption spectrophotometry and 45Ca determination. Total Ca content [2.28 +/- 0.28 (PR) and 2.19 +/- 0.12 (PP) mM/kg wet wt] extracellular [1.21 +/- 0.09 (PR) and 1.25 +/- 0.11 (PP) mM/kg wet wt] cellular [1.07 +/- 0.08 (PR) and 0.94 +/- 0.09 (PP) mM/kg wet et], total exchangeable [1.86 +/- 0.11 (PR) and 1.84 +/- 0.09 (PR) mM/kg wet wt] and inexchangeable [0.43 +/- 0.05 (PR) and 0.35 +/- 0.04 (PP) mM/kg wet wt] Ca fractions were identical in the two extreme endocrinological conditions. In contrast compartment size and rate constant of different exchangeable Ca fractions determined by kinetic analysis of 45Ca desaturation "urves (curve-peeling tecnique and computer method), revealed significant differences between PR and PP uteri. Two exchangeable phases could be identified in both endocrinological states. The rate constants of both phases of efflux were significantly higher in the PP (alpha 1 = 0.173 +/- 0.02 min-1; alpha 2 = 0.023 +/- 0.001 min-1) than in the PR uterus (alpha 1 = 0.099 +/- 0.01 min-1; alpha 2 = 0.018 +/- 0.01 min-1). Compartment size of phase 1 (fast component) was significantly higher in the PR (1.13 +/- 0.1 mM/kg wet wt) than in the PP uterus (0.77 +/- 0.06 mM/kg wet wt). In contrast, compartment size of phase 2 (slow component) was significantly smaller in PR than in PP uterine strips (0.74 +/- 0.06 and 1.08 +/- 0.11 mM/kg wet wt). The last portion of desaturation curves represents efflux from one homogenous compartment. The present results suggest that endocrinological control of the rabbit myometrium is linked to the regulation of the binding of a superficial exchangeable Ca fraction.  相似文献   

6.
The reaction velocity of glucose-6-phosphate dehydrogenase (G6PDH) and phosphogluconate dehydrogenase (PGDH) was quantified with a cytophotometer by continuous monitoring of the reaction product as it was formed in liver cryostat sections from normal, young mature female rats at 37 degrees C. Control incubations were performed in media lacking both substrate and coenzyme for G6PDH activity and lacking substrate for PGDH activity. All reaction rates were non-linear but test minus control reactions showed linearity with incubation time up to 5 min using Nitro BT as final electron acceptor. End point measurements after incubation for 5 min at 37 degrees C revealed that the highest specific activity of G6PDH was present in the intermediate area (Vmax = 7.79 +/- 1.76 mumol H2 cm-3 min-1) and of PGDH in the pericentral and intermediate areas (Vmax = 17.19 +/- 1.73 mumol H2 cm-3 min-1). In periportal and pericentral areas, Vmax values for G6PDH activity were 4.48 +/- 1.03 mumol H2 cm-3 min-1) and 3.47 +/- 0.78 mumol H2 cm-3 min-1), respectively. PGDH activity in periportal areas showed a Vmax of 10.84 +/- 0.33 mumol H2 cm3 min-1. Variation of the substrate concentration for G6PDH activity yielded similar KM values of 0.17 +/- 0.07 mM, 0.15 +/- 0.13 mM and 0.22 +/- 0.11 mM in periportal, pericentral and intermediate areas, respectively. KM values of 0.87 +/- 0.12 mM in periportal and of 1.36 +/- 0.10 mM in pericentral and intermediate areas were found for PGDH activity. The significant difference between KM values for PGDH in areas within the acinus support the hypothesis that PGDH is present in the cytoplasmic matrix and in the microsomes. A discrepancy existed between KM and Vmax values determined in cytochemical assays using cryostat sections and values calculated from biochemical assays using diluted homogenates. In cytochemical assays, the natural microenvironment for enzymes is kept for the demonstration of their activity and thus may give more accurate information on enzyme reactions as they take place in vivo.  相似文献   

7.
The effects of components of the transition state analog (creatine, MgADP, planar anion) on the kinetics and conformation of creatine kinase isozyme BB from monkey brain was studied. From analysis of the reaction time course using the pH stat assay, it was shown that during accumulation of the reaction products (ADP and creatine phosphate), among several anions added, nitrate proved the most effective in inhibiting catalytic activity. Maximum inhibition (77%) was achieved with 50 mM nitrate. The Km for ATP was 0.48 mM and in the presence of 2.5 mM nitrate, 2.2 mM; for ATP in the presence of the dead-end complex, creatine and ADP, the apparent Km was 2.0 mM and the Ki was 0.16 mM; in the presence of the transition state analog, MgADP + NO3- + creatine, the Ki was estimated to be 0.04 mM. Ultraviolet difference spectra of creatine kinase revealed significant differences only in the presence of the complete mixture of the components of the transition state analog. Comparison of gel filtration elution profiles for creatine kinase in the absence and presence of the complete mixture of components of the transition state analog did not reveal any differences in elution volume. Addition of components of the transition state analog to creatine kinase resulted in only a marginal change in intrinsic fluorescence. The presence of the components of the transition state analog increased the rate of reactivity of the enzyme with trinitrobenzenesulfonic acid from k = 6.06 +/- 0.05 M-1 min-1 to 6.96 +/- 0.11 M-1 min-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Iron release from ovotransferrin in acidic media (3 < pH < 6) occurs in at least six kinetic steps. The first is a very fast (相似文献   

9.
An important aspect of the catalytic mechanism of microsomal glutathione transferase (MGST1) is the activation of the thiol of bound glutathione (GSH). GSH binding to MGST1 as measured by thiolate anion formation, proton release, and Meisenheimer complex formation is a slow process that can be described by a rapid binding step (K(GSH)d = 47 +/- 7 mM) of the peptide followed by slow deprotonation (k2 = 0.42 +/- 0.03 s(-1). Release of the GSH thiolate anion is very slow (apparent first-order rate k(-2) = 0.0006 +/- 0.00002 s(-)(1)) and thus explains the overall tight binding of GSH. It has been known for some time that the turnover (kcat) of MGST1 does not correlate well with the chemical reactivity of the electrophilic substrate. The steady-state kinetic parameters determined for GSH and 1-chloro-2,4-dinitrobenzene (CDNB) are consistent with thiolate anion formation (k2) being largely rate-determining in enzyme turnover (kcat = 0.26 +/- 0.07 s(-1). Thus, the chemical step of thiolate addition is not rate-limiting and can be studied as a burst of product formation on reaction of halo-nitroarene electrophiles with the E.GS- complex. The saturation behavior of the concentration dependence of the product burst with CDNB indicates that the reaction occurs in a two-step process that is characterized by rapid equilibrium binding ( = 0.53 +/- 0.08 mM) to the E.GS- complex and a relatively fast chemical reaction with the thiolate (k3 = 500 +/- 40 s(-1). In a series of substrate analogues, it is observed that log k3 is linearly related (rho value 3.5 +/- 0.3) to second substrate reactivity as described by Hammett sigma- values demonstrating a strong dependence on chemical reactivity that is similar to the nonenzymatic reaction (rho = 3.4). Microsomal glutathione transferase 1 displays the unusual property of being activated by sulfhydryl reagents. When the enzyme is activated by N-ethylmaleimide, the rate of thiolate anion formation is greatly enhanced, demonstrating for the first time the specific step that is activated. This result explains earlier observations that the enzyme is activated only with more reactive substrates. Taken together, the observations show that the kinetic mechanism of MGST1 can be described by slow GSH binding/thiolate formation followed by a chemical step that depends on the reactivity of the electrophilic substrate. As the chemical reactivity of the electrophile becomes lower the rate-determining step shifts from thiolate formation to the chemical reaction.  相似文献   

10.
The determination of glyoxalase II (S-(2-hydroxyacyl)glutathione hydrolase, EC 3.1.2.6) activity is usually accomplished by monitoring the decrease of absorbance at 240 nm due to the hydrolysis of S-d-lactoylglutathione. However, it was not possible, using this assay, to detect any enzyme activity in situ, in Saccharomyces cerevisiae permeabilized cells. Glyoxalase II activity was then determined by following the formation of GSH at 412 nm using 5,5'-dithiobis(2-nitrobenzoic acid). Using this method we characterized the kinetics of glyoxalase II in situ using S-d-lactoylglutathione as substrate and compared the results with those obtained for cell-free extracts. The specific activity was found to be (4.08 +/- 0.12) x 10(-2) micromol min-1 mg-1 in permeabilized cells and (3.90 +/- 0.04) x 10(-2) micromol min1 mg-1 in cell-free extracts. Kinetic parameters were Km 0.36 +/- 0.09 mM and V (7.65 +/- 0.59) x 10(-4) mM min-1 for permeabilized cells and Km 0.15 +/- 0.10 mM and V (7.23 +/- 1.04) x 10(-4) mM min-1 for cell-free extracts. d-Lactate concentration was also determined and increased in a linear way with permeabilized cell concentration. gamma-Glutamyl transferase (EC 2.3.2.2), which also accepts S-d-lactoylglutathione as substrate and hence could interfere with glyoxalase II assays, was found to be absent in Saccharomyces cerevisiae permeabilized cells.  相似文献   

11.
The kinetics of slow onset inhibition of Proteinase K by a proteinaceous alkaline protease inhibitor (API) from a Streptomyces sp. is presented. The kinetic analysis revealed competitive inhibition of Proteinase K by API with an IC50 value 5.5 +/- 0.5 x 10-5 m. The progress curves were time-dependent, consistent with a two-step slow tight binding inhibition. The first step involved a rapid equilibrium for formation of reversible enzyme-inhibitor complex (EI) with a Ki value 5.2 +/- 0.6 x 10-6 m. The EI complex isomerized to a stable complex (EI*) in the second step because of inhibitor-induced conformational changes, with a rate constant k5 (9.2 +/- 1 x 10-3 s-1). The rate of dissociation of EI* (k6) was slower (4.5 +/- 0.5 x 10-5 s-1) indicating the tight binding nature of the inhibitor. The overall inhibition constant Ki* for two-step inhibition of Proteinase K by API was 2.5 +/- 0.3 x 10-7 m. Time-dependent dissociation of EI* revealed that the complex failed to dissociate after a time point and formed a conformationally altered, irreversible complex EI**. These conformational states of enzyme-inhibitor complexes were characterized by fluorescence spectroscopy. Tryptophanyl fluorescence of Proteinase K was quenched as a function of API concentration without any shift in the emission maximum indicating a subtle conformational change in the enzyme, which is correlated to the isomerization of EI to EI*. Time-dependent shift in the emission maxima of EI* revealed the induction of gross conformational changes, which can be correlated to the irreversible conformationally locked EI** complex. API binds to the active site of the enzyme as demonstrated by the abolished fluorescence of 5-iodoacetamidofluorescein-labeled Proteinase K. The chemoaffinity labeling experiments lead us to hypothesize that the inactivation of Proteinase K is because of the interference in the electronic microenvironment and disruption of the hydrogen-bonding network between the catalytic triad and other residues involved in catalysis.  相似文献   

12.
o-Phthaldialdehyde caused irreversible inhibition of rabbit muscle pyruvate kinase following preliminary formation of an enzyme-reagent complex. At pH 7.5, 35 degrees C, the dissociation constant for the complex and the maximal pseudo-first-order rate constant for covalent modification were 0.32 +/- 0.08 mM and 2.54 +/- 0.23 min-1, respectively. The inactivation was accompanied by uv-spectral changes pointing to isoindole formation, with a limiting stoichiometry of 1 isoindole linkage per enzyme subunit. Phosphoenolpyruvate, ADP, and ATP effectively protected the enzyme against inactivation, suggesting that the active site is the target of o-phthaldialdehyde action. As native and modified enzymes were indistinguishable with respect to mobility of the major band in sodium dodecyl sulfate-polyacrylamide gel electrophoresis, it was concluded that the crosslinkage was intrasubunit in character, and that the amino acid residues involved must be closely positioned in the polypeptide backbone. Lysine 366, previously shown to be selectively reactive toward 2',3'-dialdehyde ADP (Bezares et al., 1987, Arch, Biochem. Biophys. 253, 133-137), and cysteine 325 or 357 are implicated.  相似文献   

13.
The kinetic parameters of net exit of D-glucose from human red blood cells have been measured after the cells were loaded to 18 mM, 75 mM and 120 mM at 2 degrees C and 75 mM and 120 mM at 20 degrees C. Reducing the temperature, or raising the loading concentration raises the apparent Km for net exit. Deoxygenation also reduces the Km for D-glucose exit from red blood cells loaded initially to 120 mM at 20 degrees C from 32.9 +/- 2.3 mM (13) with oxygenated blood to 20.5 +/- 1.3 mM (17) (P less than 0.01). Deoxygenation increases the ratio Vmax/Km from 5.29 +/- 0.26 min-1 (13) for oxygenated blood to 7.13 +/- 0.29 min-1 (17) for deoxygenated blood (P less than 0.001). The counterflow of D-glucose from solutions containing 1 mM 14C-labelled D-glucose was measured at 2 degrees C and 20 degrees C. Reduction in temperature, reduced the maximal level to which labelled D-glucose was accumulated and altered the course of equilibration of the specific activity of intracellular D-glucose from a single exponential to a more complex form. Raising the internal concentration from 18 mM to 90 mM at 2 degrees C also alters the course of equilibration of labelled D-glucose within the cell to a complex form. The apparent asymmetry of the transport system may be estimated from the intracellular concentrations of labelled and unlabelled sugar at the turning point of the counterflow transient. The estimates of asymmetry obtained from this approach indicate that there is no significant asymmetry at 20 degrees C and at 2 degrees C asymmetry is between 3 and 6. This is at least 20-fold less than predicted from the kinetic parameter asymmetries for net exit and entry. None of the above results fit a kinetic scheme in which the asymmetry of the transport system is controlled by intrinsic differences in the kinetic parameters at the inner and outer membrane surface. These results are consistent with a model for sugar transport in which movement between sugar within bound and free intracellular compartments can become the rate-limiting step in controlling net movement into, or out of the cell.  相似文献   

14.
Isoprenylcysteine carboxyl methyltransferase (Icmt) catalyzes the methylation of the C-terminal prenylcysteine found on prenylated proteins. Numerous studies have shown that the methylation step is important for the correct localization and function of many prenylated proteins, most notably GTPases in the Ras superfamily. We recently reported identification of a small molecule derived from an indole core as a potent, cell-active inhibitor of Icmt whose potency was increased upon preincubation with the enzyme [Winter-Vann, A. M., Baron, R. A., et al. (2005) Proc. Natl. Acad. Sci. U.S.A. 102 (12), 4336-41]. In the study presented here, we performed a kinetic characterization of this time-dependent inhibition of Icmt by 2-[5-(3-methylphenyl)-1-octyl-1H-indol-3-yl]acetamide (cysmethynil). These analyses revealed that cysmethynil is a competitive inhibitor with respect to the isoprenylated cysteine substrate and a noncompetitive inhibitor with respect to AdoMet, the methyl donor in the reaction. The Ki of cysmethynil for Icmt, which represents the dissociation constant of the initial complex with the enzyme, was 2.39 +/- 0.02 microM, and the Ki*, which is the overall dissociation constant of the inhibitor for the final complex, was 0.14 +/- 0.01 microM. The first-order rate constant for the conversion of the initial enzyme-inhibitor complex to the final high-affinity complex was 0.87 +/- 0.06 min-1, and that for the reverse process was 0.053 +/- 0.003 min-1; the latter rate constant corresponds to a half-life for the high-affinity complex of 15 min. Structure-activity relationships of a number of closely related indole compounds revealed that the hydrophobicity of the substituent on the nitrogen of the indole core was responsible for the manifestation of time-dependent inhibition. These findings markedly enhance our understanding of the mechanism of inhibition of Icmt by this indole class of compounds and should facilitate ongoing efforts to assess the potential of targeting this enzyme in anticancer drug design.  相似文献   

15.
16.
Appert C  Zoń J  Amrhein N 《Phytochemistry》2003,62(3):415-422
The conformationally restricted phenylalanine analogue 2-aminoindan-2-phosphonic acid (AIP) inhibits phenylalanine ammonia-lyase (PAL) competitively in a time-dependent manner. This phenomenon was investigated in more detail with the heterologously expressed, highly purified homotetrameric PAL-1 isozyme from parsley. The kinetic analysis revealed that the enzyme-inhibitor complex is formed in a single "slow" step with an association rate of k(2)=2.6+/-0.04 10(4) M(-1) s(-1). The inhibition is reversible with a dissociation rate of k(-2)=1.8+/-0.04 10(-4) s(-1) and an equilibrium constant of K(i)=7+/-2 nM. The previously described PAL inhibitor (S)-2-aminooxy-3-phenylpropanoic acid [(S)-AOPP] was also found to be a slow-binding inhibitor of PAL-1. The carboxyl analogue of AIP, 2-aminoindan-2-carboxylic acid, served as a substrate of PAL-1 and was converted to indene-2-carboxylic acid.  相似文献   

17.
This is the first report that describes the inhibition mechanism of xylanase from Thermomonospora sp. by pepstatin A, a specific inhibitor toward aspartic proteases. The kinetic analysis revealed competitive inhibition of xylanase by pepstatin A with an IC50 value 3.6 +/- 0.5 microm. The progress curves were time-depended, consistent with a two-step slow tight binding inhibition. The inhibition followed a rapid equilibrium step to form a reversible enzyme-inhibitor complex (EI), which isomerizes to the second enzyme-inhibitor complex (EI*), which dissociated at a very slow rate. The rate constants determined for the isomerization of EI to EI* and the dissociation of EI* were 15 +/- 1 x 10(-5) and 3.0 +/- 1 x 10(-8) s(-1), respectively. The Ki value for the formation of EI complex was 1.5 +/- 0.5 microm, whereas the overall inhibition constant Ki* was 28.0 +/- 1 nm. The conformational changes induced in Xyl I by pepstatin A were monitored by fluorescence spectroscopy, and the rate constants derived were in agreement with the kinetic data. Thus, the conformational alterations were correlated to the isomerization of EI to EI*. Pepstatin A binds to the active site of the enzyme and disturbs the native interaction between the histidine and lysine, as demonstrated by the abolished isoindole fluorescence of o-phthalaldehyde-labeled xylanase. Our results revealed that the inactivation of xylanase is due to the interference in the electronic microenvironment and disruption of the hydrogen-bonding network between the essential histidine and other residues involved in catalysis, and a model depicting the probable interaction between pepstatin A with xylanase has been proposed.  相似文献   

18.
We purified a large amount of dynamin with high enzymatical activity from rat brain tissue by a new procedure. Dynamin 0.48 mg was obtained from 20 g of rat brain. The purity of dynamin was almost 98%. Dynamin plays a role of GTPase rather than ATPase. In the absence of microtubules, Michaelis constant (Km) and maximum velocity (Vmax) for dynamin GTPase were 370 microM and 0.25 min-1, respectively, and in their presence, both were significantly accelerated up to 25 microM and 5.5 min-1. On the other hand, the ATPase activity was very low in the absence of microtubules, and even in their presence, Km and Vmax for dynamin ATPase were 0.2 mM and 0.91 min-1. Despite slow GTPase turnover rate in the absence of microtubules, binding of GTP and its nonhydrolizing analogues was very fast, indicating that GTP binding step is not rate limiting. Dynamin did not cause a one-directional consistent microtubule sliding movement just like kinesin or dynein in the presence of 2 mM ATP or 2 mM GTP. We observed the molecular structure of dynamin with low-angle rotary shadowing technique and revealed that the dynamin molecule is globular in shape. Gel filtration assay revealed that these globules were the oligomers of 100-kDa dynamin polypeptide. Dynamin bound to microtubules with a 1:1 approximately 1.2 molar ratio in the absence of GTP. Quick-freeze deep-etch electron microscopy of the dynamin-microtubule complex showed that dynamin decorates the surface of microtubules helically, like a screw bolt, very orderly and tightly with 11.4 +/- 0.9 (SD)nm period. Contrary to the previous report, microtubules make bundles by the attachment of the dynamin helixes around each adjacent microtubule, and no cross-bridge formation was observed.  相似文献   

19.
There was a reversible inhibition of urea formation in the perfused rat liver caused by 2.25-27 mM lysine acting with a Ki of 10.8 mM in competition with ornithine. Urea formation in the presence of inhibitory concentrations of lysine ranged between 2.3 and 2.9 mumol X min-1 X (g, liver wet)-1 after addition of 1 mM of citrulline, argininosuccinate or arginine, whereas it amounted to 0.5 mumol X min-1 X (g, liver wet)-1 after addition of ornithine, showing that lysine inhibited the urea cycle between ornithine and citrulline. There was a rise of basal orotate formation of 0.03 +/- 0.02 mumol X h-1 X (g, liver wet)-1 towards a maximum of 0.6 +/- 0.04 mumol X h-1 X (g, liver wet)-1 after addition of 13.5 mM lysine, provided orotate utilization was blocked with allopurinol. Maximal rates of orotate formation were reached when ammonium concentrations exceeded 1 mM. We conclude that an inhibition of urea synthesis and a rise of orotate formation are caused by lysine in the isolated liver in vitro at rates observed in vivo. Hence, these metabolic alterations observed in the whole animal are most probably due to changes of liver metabolism.  相似文献   

20.
To evaluate a possible role of ornithine-delta-aminotransferase (EC 2.6.1.13; Orn-T) as a rate-limiting enzyme for the synthesis of transmitter glutamate and gamma-aminobutyric acid (GABA), respectively, its activity and kinetic properties were analyzed in cultured astrocytes as well as in neuronal cultures consisting mainly of glutamatergic neurons (cerebellar granule cells) or GABAergic neurons (cerebral cortex interneurons). For comparison the activity and kinetics of Orn-T were also assayed in mouse brain homogenates. The highest activity of Orn-T was found in astrocytes and in cerebral cortical neurons (5.3 +/- 0.5 and 5.3 +/- 0.4 nmol X mg-1 X min-1, respectively) whereas the activities of Orn-T in cerebellar granule cell cultures and in mouse brain were found to be about half of these values (3.1 +/- 0.3 and 2.8 +/- 0.1 nmol X min-1 X mg-1, respectively). From a kinetic study of Orn-T in the different preparations only a relatively low affinity for the enzyme with respect to ornithine was found in cerebellar granule cells, astrocytes, and whole brain [apparent Km values (at 0.5 mM alpha-ketoglutarate): 4.7 +/- 0.9, 4.3 +/- 2.2, and 6.8 +/- 2.2 mM, respectively] whereas the corresponding Km value for Orn-T in cerebral cortex interneurons was found to be significantly lower (apparent Km: 0.8 +/- 0.3 mM). The enzyme was not found to be inhibited by GABA (range 0.1 - 10 mM) in any of the preparations.  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