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.
We have purified a novel enzyme from eel white muscle which catalyzes the syntheses of imidazole dipeptides, such as carnosine (β-alanyl-l-histidine), anserine (β-alanyl-π-methyl-l-histidine), and balenine (ophidine; β-alanyl-τ-methyl-l-histidine), directly from their precursors. The enzyme was purified 1130-fold from eel muscle by a series of column chromatographies. Although eel muscle contains a large amount of carnosine and only trace amounts of anserine and balenine, the anserine synthesizing activity was by far the highest. From gel permeation chromatography, the molecular mass of the enzyme was calculated to be 275 kDa. SDS-PAGE of the purified enzyme represented a band around 43 kDa, suggesting that the native enzyme is a hexamer or heptamer. The optimal pH and temperature were around 9.5 and 60 °C, respectively. Km values for β-alanine and π-methyl-l-histidine were 44 and 89 mM, respectively. The enzyme was greatly activated by Zn2+ and inhibited by EDTA. The N-terminal amino acid sequence of 25 residues of the purified enzyme showed 52% amino acid identity to 38–62 residues of zebrafish haptoglobin precursor. The purified enzyme also exhibited hydrolytic activity against these imidazole dipeptides.  相似文献   

2.
Chemical modifications of substrate peptides are often necessary to monitor the hydrolysis of small bioactive peptides. We developed an electrospray ionization mass spectrometry (ESI–MS) assay for studying substrate distributions in reaction mixtures and determined steady-state kinetic parameters, the Michaelis–Menten constant (Km), and catalytic turnover rate (Vmax/[E]t) for three metallodipeptidases: two carnosinases (CN1 and CN2) from human and Dug1p from yeast. The turnover rate (Vmax/[E]t) of CN1 and CN2 determined at pH 8.0 (112.3 and 19.5 s−1, respectively) suggested that CN1 is approximately 6-fold more efficient. The turnover rate of Dug1p for Cys-Gly dipeptide at pH 8.0 was found to be slightly lower (73.8 s−1). In addition, we determined kinetic parameters of CN2 at pH 9.2 and found that the turnover rate was increased by 4-fold with no significant change in the Km. Kinetic parameters obtained by the ESI–MS method are consistent with results of a reverse-phase high-performance liquid chromatography (RP–HPLC)-based assay. Furthermore, we used tandem MS (MS/MS) analyses to characterize carnosine and measured its levels in CHO cell lines in a time-dependent manner. The ESI–MS method developed here obviates the need for substrate modification and provides a less laborious, accurate, and rapid assay for studying kinetic properties of dipeptidases in vitro as well as in vivo.  相似文献   

3.
Carnosine (β-alanyl-L-histidine) activates rabbit muscle phosphorylase a in the presence and absence of AMP and phosphorylase b in the presence of AMP in a biphasic manner with a maximal activation at about 50mM carnosine and with phosphorylase b showing a greater degree of activation than phosphorylase a. Anserine (β-alanyl-L-Nπ-methyl-histidine) activates phosphorylase a to a lesser extent than carnosine up to a concentration of 90mM, whereas with phosphorylase b a weak activation below 30mM and a concentration-dependent inhibition above this concentration occurs. These effects are specific for the dipeptides and are not shown by their constituent amino acids. Carnosine and anserine activate phosphorylase a in the presence of the allosteric inhibitors ATP, D-glucose and caffeine, and the inhibition of phosphorylase b by anserine is also observed in the presence of these inhibitors.  相似文献   

4.
A microchip electrophoresis (MCE) method with chemiluminescence (CL) detection was developed for the determination of carnosine-related peptides, including carnosine, homocarnosine, and anserine, in biological samples. A simple integrated MCE-CL system was built to perform the assays. The highly sensitive CL detection was achieved by means of the CL reaction between hydrogen peroxide and N-(4-aminobutyl)-N-ethylisoluminol-tagged peptides in the presence of adenine as a CL enhancer and Co2+ as a catalyst. Experimental conditions for analyte labeling, MCE separation, and CL detection were studied. MCE separation of the above-mentioned three peptides took less than 120 s. Detection limits (signal/noise ratio [S/N] = 3) of 3.0 × 10−8, 2.8 × 10−8, and 3.4 × 10−8 M were obtained for carnosine, anserine, and homocarnosine, respectively. The current MCE-CL method was applied for the determination of carnosine, anserine, and homocarnosine in human cerebrospinal fluid (CSF) and canine plasma. Homocarnosine was detected at the micromolar (μM) level in the CSF samples analyzed, whereas the levels of carnosine and anserine in these samples were below the detection limit of the assay. Interestingly, both carnosine and anserine were detected in the canine plasma samples, whereas homocarnosine was not.  相似文献   

5.
1. Muscle and brain from developing chick embryos, as well as from day-old chicks, rats, and ducks were analyzed for the histidine-containing dipeptides, anserine and carnosine. 2. Anserine was found in the brain of all species studied, whereas in muscle, anserine was found only in chicks. 3. At 15 days, the muscle of developing chick embryo contained 41 +/- 9 mumoles/100 g anserine while carnosine was present at a level of less than 3 mumoles/100 g. 4. In day-old chicks the anserine level in muscle was 100 +/- 35 mumoles/100 g while the carnosine level was 22.5 +/- 1 mumoles/100 g. 5. These findings cast doubt on earlier hypotheses relating anserine and carnosine to muscle activity.  相似文献   

6.
The phytase of Sporotrichum thermophile was purified to homogeneity using acetone precipitation followed by ion-exchange and gel-filtration column chromatography. The purified phytase is a homopentamer with a molecular mass of ~456 kDa and pI of 4.9. It is a glycoprotein with about 14% carbohydrate, and optimally active at pH 5.0 and 60 °C with a T1/2 of 16 h at 60 °C and 1.5 h at 80 °C. The activation energy of the enzyme reaction is 48.6 KJ mol?1 with a temperature quotient of 1.66, and it displayed broad substrate specificity. Mg2+ exhibited a slight stimulatory effect on the enzyme activity, while it was markedly inhibited by 2,3-butanedione suggesting a possible role of arginine in its catalysis. The chaotropic agents such as guanidinium hydrochloride, urea and potassium iodide strongly inhibited phytase activity. Inorganic phosphate inhibited enzyme activity beyond 3 mM. The maximum hydrolysis rate (Vmax) and apparent Michaelis–Menten constant (Km) for sodium phytate were 83 nmol mg?1 s?1 and 0.156 mM, respectively. The catalytic turnover number (Kcat) and catalytic efficiency (Kcat/Km) of phytase were 37.8 s?1 and 2.4 × 105 M?1 s?1, respectively. Based on the N-terminal and MALDI–LC–MS/MS identified amino acid sequences of the peptides, the enzyme did not show a significant homology with the known phytases.  相似文献   

7.
Phosphopentomutase (PPM) catalyzes the interconversion of α-d-(deoxy)-ribose 1-phosphate and α-d-(deoxy)-ribose 5-phosphate. We developed a coupled or uncoupled enzymatic assay with an enzyme nucleoside phosphorylase for determining PPM activities on d-ribose 5-phosphate at a broad temperature range from 30 to 90 °C. This assay not only is simple and highly sensitive but also does not require any costly special instrument. Via this technology, an open reading frame TM0167 from a thermophilic bacterium Thermotoga maritima putatively encoding PPM was cloned. The recombinant PPM was overexpressed in Escherichia coli Rosetta. This enzyme has the highest activity at 90 °C. MnCl2 (0.1 mM) and 50 μM α-d-glucose 1,6-bisphosphate are cofactors. The kinetic parameters of Km and kcat are 1.2 mM and 185 s−1 at 90 °C, respectively. The enzyme has a half-life time of up to 156 min at 90 °C. This enzyme is the most active and thermostable PPM reported to date.  相似文献   

8.
Vitamin B12 is destroyed by the addition of substantial amounts of vitamin C in the presence of copper. Effects of carnosine and anserine, natural water-soluble antioxidants, on the destruction of vitamin B12, were studied. Addition of carnosine (l0mM) effectively repressed the destruction of vitamin B12, but anserine had only weak inhibitory effects.  相似文献   

9.
Ensifer (Sinorhizobium) meliloti is a nitrogen-fixing α-proteobacterium able to biosynthesize the osmoprotectant glycine betaine from choline sulfate through a metabolic pathway that starts with the enzyme choline-O-sulfatase. This protein seems to be widely distributed in microorganisms and thought to play an important role in their sulfur metabolism. However, only crude extracts with choline sulfatase activity have been studied. In this work, Ensifer (Sinorhizobium) meliloti choline-O-sulfatase was obtained in a high degree of purity after expression in Escherichia coli. Gel filtration and dynamic light scattering experiments showed that the recombinant enzyme exists as a dimer in solution. Using calorimetry, its catalytic activity against its natural substrate, choline-O-sulfate, gave a kcat=2.7×10?1 s?1 and a KM=11.1 mM. For the synthetic substrates p-nitrophenyl sulfate and methylumbelliferyl sulfate, the kcat values were 3.5×10?2 s?1 and 4.3×10?2 s?1, with KM values of 75.8 and 11.8 mM respectively. The low catalytic activity of the recombinant sulfatase was due to the absence of the formylglycine post-translational modification in its active-site cysteine 54. Nevertheless, unmodified Ensifer (Sinorhizobium) meliloti choline-O-sulfatase is a multiple-turnover enzyme with remarkable catalytic efficiency.  相似文献   

10.
Dihydrodipicolinate synthase (DHDPS, E.C. 4.2.1.52), a validated antibiotic target, catalyses the first committed step in the lysine biosynthetic pathway: the condensation reaction between (S)-aspartate β-semialdehyde [(S)-ASA] and pyruvate via the formation of a Schiff base intermediate between pyruvate and the absolutely conserved active-site lysine. Escherichia coli DHDPS mutants K161A and K161R of the active-site lysine were characterised for the first time. Unexpectedly, the mutant enzymes were still catalytically active, albeit with a significant decrease in activity. The kcat values for DHDPS-K161A and DHDPS-K161R were 0.06 ± 0.02 s−1 and 0.16 ± 0.06 s−1 respectively, compared to 45 ± 3 s−1 for the wild-type enzyme. Remarkably, the KM values for pyruvate increased by only 3-fold for DHDPS-K161A and DHDPS-K161R (0.45 ± 0.04 mM and 0.57 ± 0.06 mM, compared to 0.15 ± 0.01 mM for the wild-type DHDPS), while the KM values for (S)-ASA remained the same for DHDPS-K161R (0.12 ± 0.01 mM) and increased by only 2-fold for DHDPS-K161A (0.23 ± 0.02 mM) and the Ki for lysine was unchanged. The X-ray crystal structures of DHDPS-K161A and DHDPS-K161R were solved at resolutions of 2.0 and 2.1 Å respectively and showed no changes in their secondary or tertiary structures when compared to the wild-type structure. The crystal structure of DHDPS-K161A with pyruvate bound at the active site was solved at a resolution of 2.3 Å and revealed a defined binding pocket for pyruvate that is thus not dependent upon lysine 161. Taken together with ITC and NMR data, it is concluded that although lysine 161 is important in the wild-type DHDPS-catalysed reaction, it is not absolutely essential for catalysis.  相似文献   

11.
We investigated the catalytic activity and inhibition of the β-class carbonic anhydrase (CA, EC 4.2.1.1) CahB1, from the relict cyanobacterium Coleofasciculus chthonoplastes (previously denominated Microcoleus chthonoplastes). The enzyme showed good activity as a catalyst for the CO2 hydration, with a kcat of 2.4 × 105 s−1 and a kcat/Km of 6.3 × 107 M−1 s−1. A range of inorganic anions and small molecules were investigated as inhibitors of CahB1. Perchlorate and tetrafluoroborate did not inhibit the enzyme (KIs >200 mM) whereas selenate and selenocyanide were ineffective inhibitors too, with KIs of 29.9–48.61 mM. The halides, pseudohalides, carbonate, bicarbonate, trithiocarbonate and a range of heavy metal ions-containing anions were submillimolar–millimolar inhibitors (KIs in the range of 0.15–0.90 mM). The best CahB1 inhibitors were N,N-diethyldithiocarbamate, sulfamate, sulfamide, phenylboronic acid and phenylarsonic acid, with KIs in the range of 8–75 μM, whereas acetazolamide inhibited the enzyme with a KI of 76 nM. This is the first kinetic and inhibition study of a cyanobacterial CA. As these enzymes are widespread in many cyanobacteria, being crucial for the carbon concentrating mechanism which assures substrate to RubisCO for the CO2 fixation by these organisms, a detailed kinetic/inhibition study may be essential for a better understanding of this superfamily of metalloenzymes and for potential biotechnological applications in biomimetic CO2 capture processes.  相似文献   

12.
An extracellular feruloyl esterase (PeFaeA) from the culture supernatant of Pleurotus eryngii was purified to homogeneity using cation exchange, hydrophobic interaction, and size exclusion chromatography. The length of the complete coding sequence of PeFaeA was determined to 1668 bp corresponding to a protein of 555 amino acids. The catalytic triad of Ser-Glu-His demonstrated the uniqueness of the enzyme compared to previously published FAEs. The purified PeFaeA was a monomer with an estimated molecular mass of 67 kDa. Maximum feruloyl esterase (FAE) activity was observed at pH 5.0 and 50 °C, respectively. Metal ions (5 mM), except Hg2+, had no significant influence on the enzyme activity. Substrate specificity profiling characterized the enzyme as a type A FAE preferring bulky natural substrates, such as feruloylated saccharides, rather than small synthetic ones. Km and kcat of the purified enzyme for methyl ferulate were 0.15 mM and 0.85 s−1. In the presence of 3 M NaCl activity of the enzyme increased by 28 %. PeFaeA alone released only little ferulic acid from destarched wheat bran (DSWB), whereas after addition of Trichoderma viride xylanase the concentration increased more than 20 fold.  相似文献   

13.
Glycine oxidase (GO) has great potential for use in biosensors, industrial catalysis and agricultural biotechnology. In this study, a novel GO (BliGO) from a marine bacteria Bacillus licheniformis was cloned and characterized. BliGO showed 62% similarity to the well-studied GO from Bacillus subtilis. The optimal activity of BliGO was observed at pH 8.5 and 40 °C. Interestingly, BliGO retained 60% of the maximum activity at 0 °C, suggesting it is a cold-adapted enzyme. The kinetic parameters on glyphosate (Km, kcat and kcat/Km) of BliGO were 11.22 mM, 0.08 s−1, and 0.01 mM−1 s−1, respectively. To improve the catalytic activity to glyphosate, the BliGO was engineered by directed evolution. With error-prone PCR and two rounds of DNA shuffling, the most evolved mutant SCF-4 was obtained from 45,000 colonies, which showed 7.1- and 8-fold increase of affinity (1.58 mM) and catalytic efficiency (0.08 mM−1 s−1) to glyphosate, respectively. In contrast, its activity to glycine (the natural substrate of GO) decreased by 113-fold. Structure modeling and site-directed mutation study indicated that Ser51 in SCF-4 involved in the binding of enzyme with glyphosate and played a crucial role in the improvement of catalytic efficiency.  相似文献   

14.
Cold-adapted esterases and lipases have been found to be dominant activities throughout the cold marine environment, indicating their importance in bacterial degradation of the organic matter. lip2 Gene from Psychrobacter sp. TA144, a micro-organism isolated from the Antarctic sea water, was cloned and over-expressed in Escherichia coli. The recombinant protein (PsyHSL) accumulated in the insoluble fraction from which it was recovered in active form, purified to homogeneity and deeply characterised. Temperature dependence of PsyHSL activity was typical of psychrophilic enzymes, with an optimal temperature of 35 °C at pH 8.0. The enzyme resulted to be active on pNP-esters of fatty acids with acyl chain length from C2 to C12 and the preferred substrate was pNP-pentanoate showing a kcat = 26.2 ± 0.1 s−1, KM = 0.122 ± 0.006 mM and a kcat/KM = 215 ± 11 mM−1 s−1. The enzyme was strongly inhibited by Hg2+, Zn2+, Cu2+, Fe3+, Mn2+ ions and it resulted to be activated in presence of methanol and acetonitrile, with calculated C50 values of 1.98 M and 0.92 M, respectively.  相似文献   

15.
A putative aldehyde reductase gene from Oceanospirillum sp. MED92 was overexpressed in Escherichia coli. The recombinant protein (OsAR) was characterized as a monomeric NADPH-dependent aldehyde reductase. The kinetic parameters Km and kcat of OsAR were 0.89 ± 0.08 mM and 11.07 ± 0.99 s−1 for benzaldehyde, 0.04 ± 0.01 mM and 6.05 ± 1.56 s−1 for NADPH, respectively. This enzyme exhibited high activity toward a variety of aromatic and aliphatic aldehydes, but no activity toward ketones. As such, it catalyzed the chemoselective reduction of aldehydes in the presence of ketones, as demonstrated by the reduction of 4-acetylbenzaldehyde or the mixture of hexanal and 2-nonanone, showing the application potential of this marine enzyme in such selective reduction of synthetic importance.  相似文献   

16.
Recently, we identified an allelic variant of human carnosinase 1 (CN1) that results in increased enzyme activity and is associated with susceptibility for diabetic nephropathy in humans. Investigations in diabetic (db/db) mice showed that carnosine ameliorates glucose metabolism effectively. We now investigated the renal carnosinase metabolism in db/db mice. Kidney CN1 activity increased with age and was significantly higher in diabetic mice compared to controls. Increased CN1 activity did not affect renal carnosine levels, but anserine concentrations were tenfold lower in db/db mice compared to controls (0.24±0.2 vs. 2.28±0.3 nmol/mg protein in controls; p<0.001). Homocarnosine concentrations in kidney tissue were low in both control and db/db mice (below 0.1 nmol/mg protein, p=n.s.). Carnosine treatment for 4 weeks substantially decreased renal CN1 activity in diabetic mice (0.32±0.3 in non-treated db/db vs. 0.05±0.05 μmol/mg/h in treated db/db mice; p<0.01) close to normal activities. Renal anserine concentrations increased significantly (0.24±0.2 in non-treated db/db vs. 5.7±1.2 μmol/mg/h in treated db/db mice; p<0.01), while carnosine concentrations remained unaltered (53±6.4 in non-treated vs. 61±15 nmol/mg protein in treated db/db mice; p=n.s.). Further, carnosine treatment halved proteinuria and reduced vascular permeability to one-fifth in db/db mice. In renal tissue of diabetic mice carnosinase activity is significantly increased and anserine concentrations are significantly reduced compared to controls. Carnosine treatment largely prevents the alterations of renal carnosine metabolism.  相似文献   

17.
We have cloned, purified and characterized the γ-carbonic anhydrase (CA, EC 4.2.1.1) present in the genome of the Antarctic bacterium Colwellia psychrerythraea, which is an obligate psychrophile. The enzyme shows a significant catalytic activity for the physiologic reaction of CO2 hydration to bicarbonate and protons, with the following kinetic parameters: kcat of 6.0 × 105 s−1 and a kcat/Km of 4.7 × 106 M−1 × s−1. This activity was inhibited by the sulfonamide CA inhibitor (CAI) acetazolamide, with a KI of 502 nM. A range of anions was also investigated for their inhibitory action against the new enzyme CpsCA. Perchlorate, tetrafluoroborate, fluoride and bromide were not inhibitory, whereas cyanate, thiocyanate, cyanide, hydrogensulfide, carbonate and bicarbonate showed KIs in the range of 1.4–4.4 mM. Diethyldithiocarbamate was a better inhibitor (KI of 0.58 mM) whereas sulfamide, sulfamate, phenylboronic acid and phenylarsonic acid were the most effective inhibitors detected, with KIs ranging between 8 and 38 μM. The present study may shed some more light regarding the role that γ-CAs play in the life cycle of psychrophilic bacteria as the Antarctic one investigated here.  相似文献   

18.
The glycolytic enzyme triose phosphate isomerase from Schistosoma mansoni is a potential target for drugs and vaccines. Molecular modelling of the enzyme predicted that a Ser-Ala-Asp motif which is believed to be a helminth-specific epitope is exposed. The enzyme is dimeric (as judged by gel filtration and cross-linking), resistant to proteolysis and highly stable to thermal denaturation (melting temperature of 82.0 °C). The steady-state kinetic parameters are high (Km for dihydroxyacetone phosphate is 0.51 mM; Km for glyceraldehyde 3-phosphate is 1.1 mM; kcat for dihydroxyacetone phosphate is 7800 s−1 and kcat for glyceraldehyde 3-phosphate is 6.9 s−1).  相似文献   

19.
In this study, a bioanode was developed by using layer-by-layer (LBL) assembly of sulfonated graphene (SG)/ferritin (Frt)/glucose oxidase (GOx). The SG/Frt biocomposite was used as an electron transfer elevator and mediator, respectively. Glucose oxidase (GOx) from Aspergillus niger was applied as a glucose oxidation biocatalyst. The electrocatalytic oxidation of glucose using GOx modified electrode increases with an increase in the concentration of glucose in the range of 10–50 mM. The electrochemical measurements of the electrode was carried out by using cyclic voltammetry (CV) at different scan rates (20–100 mV s−1) in 30 mM of glucose solution prepared in 0.3 M potassium ferrocyanide (K4Fe(CN)6) and linear sweep voltammetry (LSV). A saturation current density of 50 ± 2 mA cm−2 at a scan rate of 100 mV s−1 for the oxidation of 30 Mm glucose is achieved.  相似文献   

20.
We investigated the catalytic activity and inhibition of the δ-class carbonic anhydrase (CA, EC 4.2.1.1) from the marine diatom Thalassiosira weissflogii, TweCA. The enzyme, obtained by cloning the synthetic gene, was an efficient catalyst for the CO2 hydration, its physiological reaction, with a kcat of 1.3 × 105 s−1 and a kcat/KM of 3.3 × 107 M−1 s−1. A range of inorganic anions and small molecules were investigated as inhibitors of TweCA. Chloride and sulfate did not inhibit the enzyme (KIs >200 mM) whereas other halides and pseudohalides were submillimolar–millimolar inhibitors (KIs in the range of 0.93–8.3 mM). The best TweCA inhibitors were hydrogen sulfide, sulfamate, sulfamide, phenylboronic acid and phenylarsonic acid, with KIs in the range of 9–90 μM, whereas acetazolamide inhibited the enzyme with a KI of 83 nM. This is the first kinetic and inhibition study of a δ-class CA. However, these enzymes are widespread in the marine phytoplankton, being present in haptophytes, dinoflagellates, diatoms, and chlorophytic prasinophytes, contributing to the CO2 fixation by sea organisms. A phylogenetic analysis with all five genetic families of CAs showed that α- and δ-CAs are evolutionarily more related to each other with respect to the γ-CAs, although these three families clustered all together. On the contrary, the β- and ζ-CAs are also related to each other but phylogenetically much more distant from the α-, γ and δ-CA cluster. Thus, the study of δ-CAs is essential for better understanding this superfamily of metalloenzymes and their potential biotechnological applications in biomimetic CO2 capture processes, as these enzymes are part of the carbon concentrating mechanism used by many photosynthetic organisms.  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