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.
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   

2.
Pah1 is the phosphatidate phosphatase in the yeast Saccharomyces cerevisiae that produces diacylglycerol for triacylglycerol synthesis and concurrently controls the levels of phosphatidate used for phospholipid synthesis. Phosphorylation and dephosphorylation of Pah1 regulate its subcellular location and phosphatidate phosphatase activity. Compared with its phosphorylation by multiple protein kinases, Pah1 is dephosphorylated by a protein phosphatase complex consisting of Nem1 (catalytic subunit) and Spo7 (regulatory subunit). In this work, we characterized the Nem1-Spo7 phosphatase complex for its enzymological, kinetic, and regulatory properties with phosphorylated Pah1. The dephosphorylation of Pah1 by Nem1-Spo7 phosphatase resulted in the stimulation (6-fold) of phosphatidate phosphatase activity. For Pah1 phosphorylated by the Pho85-Pho80 kinase complex, maximum Nem1-Spo7 phosphatase activity required Mg2+ ions (8 mm) and Triton X-100 (0.25 mm) at pH 5.0. The energy of activation for the reaction was 8.4 kcal/mol, and the enzyme was thermally labile at temperatures above 40 °C. The enzyme activity was inhibited by sodium vanadate, sodium fluoride, N-ethylmaleimide, and phenylglyoxal but was not significantly affected by lipids or nucleotides. Nem1-Spo7 phosphatase activity was dependent on the concentrations of Pah1 phosphorylated by Pho85-Pho80, Cdc28-cyclin B, PKA, and PKC with kcat and Km values of 0.29 s−1 and 81 nm, 0.11 s−1 and 127 nm, 0.10 s−1 and 46 nm, and 0.02 s−1 and 38 nm, respectively. Its specificity constant (kcat/Km) for Pah1 phosphorylated by Pho85-Pho80 was 1.6-, 4-, and 6-fold higher, respectively, than that phosphorylated by PKA, Cdc28-cyclin B, and PKC.  相似文献   

3.
The effect of pH on the hydrolysis of α-N-benzoyl-l-arginine ethyl ester (BAEE) and α-N-benzoyl-l-argininamide (BAA) by a proteolytic enzyme component purified from Ficus carica var. Kadota latex has been studied in detail over the pH range of 3 to 9.5. kcat (lim) values for the hydrolysis of BAEE and BAA were essentially identical (5.20 and 5.01 sec−1, respectively at 30°). kcat values for hydrolysis of BAEE and BAA were dependent on prototropic groups with apparent pK values of 4.24 and 8.53 and 4.10 and 8.59, respectively. kcat (lim) values for tht hydrolysis of BAEE and BAA were essentially identical (5.20 and groups of pK 4.33 and 8.60 and 4.55 and 8.51, respectively. Thus the pH optimum is 6.5 for both substrates. Km (app) values for BAEE and BAA were 3.32 × 10−2m and 6.03 × 10−2m respectively over the pH range of 3.9 to 8.0. These data are interpreted in terms of the involvement of a carboxyl and a sulfhydryl group in the active center of the enzyme. The data do not support the concept that deacylation of the acyl-enzyme is completely the rate controlling step in the hydrolyses. Rather, it appears that the magnitude of k2 and k3 are not greatly different.  相似文献   

4.
Nitric -oxide reductase (NOR) from Paracoccus denitrificans catalyzes the reduction of nitric oxide (NO) to nitrous oxide (N2O) (2NO + 2H+ + 2e →N2O + H2O) by a poorly understood mechanism. NOR contains two low spin hemes c and b, one high spin heme b3, and a non-heme iron FeB. Here, we have studied the reaction between fully reduced NOR and NO using the “flow-flash” technique. Fully (four-electron) reduced NOR is capable of two turnovers with NO. Initial binding of NO to reduced heme b3 occurs with a time constant of ∼1 μs at 1.5 mm NO, in agreement with earlier studies. This reaction is [NO]-dependent, ruling out an obligatory binding of NO to FeB before ligation to heme b3. Oxidation of hemes b and c occurs in a biphasic reaction with rate constants of 50 s−1 and 3 s−1 at 1.5 mm NO and pH 7.5. Interestingly, this oxidation is accelerated as [NO] is lowered; the rate constants are 120 s−1 and 12 s−1 at 75 μm NO. Protons are taken up from solution concomitantly with oxidation of the low spin hemes, leading to an acceleration at low pH. This effect is, however, counteracted by a larger degree of substrate inhibition at low pH. Our data thus show that substrate inhibition in NOR, previously observed during multiple turnovers, already occurs during a single oxidative cycle. Thus, NO must bind to its inhibitory site before electrons redistribute to the active site. The further implications of our data for the mechanism of NO reduction by NOR are discussed.  相似文献   

5.
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   

6.
V-ATPases are rotary molecular motors that generally function as proton pumps. We recently solved the crystal structures of the V1 moiety of Enterococcus hirae V-ATPase (EhV1) and proposed a model for its rotation mechanism. Here, we characterized the rotary dynamics of EhV1 using single-molecule analysis employing a load-free probe. EhV1 rotated in a counterclockwise direction, exhibiting two distinct rotational states, namely clear and unclear, suggesting unstable interactions between the rotor and stator. The clear state was analyzed in detail to obtain kinetic parameters. The rotation rates obeyed Michaelis-Menten kinetics with a maximal rotation rate (Vmax) of 107 revolutions/s and a Michaelis constant (Km) of 154 μm at 26 °C. At all ATP concentrations tested, EhV1 showed only three pauses separated by 120°/turn, and no substeps were resolved, as was the case with Thermus thermophilus V1-ATPase (TtV1). At 10 μm ATP (⪡Km), the distribution of the durations of the ATP-waiting pause fit well with a single-exponential decay function. The second-order binding rate constant for ATP was 2.3 × 106 m−1 s−1. At 40 mm ATP (⪢Km), the distribution of the durations of the catalytic pause was reproduced by a consecutive reaction with two time constants of 2.6 and 0.5 ms. These kinetic parameters were similar to those of TtV1. Our results identify the common properties of rotary catalysis of V1-ATPases that are distinct from those of F1-ATPases and will further our understanding of the general mechanisms of rotary molecular motors.  相似文献   

7.
Human cystathionine β-synthase (CBS), a novel heme-containing pyridoxal 5′-phosphate enzyme, catalyzes the condensation of homocysteine and serine or cysteine to produce cystathionine and H2O or H2S, respectively. The presence of heme in CBS has limited spectrophotometric characterization of reaction intermediates by masking the absorption of the pyridoxal 5′-phosphate cofactor. In this study, we employed difference stopped-flow spectroscopy to characterize reaction intermediates formed under catalytic turnover conditions. The reactions of l-serine and l-cysteine with CBS resulted in the formation of a common aminoacrylate intermediate (kobs = 0.96 ± 0.02 and 0.38 ± 0.01 mm−1 s−1, respectively, at 24 °C) with concomitant loss of H2O and H2S and without detectable accumulation of the external aldimine or other intermediates. Homocysteine reacted with the aminoacrylate intermediate with kobs = 40.6 ± 3.8 s−1 and re-formed the internal aldimine. In the reverse direction, CBS reacted with cystathionine, forming the aminoacrylate intermediate with kobs = 0.38 ± 0.01 mm−1 s−1. This study provides the first insights into the pre-steady-state kinetic mechanism of human CBS and indicates that the reaction is likely to be limited by a conformational change leading to product release.  相似文献   

8.
Glutathione (GSH) is an essential antioxidant responsible for the maintenance of intracellular redox homeostasis. As tumors outgrow their blood supply and become hypoxic, their redox homeostasis is challenged by the production of nitric oxide and reactive oxygen species (ROS). In gliomas, the sustained import of l-cystine via the l-cystine/l-glutamate exchanger, system xc, is rate-limiting for the synthesis of GSH. We show that hypoxia causes a significant increase in NO and ROS but without affecting glioma cell growth. This is explained by a concomitant increase in the utilization of GSH, which is accompanied by an increase in the cell-surface expression of xCT, the catalytic subunit of system xc, and l-cystine uptake. Growth was inhibited when GSH synthesis was blocked by buthionine sulfoximine (BSO), an inhibitor of the enzyme required for GSH synthesis, or when cells were deprived of l-cystine. These findings suggest that glioma cells show an increased requirement for GSH to maintain growth under hypoxic conditions. Therefore, approaches that limit GSH synthesis such as blocking system xc may be considered as an adjuvant to radiation or chemotherapy.  相似文献   

9.
The reversibility of adenosine triphosphate cleavage by myosin   总被引:12,自引:12,他引:0  
For the simplest kinetic model the reverse rate constants (k−1 and k−2) associated with ATP binding and cleavage on purified heavy meromyosin and heavy meromyosin subfragment 1 from rabbit skeletal muscle in the presence of 5mm-MgCl2, 50mm-KCl and 20mm-Tris–HCl buffer at pH8.0 and 22°C are: k−1<0.02s−1 and k−1=16s−1. Apparently, higher values of k−1 and k−2 are found with less-purified protein preparations. The values of k−1 and k−2 satisfy conditions required by previous 18O-incorporation studies of H218O into the Pi moiety on ATP hydrolysis and suggest that the cleavage step does involve hydrolysis of ATP or formation of an adduct between ATP and water. The equilibrium constant for the cleavage step at the myosin active site is 9. If the cycle of events during muscle contraction is described by the model proposed by Lymn & Taylor (1971), the fact that there is only a small negative standard free-energy change for the cleavage step is advantageous for efficient chemical to mechanical energy exchange during muscle contraction.  相似文献   

10.
When supplied under low chloride concentrations, vanadate inhibits the blue light-stimulated swelling of Vicia faba L. guard cell protoplasts in a dose-dependent fashion. The volume of guard cell protoplasts incubated in 10 mm K-imino-diacetic acid, 0.4 m mannitol, and 1 mm CaCl2 remained essentially constant under 1000 μmol m−2 s−1 red light, but increased an average of 27% after 8 min of the addition of 50 μmol m−2 s−1 blue light to the background red light. At 500 μm, vanadate completely inhibits the response to blue light. Vanadate also inhibits the swelling of guard cell protoplasts stimulated by the H+-ATPase agonist fusicoccin. The vanadate sensitivity of the blue light-stimulated swelling implicates a proton-pumping ATPase as a component of the sensory transduction of blue light in guard cells.  相似文献   

11.

Background and Aims

Previous research has suggested a trade-off between the capacity of plants to downregulate their phosphorus (P) uptake capacity and their efficiency of P resorption from senescent leaves in species from P-impoverished environments.

Methods

To investigate this further, four Australian native species (Banksia attenuata, B. menziesii, Acacia truncata and A. xanthina) were grown in a greenhouse in nutrient solutions at a range of P concentrations [P]. Acacia plants received between 0 and 500 µm P; Banksia plants received between 0 and 10 µm P, to avoid major P-toxicity symptoms in these highly P-sensitive species.

Key Results

For both Acacia species, the net P-uptake rates measured at 10 µm P decreased steadily with increasing P supply during growth. In contrast, in B. attenuata, the net rate of P uptake from a solution with 10 µm P increased linearly with increasing P supply during growth. The P-uptake rate of B. menziesii showed no significant response to P supply in the growing medium. Leaf [P] of the four species supported this finding, with A. truncata and A. xanthina showing an increase up to a saturation value of 19 and 21 mg P g−1 leaf dry mass, respectively (at 500 µm P), whereas B. attenuata and B. menziesii both exhibited a linear increase in leaf [P], reaching 10 and 13 mg P g−1 leaf dry mass, respectively, without approaching a saturation point. The Banksia plants grown at 10 µm P showed mild symptoms of P toxicity, i.e. yellow spots on some leaves and drying and curling of the tips of the leaves. Leaf P-resorption efficiency was 69 % (B. attenuata), 73 % (B. menziesii), 34 % (A. truncata) and 36 % (A. xanthina). The P-resorption proficiency values were 0·08 mg P g−1 leaf dry mass (B. attenuata and B. menziesii), 0·32 mg P g−1 leaf dry mass (A. truncata) and 0·36 mg P g−1 leaf dry mass (A. xanthina). Combining the present results with additional information on P-remobilization efficiency and the capacity to downregulate P-uptake capacity for two other Australian woody species, we found a strong negative correlation between these traits.

Conclusions

It is concluded that species that are adapted to extremely P-impoverished soils, such as many south-western Australian Proteaceae species, have developed extremely high P-resorption efficiencies, but lost their capacity to downregulate their P-uptake mechanisms. The results support the hypothesis that the ability to resorb P from senescing leaves is inversely related to the capacity to downregulate net P uptake, possibly because constitutive synthesis of P transporters is a prerequisite for proficient P remobilization from senescing tissues.  相似文献   

12.
Five fungal genomes from the Ascomycota (sac fungi) were found to contain a gene with sequence similarity to a recently discovered small group of bacterial prenyltransferases that catalyze the C-prenylation of aromatic substrates in secondary metabolism. The genes from Aspergillus terreus NIH2624, Botryotinia fuckeliana B05.10 and Sclerotinia sclerotiorum 1980 were expressed in Escherichia coli, and the resulting His8-tagged proteins were purified and investigated biochemically. Their substrate specificity was found to be different from that of any other prenyltransferase investigated previously. Using 2,7-dihydroxynaphthalene (2,7-DHN) and dimethylallyl diphosphate as substrates, they catalyzed a regiospecific Friedel-Crafts alkylation of 2,7-DHN at position 3. Using the enzyme of A. terreus, the Km values for 2,7-DHN and dimethylallyl diphosphate were determined as 324 ± 25 μm and 325 ± 35 μm, respectively, and kcat as 0.026 ± 0.001 s−1. A significantly lower level of prenylation activity was found using dihydrophenazine-1-carboxylic acid as aromatic substrate, and only traces of products were detected with aspulvinone E, flaviolin, or 4-hydroxybenzoic acid. No product was formed with l-tryptophan, l-tyrosine, or 4-hydroxyphenylpyruvate. The genes for these fungal prenyltransferases are not located within recognizable secondary metabolic gene clusters. Their physiological function is yet unknown.  相似文献   

13.
1. all-trans-Retinoic acid at concentrations greater than 10−7m stimulated the incorporation of d-[3H]glucosamine into 8m-urea/5% (w/v) sodium dodecyl sulphate extracts of 1m-CaCl2-separated epidermis from pig ear skin slices cultured for 18h. The incorporation of 35SO42−, l-[14C]fucose and U-14C-labelled l-amino acids was not significantly affected. 2. Electrophoresis of the solubilized epidermis showed increased incorporation of d-[3H]glucosamine into a high-molecular-weight glycosaminoglycan-containing peak when skin slices were cultured in the presence of 10−5m-all-trans-retinoic acid. The labelling of other epidermal components with d-[3H]glucosamine, 35SO42−, l-[14C]fucose and U-14C-labelled l-amino acids was not significantly affected by 10−5m-all-trans-retinoic acid. 3. Trypsinization dispersed the epidermal cells and released 75–85% of the total d-[3H]glucosamine-labelled material in the glycosaminoglycan peak. Thus most of this material was extracellular in both control and 10−5m-all-trans-retinoic acid-treated epidermis. 4. Increased labelling of extracellular epidermal glycosaminoglycans was also observed when human skin slices were treated with all-trans-retinoic acid, indicating a similar mechanism in both tissues. Increased labelling was also found when the epidermis was cultured in the absence of the dermis, suggesting a direct effect of all-trans-retinoic acid on the epidermis. 5. Increased incorporation of d-[3H]-glucosamine into extracellular epidermal glycosaminoglycans in 10−5m-all-trans-retinoic acid-treated skin slices was apparent after 4–8h in culture and continued up to 48h. all-trans-Retinoic acid (10−5m) did not affect the rate of degradation of this material in cultures `chased' with 5mm-unlabelled glucosamine after 4 or 18h. 6. Cellulose acetate electrophoresis at pH7.2 revealed that hyaluronic acid was the major labelled glycosaminoglycan (80–90%) in both control and 10−5m-all-trans-retinoic acid-treated epidermis. 7. The labelling of epidermal plasma membranes isolated from d-[3H]glucosamine-labelled skin slices by sucrose density gradient centrifugation was similar in control and 10−5m-all-trans-retinoic acid-treated tissue. 8. The results indicate that increased synthesis of mainly extracellular glycosaminoglycans (largely hyaluronic acid) may be the first response of the epidermis to excess all-trans-retinoic acid.  相似文献   

14.
Enzymatic processes are useful for industrially important sugar production, and in vitro two-step isomerization has proven to be an efficient process in utilizing readily available sugar sources. A hypothetical uncharacterized protein encoded by ydaE of Bacillus licheniformis was found to have broad substrate specificities and has shown high catalytic efficiency on d-lyxose, suggesting that the enzyme is d-lyxose isomerase. Escherichia coli BL21 expressing the recombinant protein, of 19.5 kDa, showed higher activity at 40 to 45°C and pH 7.5 to 8.0 in the presence of 1.0 mM Mn2+. The apparent Km values for d-lyxose and d-mannose were 30.4 ± 0.7 mM and 26 ± 0.8 mM, respectively. The catalytic efficiency (kcat/Km) for lyxose (3.2 ± 0.1 mM−1 s−1) was higher than that for d-mannose (1.6 mM−1 s−1). The purified protein was applied to the bioproduction of d-lyxose and d-glucose from d-xylose and d-mannose, respectively, along with the thermostable xylose isomerase of Thermus thermophilus HB08. From an initial concentration of 10 mM d-lyxose and d-mannose, 3.7 mM and 3.8 mM d-lyxose and d-glucose, respectively, were produced by two-step isomerization. This two-step isomerization is an easy method for in vitro catalysis and can be applied to industrial production.  相似文献   

15.
DNA polymerase ν (pol ν) is a low fidelity A-family polymerase with a putative role in interstrand cross-link repair and homologous recombination. We carried out pre-steady-state kinetic analysis to elucidate the kinetic mechanism of this enzyme. We found that the mechanism consists of seven steps, similar that of other A-family polymerases. pol ν binds to DNA with a Kd for DNA of 9.2 nm, with an off-rate constant of 0.013 s−1and an on-rate constant of 14 μm−1 s−1. dNTP binding is rapid with Kd values of 20 and 476 μm for the correct and incorrect dNTP, respectively. Pyrophosphorylation occurs with a Kd value for PPi of 3.7 mm and a maximal rate constant of 11 s−1. Pre-steady-state kinetics, examination of the elemental effect using dNTPαS, and pulse-chase experiments indicate that a rapid phosphodiester bond formation step is flanked by slow conformational changes for both correct and incorrect base pair formation. These experiments in combination with computer simulations indicate that the first conformational change occurs with rate constants of 75 and 20 s−1; rapid phosphodiester bond formation occurs with a Keq of 2.2 and 1.7, and the second conformational change occurs with rate constants of 2.1 and 0.5 s−1, for correct and incorrect base pair formation, respectively. The presence of a mispair does not induce the polymerase to adopt a low catalytic conformation. pol ν catalyzes both correct and mispair formation with high catalytic efficiency.  相似文献   

16.
Citrulline formation by both human neuronal nitric-oxide synthase (nNOS) and mouse macrophage inducible NOS was inhibited by the hydrogen sulfide (H2S) donor Na2S with IC50 values of ∼2.4·10−5 and ∼7.9·10−5 m, respectively, whereas human endothelial NOS was hardly affected at all. Inhibition of nNOS was not affected by the concentrations of l-arginine (Arg), NADPH, FAD, FMN, tetrahydrobiopterin (BH4), and calmodulin, indicating that H2S does not interfere with substrate or cofactor binding. The IC50 decreased to ∼1.5·10−5 m at pH 6.0 and increased to ∼8.3·10−5 m at pH 8.0. Preincubation of concentrated nNOS with H2S under turnover conditions decreased activity after dilution by ∼70%, suggesting irreversible inhibition. However, when calmodulin was omitted during preincubation, activity was not affected, suggesting that irreversible inhibition requires both H2S and NO. Likewise, NADPH oxidation was inhibited with an IC50 of ∼1.9·10−5 m in the presence of Arg and BH4 but exhibited much higher IC50 values (∼1.0–6.1·10−4 m) when Arg and/or BH4 was omitted. Moreover, the relatively weak inhibition of nNOS by Na2S in the absence of Arg and/or BH4 was markedly potentiated by the NO donor 1-(hydroxy-NNO-azoxy)-l-proline, disodium salt (IC50 ∼ 1.3–2.0·10−5 m). These results suggest that nNOS and inducible NOS but not endothelial NOS are irreversibly inhibited by H2S/NO at modest concentrations of H2S in a reaction that may allow feedback inhibition of NO production under conditions of excessive NO/H2S formation.  相似文献   

17.
18.
Alginate lyases are enzymes that degrade alginate through β-elimination of the glycosidic bond into smaller oligomers. We investigated the alginate lyases from Vibrio splendidus 12B01, a marine bacterioplankton species that can grow on alginate as its sole carbon source. We identified, purified, and characterized four polysaccharide lyase family 7 alginates lyases, AlyA, AlyB, AlyD, and AlyE, from V. splendidus 12B01. The four lyases were found to have optimal activity between pH 7.5 and 8.5 and at 20 to 25°C, consistent with their use in a marine environment. AlyA, AlyB, AlyD, and AlyE were found to exhibit a turnover number (kcat) for alginate of 0.60 ± 0.02 s−1, 3.7 ± 0.3 s−1, 4.5 ± 0.5 s−1, and 7.1 ± 0.2 s−1, respectively. The Km values of AlyA, AlyB, AlyD, and AlyE toward alginate were 36 ± 7 μM, 22 ± 5 μM, 60 ± 2 μM, and 123 ± 6 μM, respectively. AlyA and AlyB were found principally to cleave the β-1,4 bonds between β-d-mannuronate and α-l-guluronate and subunits; AlyD and AlyE were found to principally cleave the α-1,4 bonds involving α-l-guluronate subunits. The four alginate lyases degrade alginate into longer chains of oligomers.  相似文献   

19.
The mono-ADP-ribosyltransferase (mART) toxins are contributing factors to a number of human diseases, including cholera, diphtheria, traveler''s diarrhea, and whooping cough. VahC is a cytotoxic, actin-targeting mART from Aeromonas hydrophila PPD134/91. This bacterium is implicated primarily in diseases among freshwater fish species but also contributes to gastrointestinal and extraintestinal infections in humans. VahC was shown to ADP-ribosylate Arg-177 of actin, and the kinetic parameters were Km(NAD+) = 6 μm, Km(actin) = 24 μm, and kcat = 22 s−1. VahC activity caused depolymerization of actin filaments, which induced caspase-mediated apoptosis in HeLa Tet-Off cells. Alanine-scanning mutagenesis of predicted catalytic residues showed the predicted loss of in vitro mART activity and cytotoxicity. Bioinformatic and kinetic analysis also identified three residues in the active site loop that were critical for the catalytic mechanism. A 1.9 Å crystal structure supported the proposed roles of these residues and their conserved nature among toxin homologues. Several small molecules were characterized as inhibitors of in vitro VahC mART activity and suramin was the best inhibitor (IC50 = 20 μm). Inhibitor activity was also characterized against two other actin-targeting mART toxins. Notably, these inhibitors represent the first report of broad spectrum inhibition of actin-targeting mART toxins.  相似文献   

20.
1. The `30s' and `50s' ribosomes from ribonuclease-active (Escherichia coli B) and -inactive (Pseudomonas fluorescens and Escherichia coli MRE600) bacteria have been studied in the ultracentrifuge. Charge anomalies were largely overcome by using sodium chloride–magnesium chloride solution, I 0·16, made 0–50mm with respect to Mg2+. 2. Differentiation of enzymic and physical breakdown at Mg2+ concentrations less than 5mm was made by comparing the properties of E. coli B and P. fluorescens ribosomes. 3. Ribonuclease-active ribosomes alone showed a transformation of `50s' into 40–43s components. This was combined with the release of a small amount of `5s' material which may be covalently bound soluble RNA. Other transformations of the `50s' into 34–37s components were observed in both ribonuclease-active and -inactive ribosomes at 1·0–2·5mm-Mg2+, and also with E. coli MRE600 when EDTA (0·2mm) was added to a solution in 0·16m-sodium chloride. 4. Degradation of ribonuclease-active E. coli B ribosomes at Mg2+ concentration 0·25mm or less was coincident with the formation of 16s and 21s ribonucleoprotein in P. fluorescens, and this suggested that complete dissociation of RNA from protein was not an essential prelude to breakdown of the RNA by the enzyme. 5. As high Cs+/Mg2+ ratios cause ribosomal degradation great care is necessary in the interpretation of equilibrium-density-gradient experiments in which high concentrations of caesium chloride or similar salts are used. 6. The importance of the RNA moiety in understanding the response of ribosomes to their ionic environment 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