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.
A chitinase was purified from the stomach of a fish, the silver croaker Pennahia argentatus, by ammonium sulfate fractionation and column chromatography using Chitopearl Basic BL-03, CM-Toyopearl 650S, and Butyl-Toyopearl 650S. The molecular mass and isoelectric point were estimated at 42 kDa and 6.7, respectively. The N-terminal amino acid sequence showed a high level of homology with family 18 chitinases. The optimum pH of silver croaker chitinase toward p-nitrophenyl N-acetylchitobioside (pNp-(GlcNAc)2) and colloidal chitin were observed to be pH 2.5 and 4.0, respectively, while chitinase activity increased about 1.5- to 3-fold with the presence of NaCl. N-Acetylchitooligosaccharide ((GlcNAc)n, n = 2–6) hydrolysis products and their anomer formation ratios were analyzed by HPLC using a TSK-GEL Amide-80 column. Since the silver croaker chitinase hydrolyzed (GlcNAc)4–6 and produced (GlcNAc)2–4, it was judged to be an endo-type chitinase. Meanwhile, an increase in β-anomers was recognized in the hydrolysis products, the same as with family 18 chitinases. This enzyme hydrolyzed (GlcNAc)5 to produce (GlcNAc)2 (79.2%) and (GlcNAc)3 (20.8%). Chitinase activity towards various substrates in the order pNp-(GlcNAc)n (n = 2–4) was pNp-(GlcNAc)2 >> pNp-(GlcNAc)4 > pNp-(GlcNAc)3. From these results, silver croaker chitinase was judged to be an enzyme that preferentially hydrolyzes the 2nd glycosidic link from the non-reducing end of (GlcNAc)n. The chitinase also showed wide substrate specificity for degrading α-chitin of shrimp and crab shell and β-chitin of squid pen. This coincides well with the feeding habit of the silver croaker, which feeds mainly on these animals.  相似文献   

2.
The in vitro production of chitinases and β-1,3-glucanases by Stachybotrys elegans, a mycoparasite of Rhizoctonia solani, was examined under various culture conditions, such as carbon and nitrogen sources, pH, and incubation period. Production of both enzymes was influenced by the carbon source incorporated into the medium and was stimulated by acidic pH and NaNO3. The activity of both enzymes was very low in culture filtrates from cells grown on glucose and sucrose compared with that detected on chitin (for chitinases) and cell wall fragments (for β-1,3-glucanases). Protein electrophoresis revealed that, depending on the carbon source used, different isoforms of chitinases and β-1,3-glucanases were detected. S. elegans culture filtrates, possessing β-1,3-glucanase and chitinase activities, were capable of degrading R. solani mycelium.  相似文献   

3.
The most abundant N-glycan in plants is the paucimannosidic N-glycan with core β1,2-xylose and α1,3-fucose residues (Man3XylFuc(GlcNAc)2). Here, we report a mechanism in Arabidopsis thaliana that efficiently produces the largest N-glycan in plants. Genetic and biochemical evidence indicates that the addition of the 6-arm β1,2-GlcNAc residue by N-acetylglucosaminyltransferase II (GnTII) is less effective than additions of the core β1,2-xylose and α1,3-fucose residues by XylT, FucTA, and FucTB in Arabidopsis. Furthermore, analysis of gnt2 mutant and 35S:GnTII transgenic plants shows that the addition of the 6-arm non-reducing GlcNAc residue to the common N-glycan acceptor GlcNAcMan3(GlcNAc)2 inhibits additions of the core β1,2-xylose and α1,3-fucose residues. Our findings indicate that plants limit the rate of the addition of the 6-arm GlcNAc residue to the common N-glycan acceptor as a mechanism to facilitate formation of the prevalent N-glycans with Man3XylFuc(GlcNAc)2 and (GlcNAc)2Man3XylFuc(GlcNAc)2 structures.  相似文献   

4.
There is emerging evidence that chitinases have additional functions beyond degrading environmental chitin, such as involvement in innate and acquired immune responses, tissue remodeling, fibrosis, and serving as virulence factors of bacterial pathogens. We have recently shown that both the human chitotriosidase and a chitinase from Salmonella enterica serovar Typhimurium hydrolyze LacNAc from Galβ1–4GlcNAcβ-tetramethylrhodamine (LacNAc-TMR (Galβ1–4GlcNAcβ(CH2)8CONH(CH2)2NHCO-TMR)), a fluorescently labeled model substrate for glycans found in mammals. In this study we have examined the binding affinities of the Salmonella chitinase by carbohydrate microarray screening and found that it binds to a range of compounds, including five that contain LacNAc structures. We have further examined the hydrolytic specificity of this enzyme and chitinases from Sodalis glossinidius and Polysphondylium pallidum, which are phylogenetically related to the Salmonella chitinase, as well as unrelated chitinases from Listeria monocytogenes using the fluorescently labeled substrate analogs LacdiNAc-TMR (GalNAcβ1–4GlcNAcβ-TMR), LacNAc-TMR, and LacNAcβ1–6LacNAcβ-TMR. We found that all chitinases examined hydrolyzed LacdiNAc from the TMR aglycone to various degrees, whereas they were less active toward LacNAc-TMR conjugates. LacdiNAc is found in the mammalian glycome and is a common motif in invertebrate glycans. This substrate specificity was evident for chitinases of different phylogenetic origins. Three of the chitinases also hydrolyzed the β1–6 bond in LacNAcβ1–6LacNAcβ-TMR, an activity that is of potential importance in relation to mammalian glycans. The enzymatic affinities for these mammalian-like structures suggest additional functional roles of chitinases beyond chitin hydrolysis.  相似文献   

5.
This paper evaluates CHCl3 and CH3OH extracts of the stem bark, branches and leaves of Drimys brasiliensis and drimane sesquiterpenes isolated from the stem bark against strains of Leishmania amazonensis and Leishmania braziliensis promastigotes and Plasmodium falciparum trophozoites. All of the extracts and compounds were tested in cell lines in comparison with reference standards and cell viability was determined by the XTT method. The CHCl3 and CH3OH extracts from the stem bark and branches yielded promising results against two strains of Leishmania, with 50% inhibitory concentrations (IC50 ) values ranging from 39-100 µg/mL. The CHCl3 extract of the stem bark returned IC50 values of 39 and 40.6 µg/mL for L. amazonensis and L. braziliensis, respectively. The drimanes were relatively effective: 1-β-(p-coumaroyloxy)-polygodial produced IC50 values of 5.55 and 2.52 µM for L. amazonensis and L. braziliensis, respectively, compared with 1-β-(p-methoxycinnamoyl)-polygodial, which produced respective IC50 values of 15.85 and 17.80 µM. The CHCl3 extract demonstrated activity (IC50 of 3.0 µg/mL) against P. falciparum. The IC50 values of 1-β-(p-cumaroyloxyl)-polygodial and 1-β-(p-methoxycinnamoyl)-polygodial were 1.01 and 4.87 µM, respectively, for the trophozoite strain. Therefore, the results suggest that D. brasiliensis is a promising plant from which to obtain new and effective antiparasitic agents.  相似文献   

6.
Chitin, a homopolymer of β1,4-linked N-acetylglucosamine (GlcNAc) residues, is a key component of the cell walls of fungi and the exoskeletons of arthropods. Chitin synthases transfer GlcNAc from UDP-GlcNAc to preexisting chitin chains in reactions that are typically stimulated by free GlcNAc. The effect of GlcNAc was probed by using a yeast strain expressing a single chitin synthase, Chs2, by examining formation of chitin oligosaccharides (COs) and insoluble chitin, and by replacing GlcNAc with 2-acylamido analogues of GlcNAc. Synthesis of COs was strongly dependent on inclusion of GlcNAc in chitin synthase incubations, and N,N′-diacetylchitobiose (GlcNAc2) was the major reaction product. Formation of both COs and insoluble chitin was also stimulated by GlcNAc2 and by N-propanoyl-, N-butanoyl-, and N-glycolylglucosamine. MALDI analyses of the COs made in the presence of 2-acylamido analogues of GlcNAc showed they that contained a single GlcNAc analogue and one or more additional GlcNAc residues. These results indicate that Chs2 can use certain 2-acylamido analogues of GlcNAc, and likely free GlcNAc and GlcNAc2 as well, as GlcNAc acceptors in a UDP-GlcNAc-dependent glycosyltransfer reaction. Further, formation of modified disaccharides indicates that CSs can transfer single GlcNAc residues.  相似文献   

7.
To isolate chitinases and β-1,3-glucanases from the intercellular space of oats (Avena sativa L.), primary leaves were infiltrated with buffer and subjected to gentle centrifugation to obtain intercellular washing fluid (IWF). Approximately 5% of the chitinase and 10% of the β-1,3-glucanase activity of the whole leaf were released. Only small amounts (0.01-0.03%) of the intracellular marker malate-dehydrogenase were released into the IWF during infiltration. Activities of chitinase and β-1,3-glucanase in the IWF and in the leaf extract were compared by different chromatographic methods. On Sephadex G-75, chitinase appeared as a single peak (Mr 29.8 kD) both in IWF and homogenate. β-1,3-Glucanase, however, showed two peaks in the IWF (Mr 52 and 31.3 kD), whereas the elution pattern of the homogenate showed only one major peak at 22 kD. Chromatofocusing indicated that the IWF contained four chitinases and five β-1,3-glucanases. The elution pattern of the homogenate and IWF were similar with regard to the elution pH, but the peak intensities were distinctly different. Our results demonstrate that extracellular β-1,3-glucanases are different from those located intracellularly. Extracellular and intracellular chitinases do not differ in molecular properties, except for one isozyme which seems to be confined to the extracellular space. We suggest that both enzymes might play a special role in pathogenesis during fungal infection.  相似文献   

8.
Seven endochitinases (EC 3.2.1.14) (relative molecular masses 23000–28000 and isoelectric points 10.3–10.4) were purified from nonembryogenic Citrus sinensis L. Osbeck cv. Valencia callus tissue. The basic chitinase/lysozyme from this tissue (BCLVC) exhibited lysozyme, chitinase and chitosanase activities and was determined to be a class III chitinase. While BCLVC acted as a lysozyme at pH 4.5 and low ionic strength (0.03) it acted as a chitinase/chitosanase at high ionic strengths (0.2) with a pH optimum of ca. 5. The lysozyme activity of BCLVC was inhibited by histamine, imidazole, histidine and the N-acetyl-d-glucosamine oligosaccharide (GlcNAc)3. The basic chitinase from cv. Valencia callus, BCVC-2, had an N-terminal amino acid sequence similar to tomato and tobacco AP24 proteins. The sequences of the other five chitinases were N-terminal blocked. Whereas BCLVC was capable of hydrolyzing 13.8–100% acetylated chitosans and (GlcNAc)4–6 oligosaccharides, BCVC-2 hydrolyzed only 100% acetylated chitosan, and the remaining enzymes expressed varying degrees of hydrolytic capabilities. Experiments with (GlcNAc)2–6 suggest that BCLVC hydrolysis occurs in largely tetrasaccharide units whereas hydrolysis by the other chitinases occurs in disaccharide units. Cross-reactivities of the purified proteins with antibodies for a potato leaf chitinase (AbPLC), BCLVC, BCVC-3, and tomato AP24 indicate that these are separate and distinct proteins.Mention of a trademark, warranty, propriety, or vendor does not constitute a guarantee by the U.S. Department of Agriculture and does not imply its approval to the exclusion of other products or vendors that may also be suitable.Abbreviations Ab antibody - BCLVC basic chitinase/lysozyme cv. Valencia callus - BCVC basic chitinase cv. Valencia callus - CE capillary electrophoresis - CM-chitin-RBV carboxymethyl-chitin-remazol brilliant violet - GlcNAc N-acetyl-d-glucosamine - HEWL hen egg-white lysozyme - Mr relativemolecular mass - pI isoelectric point - PLC potato leaf chitinase - PR pathogenesis-related - SEC size exclusion chromatography We thank Mr. M. Burkhart, Ms. T.-T. Ho, and Ms. M. Doherty for their valuable technical assistance. A portion of the funding for this work was made available from the Citrus Production Research Marketing Order by the Division of Marketing and Development, Florida Department of Agriculture and Consumer Services, Bob Crawford, Commissioner.  相似文献   

9.
The protein components of the 2-nitrotoluene (2NT) and nitrobenzene dioxygenase enzyme systems from Acidovorax sp. strain JS42 and Comamonas sp. strain JS765, respectively, were purified and characterized. These enzymes catalyze the initial step in the degradation of 2-nitrotoluene and nitrobenzene. The identical shared reductase and ferredoxin components were monomers of 35 and 11.5 kDa, respectively. The reductase component contained 1.86 g-atoms iron, 2.01 g-atoms sulfur, and one molecule of flavin adenine dinucleotide per monomer. Spectral properties of the reductase indicated the presence of a plant-type [2Fe-2S] center and a flavin. The reductase catalyzed the reduction of cytochrome c, ferricyanide, and 2,6-dichlorophenol indophenol. The ferredoxin contained 2.20 g-atoms iron and 1.99 g-atoms sulfur per monomer and had spectral properties indicative of a Rieske [2Fe-2S] center. The ferredoxin component could be effectively replaced by the ferredoxin from the Pseudomonas sp. strain NCIB 9816-4 naphthalene dioxygenase system but not by that from the Burkholderia sp. strain LB400 biphenyl or Pseudomonas putida F1 toluene dioxygenase system. The oxygenases from the 2-nitrotoluene and nitrobenzene dioxygenase systems each had spectral properties indicating the presence of a Rieske [2Fe-2S] center, and the subunit composition of each oxygenase was an α3β3 hexamer. The apparent Km of 2-nitrotoluene dioxygenase for 2NT was 20 μM, and that for naphthalene was 121 μM. The specificity constants were 7.0 μM−1 min−1 for 2NT and 1.2 μM−1 min−1 for naphthalene, indicating that the enzyme is more efficient with 2NT as a substrate. Diffraction-quality crystals of the two oxygenases were obtained.  相似文献   

10.
4-O-β-Di-N-acetylchitobiosyl moranoline (2) and 4-O-β-tri-N-acetylchitotriosyl moranoline (3) were produced by lysozyme-mediated transglycosylation from the substrates tetra-N-acetylchitotetraose, (GlcNAc)4, and moranoline, and the binding modes of 2 and 3 to hen egg white lysozyme (HEWL) was examined by inhibition kinetics, isothermal titration calorimetry (ITC), and x-ray crystallography. Compounds 2 and 3 specifically bound to HEWL, acting as competitive inhibitors with Ki values of 2.01 × 10−5 and 1.84 × 10−6 m, respectively. From ITC analysis, the binding of 3 was found to be driven by favorable enthalpy change (ΔHr°), which is similar to those obtained for 2 and (GlcNAc)4. However, the entropy loss (−TΔSr°) for the binding of 3 was smaller than those of 2 and (GlcNAc)4. Thus the binding of 3 was found to be more favorable than those of the others. Judging from the Kd value of 3 (760 nm), the compound appears to have the highest affinity among the lysozyme inhibitors identified to date. X-ray crystal structure of HEWL in a complex with 3 showed that compound 3 binds to subsites −4 to −1 and the moranoline moiety adopts an undistorted 4C1 chair conformation almost overlapping with the −1 sugar covalently bound to Asp-52 of HEWL (Vocadlo, Davies, G. J., Laine, R., and Withers, S. G. (2001) Nature 412, 835–838). From these results, we concluded that compound 3 serves as a transition-state analogue for lysozyme providing additional evidence supporting the covalent glycosyl-enzyme intermediate in the catalytic reaction.  相似文献   

11.
4-Methyl-5-nitrocatechol (4M5NC) monooxygenase (DntB) from Burkholderia sp. strain DNT catalyzes the second step of 2,4-dinitrotoluene degradation by converting 4M5NC to 2-hydroxy-5-methylquinone with the concomitant removal of the nitro group. DntB is a flavoprotein that has a very narrow substrate range. Here, error-prone PCR was used to create variant DntB M22L/L380I, which accepts the two new substrates 4-nitrophenol (4NP) and 3-methyl-4-nitrophenol (3M4NP). At 300 μM of 4NP, the initial rate of the variant expressing M22L/L380I enzyme (39 ± 6 nmol/min/mg protein) was 10-fold higher than that of the wild-type enzyme (4 ± 2 nmol/min/mg protein). The values of kcat/Km of the purified wild-type DntB enzyme and purified variant M22L/L380I were 40 and 450 (s−1 M−1), respectively, which corroborates that the variant M22L/L380I enzyme has 11-fold-higher efficiency than the wild-type enzyme for 4NP degradation. In addition, the variant M22L/L380I enzyme has fourfold-higher activity toward 3M4NP; at 300 μM, the initial nitrite release rate of M22L/L380I enzyme was 17 ± 4 nmol/min/mg protein, while that of the wild-type enzyme was 4.4 ± 0.7 nmol/min/mg protein. Saturation mutagenesis was also used to further investigate the role of the individual amino acid residues at positions M22, L380, and M22/L380 simultaneously. Mutagenesis at the individual positions M22L and L380I did not show appreciable enhancement in 4NP activity, which suggested that these two sites should be mutated together; simultaneous saturation mutagenesis led to the identification of the variant M22S/L380V, with 20% enhanced degradation of 4NP compared to the variant M22L/L380I. This is the first report of protein engineering for nitrite removal by a flavoprotein.  相似文献   

12.
An N-acetyl-β-d-hexosaminidase has been purified from primary wheat leaves (Triticum aestivum L.) by freeze-thawing, (NH4)2SO4 precipitation, methanol precipitation, gel filtration, cation exchange chromatography and affinity chromatography on concanavalin A-Sepharose. The activity of the purified preparations could be stabilised by addition of Triton X-100 and the enzyme was stored at -20°C without significant loss of activity. The enzyme hydrolysed pNP-β-d-GlcNAc (optimum pH 5.2, Km 0.29 mM, Vmax 2.56 μkat mg−1) and pNP-β-d-GalNAc (optimum pH 4.4, Km 0.27 mM, Vmax 2.50 μkat mg−1). Five major isozymes were identified, with isoelectric points in the range 5.13–5.36. All five isozymes possessed both N-acety-β-d-glucosaminidase and N-acetyl-β-d-galactosaminidase activity. Inhibition studies and mixed substrate analysis suggested that both substrates are catalysed by the same active site. Both activities were inhibited by GlcNAc, 2-acetamido-2-deoxygluconolactone, GalNAc and the ions of mercury, silver and copper. The Kis for inhibition of N-acetyl-β-d-glucosaminidase activity were: GlcNAc (15.3 mM) and GalNAc (3.4mM). For inhibition of N-acety-β-d-galactosaminidase activity the corresponding values were: GlcNAc (18.2 mM) and GalNac (2.5 mM). The enzyme was considerably less active at releasing pNP from pNP-β-d-(GlcNAc)2 and pNP-β-d-(GlcNAc)3 than from pNP-β-d-GlcNAc. The ability of the N-acetyl-β-d-hexosaminidase to relase GlcNAc from chitin oligomers (GlcNAc)2 (optimum pH 5.0) and (GlcNAc)3−6 (optimum pH 4.4) was also low. Analysis of the reaction products revealed that the initial products from the hydrolysis of (GlcNAc)n were predominantly (GlcNAc)n−1 and GlcNAc.  相似文献   

13.

Background

The enzymatic hydrolysis of α−mannosides is catalyzed by glycoside hydrolases (GH), termed α−mannosidases. These enzymes are found in different GH sequence–based families. Considerable research has probed the role of higher eukaryotic “GH38” α−mannosides that play a key role in the modification and diversification of hybrid N-glycans; processes with strong cellular links to cancer and autoimmune disease. The most extensively studied of these enzymes is the Drosophila GH38 α−mannosidase II, which has been shown to be a retaining α−mannosidase that targets both α−1,3 and α−1,6 mannosyl linkages, an activity that enables the enzyme to process GlcNAc(Man)5(GlcNAc)2 hybrid N-glycans to GlcNAc(Man)3(GlcNAc)2. Far less well understood is the observation that many bacterial species, predominantly but not exclusively pathogens and symbionts, also possess putative GH38 α−mannosidases whose activity and specificity is unknown.

Methodology/Principal Findings

Here we show that the Streptococcus pyogenes (M1 GAS SF370) GH38 enzyme (Spy1604; hereafter SpGH38) is an α−mannosidase with specificity for α−1,3 mannosidic linkages. The 3D X-ray structure of SpGH38, obtained in native form at 1.9 Å resolution and in complex with the inhibitor swainsonine (K i 18 µM) at 2.6 Å, reveals a canonical GH38 five-domain structure in which the catalytic “–1” subsite shows high similarity with the Drosophila enzyme, including the catalytic Zn2+ ion. In contrast, the “leaving group” subsites of SpGH38 display considerable differences to the higher eukaryotic GH38s; features that contribute to their apparent specificity.

Conclusions/Significance

Although the in vivo function of this streptococcal GH38 α−mannosidase remains unknown, it is shown to be an α−mannosidase active on N-glycans. SpGH38 lies on an operon that also contains the GH84 hexosaminidase (Spy1600) and an additional putative glycosidase. The activity of SpGH38, together with its genomic context, strongly hints at a function in the degradation of host N- or possibly O-glycans. The absence of any classical signal peptide further suggests that SpGH38 may be intracellular, perhaps functioning in the subsequent degradation of extracellular host glycans following their initial digestion by secreted glycosidases.  相似文献   

14.
Genes encoding three putative endopeptidases were identified from a draft-quality genome sequence of Lactobacillus helveticus CNRZ32 and designated pepO3, pepF, and pepE2. The ability of cell extracts from Escherichia coli DH5α derivatives expressing CNRZ32 endopeptidases PepE, PepE2, PepF, PepO, PepO2, and PepO3 to hydrolyze the model bitter peptides, β-casein (β-CN) (f193-209) and αS1-casein (αS1-CN) (f1-9), under cheese-ripening conditions (pH 5.1, 4% NaCl, and 10°C) was examined. CNRZ32 PepO3 was determined to be a functional paralog of PepO2 and hydrolyzed both peptides, while PepE and PepF had unique specificities towards αS1-CN (f1-9) and β-CN (f193-209), respectively. CNRZ32 PepE2 and PepO did not hydrolyze either peptide under these conditions. To demonstrate the utility of these peptidases in cheese, PepE, PepO2, and PepO3 were expressed in Lactococcus lactis, a common cheese starter, using a high-copy vector pTRKH2 and under the control of the pepO3 promoter. Cell extracts of L. lactis derivatives expressing these peptidases were used to hydrolyze β-CN (f193-209) and αS1-CN (f1-9) under cheese-ripening conditions in single-peptide reactions, in a defined peptide mix, and in Cheddar cheese serum. Peptides αS1-CN (f1-9), αS1-CN (f1-13), and αS1-CN (f1-16) were identified from Cheddar cheese serum and included in the defined peptide mix. Our results demonstrate that in all systems examined, PepO2 and PepO3 had the highest activity with β-CN (f193-209) and αS1-CN (f1-9). Cheese-derived peptides were observed to affect the activity of some of the enzymes examined, underscoring the importance of incorporating such peptides in model systems. These data indicate that L. helveticus CNRZ32 endopeptidases PepO2 and PepO3 are likely to play a key role in this strain's ability to reduce bitterness in cheese.  相似文献   

15.
Three strains of Steinernema feltiae Filipjev (All, Mexican, and Breton strains) and one of Heterorhabditis heliothidis (Khan, Brooks, and Hirschmann) were evaluated for their potential to control Colorado potato beetle (CPB), Leptinotarsa decemlineata (Say), larvae and pupae in the soil. In laboratory studies, H. heliothidis and S. feltiae (Mexican strain) produced the highest mortality (6 days posttreatment) of CPB when applied to the surface of a soil column containing mature CPB larvae 5 cm below. Mortality ranged from 80 to 90% at rates of 79-158 nematodes/cm². Similar results were seen in a field microplot study with all four nematodes; S. feltiae (Mexican strain) and H. heliothidis were most effective. Adult CPB emergence was reduced 86.5-100% after application of 31-93 H. heliothidis/cm² and 88.4-100% with 93-155 S. feltiae (Mexican strain)/cm². The All strain of S. feltiae was moderately effective (ca. 80% reduction at 93-155 nematodes/cm²), while the Breton strain was ineffective (< 40% reduction at 155 nematodes/cm²). In small plots of potatoes enclosed in field cages, application of H. heliothidis and S. feltiae (Mexican strain) at rates of 93-155 nematodes/cm² before larval CPB burial in the soil resulted in 66-77% reduction in adult CPB emergence. Soil applications of these nematodes show potential for biological control of CPB.  相似文献   

16.
α-Ketoglutarate-dependent (R)-dichlorprop dioxygenase (RdpA) and α-ketoglutarate-dependent (S)-dichlorprop dioxygenase (SdpA), which are involved in the degradation of phenoxyalkanoic acid herbicides in Sphingomonas herbicidovorans MH, were expressed and purified as His6-tagged fusion proteins from Escherichia coli BL21(DE3)(pLysS). RdpA and SdpA belong to subgroup II of the α-ketoglutarate-dependent dioxygenases and share the specific motif HXDX24TX131HX10R. Amino acids His-111, Asp-113, and His-270 and amino acids His-102, Asp-104, and His 257 comprise the 2-His-1-carboxylate facial triads and were predicted to be involved in iron binding in RdpA and SdpA, respectively. RdpA exclusively transformed the (R) enantiomers of mecoprop [2-(4-chloro-2-methylphenoxy)propanoic acid] and dichlorprop [2-(2,4-dichlorophenoxy)propanoic acid], whereas SdpA was specific for the (S) enantiomers. The apparent Km values were 99 μM for (R)-mecoprop, 164 μM for (R)-dichlorprop, and 3 μM for α-ketoglutarate for RdpA and 132 μM for (S)-mecoprop, 495 μM for (S)-dichlorprop, and 20 μM for α-ketoglutarate for SdpA. Both enzymes had high apparent Km values for oxygen; these values were 159 μM for SdpA and >230 μM for RdpA, whose activity was linearly dependent on oxygen at the concentration range measured. Both enzymes had narrow cosubstrate specificity; only 2-oxoadipate was able to replace α-ketoglutarate, and the rates were substantially diminished. Ferrous iron was necessary for activity of the enzymes, and other divalent cations could not replace it. Although the results of growth experiments suggest that strain MH harbors a specific 2,4-dichlorophenoxyacetic acid-converting enzyme, tfdA-, tfdAα-, or cadAB-like genes were not discovered in a screening analysis in which heterologous hybridization and PCR were used.  相似文献   

17.
Two Salmonella typhimurium strains, which could be used as sources for the leucine biosynthetic intermediates α- and β-isopropylmalate were constructed by a series of P22-mediated transductions. One strain, JK527 [flr-19 leuA2010 Δ(leuD-ara)798 fol-162], accumulated and excreted α-isopropylmalate, whereas the second strain, JK553 (flr-19 leuA2010 leuB698), accumulated and excreted α- and β-isopropylmalate. The yield of α-isopropylmalate isolated from the culture medium of JK527 was more than five times the amount obtained from a comparable volume of medium in which Neurospora crassa strain FLR92-1-216 (normally used as the source for α- and β-isopropylmalate) was grown. Not only was the yield greater, but S. typhimurium strains are much easier to handle and grow to saturation much faster than N. crassa strains. The combination of the two regulatory mutations flr-19, which results in constitutive expression of the leucine operon, and leuA2010, which renders the first leucine-specific biosynthetic enzyme insensitive to feedback inhibition by leucine, generated limitations in the production of valine and pantothenic acid. The efficient, irreversible, and unregulated conversion of α-ketoisovaleric acid into α-isopropylmalate (α-isopropylmalate synthetase Km for α-ketoisovaleric acid, 6 × 10−5 M) severely restricted the amount of α-ketoisovaleric acid available for conversion into valine and pantothenic acid (ketopantoate hydroxymethyltransferase Km for α-ketoisovaleric acid, 1.1 × 10−3 M; transaminase B Km for α-ketoisovaleric acid, 2 × 10−3 M).  相似文献   

18.
Streptomyces albovinaceus, Streptomyces caviscabies, Streptomyces griseus, Streptomyces setonii, and Streptomyces virginiae selected as antagonists of Moniliophthora (ex Crinipellis) perniciosa, the causal agent of cacao Witches’ broom, were examined in vitro to detect production of chitinases, β-1,3-glucanases, and cellulases. All the species produced chitinases, but not β-1,3-glucanases or cellulases, when grown on a liquid mineral medium containing glucose, colloidal chitin, or cell walls of M. perniciosa as a carbon source. There were no quantitative differences among species in the production of chitinase, however, the germination inhibition of basidiospores of M. perniciosa was higher when they were cultivated using glucose as a carbon source, followed by colloidal chitin and cell walls. All the species also produced hydroxymate type siderophores in similar quantities, and the quantity of siderophores did not correlate with the inhibition of basidiospore germination. The germination inhibition was more pronounced when S. albovinaceus, S. griseus, and S. virginiae were cultivated on iron-deficient medium, suggesting involvement of siderophores in the antagonism by these species of actinomycetes.  相似文献   

19.
Two chitinolytic fungal strains, Trichoderma aureoviride DY-59 and Rhizopus microsporus VS-9, were isolated from soil samples of Korea and Vietnam, respectively. DY-59 and VS-9 crude chitinases secreted by these fungi in the 0.5% swollen chitin culture medium had an optimal pH of 4 and the optimal temperatures of 40°C and 60°C, respectively. Enzymatic hydrolysis products from crab swollen chitin were N-acetyl-β-D-glucosamine (GlcNAc) by DY-59 chitinase, and GlcNAc and N, N′-diacetylchitobiose (GlcNAc)2 by VS-9 chitinases. The chitinases degraded the cell wall of Fusarium solani hyphae to produce oligosaccharides, among which GlcNAc, (GlcNAc)2, and pentamer (GlcNAc)5 were identified by high-pressure liquid chromatography. DY-59 and VS-9 chitinases inhibited F. solani microconidial germination by more than 70% and 60% at final protein concentrations of 5 and 27 μg mL−1, respectively, at 30°C for 20 h treatment.  相似文献   

20.
We studied nitrogen oxide production and consumption by methanotrophs Methylobacter luteus (group I), Methylosinus trichosporium OB3b (group II), and an isolate from a hardwood swamp soil, here identified by 16S ribosomal DNA sequencing as Methylobacter sp. strain T20 (group I). All could consume nitric oxide (nitrogen monoxide, NO), and produce small amounts of nitrous oxide (N2O). Only Methylobacter strain T20 produced large amounts of NO (>250 parts per million by volume [ppmv] in the headspace) at specific activities of up to 2.0 × 10−17 mol of NO cell−1 day−1, mostly after a culture became O2 limited. Production of NO by strain T20 occurred mostly in nitrate-containing medium under anaerobic or nearly anaerobic conditions, was inhibited by chlorate, tungstate, and O2, and required CH4. Denitrification (methanol-supported N2O production from nitrate in the presence of acetylene) could not be detected and thus did not appear to be involved in the production of NO. Furthermore, cd1 and Cu nitrite reductases, NO reductase, and N2O reductase could not be detected by PCR amplification of the nirS, nirK, norB, and nosZ genes, respectively. M. luteus and M. trichosporium produced some NO in ammonium-containing medium under aerobic conditions, likely as a result of methanotrophic nitrification and chemical decomposition of nitrite. For Methylobacter strain T20, arginine did not stimulate NO production under aerobiosis, suggesting that NO synthase was not involved. We conclude that strain T20 causes assimilatory reduction of nitrate to nitrite, which then decomposes chemically to NO. The production of NO by methanotrophs such as Methylobacter strain T20 could be of ecological significance in habitats near aerobic-anaerobic interfaces where fluctuating O2 and nitrate availability occur.  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