首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   50篇
  免费   7篇
  2022年   1篇
  2019年   1篇
  2018年   5篇
  2017年   1篇
  2016年   4篇
  2015年   3篇
  2014年   3篇
  2013年   8篇
  2012年   2篇
  2011年   5篇
  2010年   5篇
  2009年   2篇
  2006年   3篇
  2005年   5篇
  1995年   2篇
  1993年   1篇
  1991年   1篇
  1983年   1篇
  1982年   2篇
  1980年   1篇
  1972年   1篇
排序方式: 共有57条查询结果,搜索用时 31 毫秒
21.
22.
23.
Shrub communities are important components of the subalpine vegetation. We focused on habitat conditions and stage structure (flowering and non-reproducing individuals) of stands of the endangered Salix lapponum in the Krkono?e Mts, Czech Republic. Habitat conditions were determined using soil sample analyses and Ellenberg indicator values (EIVs) derived from fine-scale (1?×?1 m) vegetation plots. The fine-scale plots were compared with coarse-scale relevés with the occurrence of S. lapponum acquired from the Czech National Phytosociological Database. We found that S. lapponum grows on nutrient-poor, acidic soils with high relative water contents, high amounts of total nitrogen, low amounts of phosphorus, moderate amounts of magnesium and low to moderate amounts of calcium. The overall proportion of non-reproducing individuals was 35.2%, but strong variations were observed among populations (0–100%). Co-occurring species and EIVs data indicated that flowering individuals are relatively more common in humid, nutrient-rich and warmer microhabitats than non-reproducing ones. Well-developed (“typical”) stands of S. lapponum with a dominance of flowering individuals occur along alpine springs, streams and in glacial cirques (= association Salicetum lapponum Zlatník 1928), but S. lapponum also grows along transitional mires and peat bogs and in subalpine grasslands. Coarse-scale relevés were similar to “typical” stands of S. lapponum at the fine scale but were differentiated from them by high frequencies of several, mainly broad-leaved forbs and herbs typical of moist to wet and fertile soils. Threats to S. lapponum and the future prospects of its stands in the Krkono?e Mts are briefly discussed.  相似文献   
24.

Proteins were obtained from effluent of a starch manufacture by using different isolation temperatures (40, 60, 80, and 100 °C). The proteins, remaining in effluent after treatment of potato juice at 80 and 100 °C differed significantly in composition and in structural stability as well as in trypsin inhibitory and antifungal activities in comparison with the variants of 40 and 60 °C. The protein samples of 80 °C exhibited the highest antifungal activity and its average value of IC50 against five strains of two Fusarium species was determined in average at 0.18 mg ml−1. The 80 °C protein samples consisted predominantly of low-molecular proteins (7–17 kDa) identified as potato tuber protease inhibitors I and II. Predominantly, protease inhibitors II were identified for the protein samples obtained by 100 °C and here we identified 7 spots in comparison with 12 identified for the 80 °C samples. Samples of 40 and 60 °C with low antifungal activities represent high variability of detected and identified proteins. We identified various representatives of aspartic, cysteine, and serine protease inhibitors in both types of samples. These samples also contained Kunitz-type protease inhibitors that were not found in the 80 and 100 °C samples which documented thermal unstableness of Kunitz-type protease inhibitors. Functional stability at high temperatures and antifungal activity of isolated potato protease inhibitors I and II support the potential of this fraction usage in food, feed, pharmaceutical, or agricultural industry and offer new products for starch manufactures. At the same time, utilization of the stable protein fraction of waste deproteinized potato water promotes exploitation of potato starch production resources.

  相似文献   
25.
26.
27.
This study focuses on two representatives of experimentally uncharacterized haloalkane dehalogenases from the subfamily HLD-III. We report biochemical characterization of the expression products of haloalkane dehalogenase genes drbA from Rhodopirellula baltica SH1 and dmbC from Mycobacterium bovis 5033/66. The DrbA and DmbC enzymes show highly oligomeric structures and very low activities with typical substrates of haloalkane dehalogenases.Haloalkane dehalogenases (EC 3.8.1.5.) acting on halogenated aliphatic hydrocarbons catalyze carbon-halogen bond cleavage, leading to an alcohol, a halide ion, and a proton as the reaction products (7). Haloalkane dehalogenases originating from various bacterial strains have potential for application in bioremediation technologies (4, 6, 22), construction of biosensors (2), decontamination of warfare agents (17), and synthesis of optically pure compounds (19). Recent evolutionary study of haloalkane dehalogenase sequences revealed the existence of three subfamilies, denoted HLD-I, HLD-II, and HLD-III (3). In contrast to subfamilies HLD-I and HLD-II, the subfamily HLD-III is currently lacking experimentally characterized proteins. We have therefore focused on the isolation and study of two selected representatives of the HLD-III subfamily, DrbA and DmbC.The drbA gene was amplified by PCR using the cosmid pircos.a3g10 originating from marine bacterium Rhodopirellula baltica SH1, and the dmbC gene was amplified from DNA originating from obligatory pathogen Mycobacterium bovis 5033/66. Six-histidine tails were added to the C termini of DrbA and DmbC in a cloning step, enabling single-step purification using Ni-nitrilotriacetic acid resin. Haloalkane dehalogenase DrbA was expressed under the T7 promoter and purified, with a resulting yield of 0.1 mg of protein per gram of cell mass. Haloalkane dehalogenase DmbC was obtained by expression in Mycobacterium smegmatis, with a yield of 0.07 mg of purified protein per gram of cell mass.The correct folding and secondary structures of the newly prepared enzymes were verified by circular dichroism (CD) spectroscopy. Far-UV CD spectra were recorded for DrbA and DmbC enzymes and other, related haloalkane dehalogenases. All enzymes tested exhibited CD spectra with two negative features at 208 and 222 nm and one positive peak at 195 nm, which are characteristic of α-helical content (Fig. (Fig.1).1). This suggested that both new enzymes, DrbA and DmbC, were folded correctly. However, DmbC exhibited more intense negative maxima which differed from other haloalkane dehalogenases in the θ222208 ratio. This finding indicated a slight variation in the arrangement of secondary structure elements of the DmbC enzyme. Thermally induced denaturations of DrbA and DmbC were tested in parallel. Both enzymes showed changes in ellipticity during increasing temperature. The melting temperatures calculated from these curves were 45.8 ± 0.4°C for DmbC and 39.4 ± 0.1°C for DrbA. The thermostability results obtained for DrbA and DmbC were in good agreement with the range of melting temperatures determined for other, related haloalkane dehalogenases.Open in a separate windowFIG. 1.Far-UV CD spectra of DrbA, DmbC, and seven different biochemically characterized haloalkane dehalogenases. Protein concentration used for far-UV CD spectrum measurement was 0.2 mg/ml.The sizes of the purified proteins were estimated by electrophoresis under native conditions conducted using a 10% polyacrylamide gel (Fig. (Fig.2).2). More precise determination of the sizes of DrbA and DmbC was achieved by gel filtration chromatography performed on Sephacryl S-500 HR (GE Healthcare, Uppsala, Sweden), calibrated with blue dextran 2000, thyroglobulin (669 kDa), ferritin (440 kDa), catalase (240 kDa), conalbumin (75 kDa), and ovalbumin (43 kDa) (Fig. (Fig.3A).3A). Both DrbA and DmbC were eluted from the column in the fraction prior to blue dextran, indicating that both enzymes form oligomeric complexes of a size larger than 2,000 kDa (Fig. 3B and C). The haloalkane dehalogenases which have been biochemically characterized so far form monomers, except for DbjA isolated from Bradyrhizobium japonicum USDA110 (21), which shows monomeric, dimeric, and tetrameric forms according to the pH of the buffer (R. Chaloupkova, submitted for publication).Open in a separate windowFIG. 2.Native protein electrophoresis of DrbA and DmbC. Lane 1, carbonic anhydrase (29 kDa); lane 2, ovalbumin (43 kDa); lane 3, bovine albumin (67 kDa); lane 4, conalbumin (75 kDa); lane 5, catalase (240 kDa); lane 6, ferritin (440 kDa); lane 7, DrbA; lane 8, DmbC.Open in a separate windowFIG. 3.Gel filtration chromatogram of DrbA and DmbC. (A) The following calibration kit samples (0.5 ml of a concentration of 2 mg/ml protein loaded) were analyzed using 50 mM Tris-HCl with 150 mM NaCl, pH 7.5, as elution buffer: blue dextran (line 1, 9.6-ml fraction), thyroglobulin (line 2, 15.95-ml fraction), ferritin (line 3, 16.78-ml fraction), ovalbumin (line 4, 18.55-ml fraction), and RNase A (line 5, 20.08-ml fraction). (B and C) Haloalkane dehalogenase DrbA eluted in the 9.03-ml fraction (B), and haloalkane dehalogenase DmbC in the 9.31-ml fraction (C).The substrate specificities of DrbA and DmbC were investigated with a set of 30 selected chlorinated, brominated, and iodinated hydrocarbons. Standardized specific activities related to 1-chlorobutane (summarized in Table Table1)1) were compared with the activity profiles of other haloalkane dehalogenases (Fig. (Fig.4).4). DrbA and DmbC displayed similar activity patterns, with catalytic activities approximately two orders of magnitude lower than those of other known haloalkane dehalogenases (1, 5, 8-11, 13-16, 18, 20, 23). HLD-III subfamily enzymes showed a restricted specificity range and a preference for iodinated short-chain hydrocarbons. Both phenomena may be related to the composition of the catalytic pentad Asp-His-Asp+Asn-Trp, which is unique to the members of the HLD-III subfamily (3). The preference for substrates carrying an iodine substituent can be related to a pair of halide-binding residues and their spatial arrangement with the catalytic triad. These residues make up the catalytic pentad, playing a critical role in substrate binding, formation of the transition states, and the reaction intermediates of the dehalogenation reaction (12).Open in a separate windowFIG. 4.Substrate specificity profiles of DrbA, DmbC, and seven different biochemically characterized haloalkane dehalogenases. Activities were determined using a consistent set of 30 halogenated substrates (see Table Table1).1). Data were standardized by dividing each value by the sum of all activities determined for individual enzymes in order to mask the differences in absolute activities. Specific activities (in μmol·s−1·mg−1) with 1-chlorobutane are 0.0003 (DrbA), 0.0001 (DmbC), 0.0003 (DatA), 0.0133 (DbjA), 0.0010 (DbeA), 0.0128 (DhaA), 0.0231 (LinB), 0.0171 (DmbA), and 0.0117 (DhlA).

TABLE 1.

Specific activities of haloalkane dehalogenases DrbA and DmbC toward a set of 30 halogenated hydrocarbonsa
SubstrateDrbA
DmbC
Sp act (nmol product·s−1· mg−1 protein)Relative activity (%)Sp act (nmol product·s−1· mg−1 protein)Relative activity (%)
1-Chlorobutane0.2911000.122100
1-Chlorohexane0.129440.122100
1-Bromobutane0.081281.2211,000
1-Bromohexane0.181620.977800
1-Iodopropane0.143492.1981,800
1-Iodobutane0.5061742.5642,100
1-Iodohexane0.095330.244200
1,2-DichloroethaneNANANANA
1,3-DichloropropaneNANA0.01210
1,5-DichloropentaneNANA0.06150
1,2-Dibromoethane0.098340.855700
1,3-DibromopropaneNANA5.0074,100
1-Bromo-3-chloropropane0.00101.4651,200
1,3-Diiodopropane0.3581236.7165,500
2-Iodobutane0.0289NANA
1,2-DichloropropaneNANANANA
1,2-Dibromopropane0.148510.244200
2-Bromo-1-chloropropane0.091310.488400
1,2,3-TrichloropropaneNANANANA
Bis-(2-chloroethyl) etherNANANANA
ChlorocyclohexaneNANANANA
Bromocyclohexane0.0269NANA
(1-Bromomethyl)-cyclohexaneNANA0.08973
1-Bromo-2-chloroethane0.167570.11191
ChlorocyclopentaneNANANANA
4-Bromobutyronitrile0.200690.444364
1,2,3-TribromopropaneNANA0.222182
3-Chloro-2-methyl propeneNANANANA
2,3-Dichloropropene0.27695NANA
1,2-Dibromo-3-chloropropane0.01030.04436
Open in a separate windowaNA, no activity detected.Substrates 1-iodobutane and 1,3-diiodopropane, identified as the best substrates for haloalkane dehalogenases DrbA and DmbC, were used for measuring the dependency of enzyme activity on temperature and for determination of the pH optima. DrbA exhibited the highest activity with 1-iodobutane at 50°C, although above this temperature, the enzyme rapidly became inactivated. DmbC showed the highest activity toward 1,3-diiodopropane at 40°C, which is similar to the temperature determined with the haloalkane dehalogenases DmbA and DmbB (45°C), isolated from the same species (10). Irrespective of the reaction temperature, DrbA showed the maximum activity at pH 9.15. DrbA kept 80% of its activity throughout a relatively wide range of pH values (pH 7.00 and 9.91) compared to DmbC, which showed a sharp maximum at pH 8.30. The Michaelis-Menten kinetics of DrbA and DmbC determined by isothermal titration microcalorimetry were investigated with 1-iodobutane, which is an iodinated analogue of 1-chlorobutane routinely used for characterization of haloalkane dehalogenases. The low magnitudes of the Michaelis constants (Km = 0.063 ± 0.003 mM for DrbA and 0.018 ± 0.001 mM for DmbC) suggest a high affinity of both enzymes for 1-iodobutane. The catalytic constants determined with 1-iodobutane (kcat = 0.128 ± 0.002 s−1 for DrbA and 0.0715 ± 0.0004 s−1 for DmbC) suggest that the low specific activities observed during substrate screening are not due to poor affinity but are instead due to a low conversion rate.The biochemical characteristics of purified DrbA and DmbC suggest that these proteins represent novel enzymes differing from previously characterized haloalkane dehalogenases by (i) their unique ability to form oligomers and (ii) low levels of dehalogenating activity with typical substrates of haloalkane dehalogenases. This study further illustrates how genome sequencing projects and phylogenetic analyses contribute to the identification of novel enzymes. Characterization of DrbA and DmbC, belonging to the subfamily HLD-III, partially filled a gap in the knowledge of the haloalkane dehalogenase family and provided an additional insight into evolutionary relationships among its members.  相似文献   
28.
The influence of the matrix solution, sample form and deposition technique on the quality MALDI-TOF mass spectra was examined and assessed with the aim to improve MALDI-TOF MS performance for the identification of microorganisms and to enable automatic spectra acquisition. It was observed that the use of matrix compounds ferulic and sinapinic acid may result in improved mass spectral features, in terms of signal resolution and S/N ratio, as compared to alpha-cyano-4-hydroxycinnamic acid, which was, on the other hand, found to be the only matrix compound that enabled fully automatic mass spectra acquisition. The robustness of the whole sample preparation procedure was then assessed on a set of 25 strains of four Acinetobacter species. Results showed reproducible detection of subtle mass spectral differences between strains belonging to the same species, although they do not confirm the possibility of reliable strain typing.  相似文献   
29.
The newly rediscovered specimen ofEopelobates from the uppermost Oligocène brown coal measures near Bonn, Germany, deposited since the last century in the collections of the Institut und Museum für Geologie und Paläontologie der Universität Tübingen, most probably belongs toE. anthracinus. Its small but fully ossified skeleton suggests that it is reasonable to keep the speciesE. anthracinus separate fromE. bayeri, although both species are closely related.  相似文献   
30.
β-Hexachlorocyclohexane (β-HCH) is the most recalcitrant among the α-, β-, γ-, and δ-isomers of HCH and causes serious environmental pollution problems. We demonstrate here that the haloalkane dehalogenase LinB, reported earlier to mediate the second step in the degradation of γ-HCH in Sphingomonas paucimobilis UT26, metabolizes β-HCH to produce 2,3,4,5,6-pentachlorocyclohexanol.  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