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 31 毫秒
1.
Ten gut and ten vaginalLactobacillus strains were investigated for their ability to bind type I collagen (Cn-I) and four selected gut lactobacilli were investigated for their binding to other extracellular matrix (ECM) molecules. Immobilized Cn-I (100 mg/L) in wells of microtitre plates was bound by all 10 autoaggregating vaginal strains and by 3 strains of gut lactobacilli from piglets in the range ofA 570 readings 0.114–1.806.L. acidophilus strain SV31 was much more adherent than the rest of strains. All four gut lactobacilli tested for binding to other ECM molecules displayed good binding to porcine fibronectin and heparin and some of them bound weakly to fetuin and porcine mucin. No binding of these strains was observed to bovine mucin, bovine fibrinogen and bovine lactoferrin.  相似文献   

2.
Summary The aerobic growth and metabolism of eleven homofermentative and three heterofermentative Lactobacillus strains, three Leuconostoc strains, two Brochothrix thermosphacta strains and two Carnobacterium strains were studied in batch cultures at pH 6.0 and 25°C on a complex substrate containing 10.0 g glucose per litre. All strains, except Carnobacterium divergens 69, grew well aerobically. An oxygen consumption was registered for 18 of the strains—the exceptions being Lactobacillus alimentarius DSM 20249T, Lactobacillus farciminis DSM 20284T and Lactobacillus sharpeae DSM 20505T. The homofermentative lactobacilli showed a maximal oxygen consumption during the stationary growth phase and this was coupled with a low final viable count. Leuconostoc strains, heterofermentative lactobacilli, Brochothrix thermosphacta and Carnobacterium strains showed a maximal oxygen consumption during the exponential growth phase together with a high final viable count. The maximum specific growth rate varied from 0.19 to 0.54 h-1 while the growth yield varied from 19 to 86 g dry weight per mol glucose consumed. In general, homofermentative lactobacilli produced dl-lactic acid, acetic acid and acetoin. The three heterofermentative lactobacilli produced dl-lactic acid and acetic acid, two strains also produced ethanol Leuconostoc spp. formed d-lactic acid, acetic acid, and ethanol. B. thermosphacta produced acetoin, acetic acid, formic acid, isobutyric acid and isovaleric acid but no lactic acid. Carnobacterium produced l-lactic acid, acetic acid and acetoin. All strains accumulated hydrogen peroxide except L. alimentarius DSM 20249T, Carnobacterium piscicola 3 and B. thermosphacta.née Blickstad  相似文献   

3.
Abstract Toxic shock syndrome toxin-1 (TSST-1) producing strains of Staphylococcus aureus isolated from 18 patients with toxic shock syndrome (TSS) and from 56 patients with other diagnoses were compared for capacity to interact with various serum and connective tissue proteins. TSS associated isolates showed significantly stronger binding of Type-I collagen (Cn-I) and Cn-II than non-TSS strains, in a particle agglutination assay (PAA) as well as in 125I labelled Cn uptake experiments. 125I Cn-IV binding, was similar between the two groups, whereas in PAA, a stronger interaction was observed for non-TSS than TSS associated strains. The median binding of 125I Cn to TSS-associated strains were 52.2 (Cn-I), 30.6 (Cn-II) and 20.0 (Cn-IV) compared to 20.0 (Cn-I), 14.4 (Cn-II) and 24.4 (Cn-IV) values of non-TSS strains. A saturation with 125I Cn-I and Cn-II binding was established for TSS (30 min) and non-TSS (15 min) strains. 125I Cn-IV binding reached a saturation in 10 min and 90 min with TSS and non-TSS strains respectively. Finally, the binding profiles of TSS associated and non-TSS strains to fibronectin, fibrinogen, laminin and IgG did not differ in both PAA and radioisotope assays. In scanning electron microscopy, cells of TSS associated strains bound to the reprecipitated native Cn-I fibrils. In contrast, most cells of non-TSS strains were localized to the distal end or were trapped between the Cn fibrils. The stronger interaction with Cn-I and II in particular, shown by TSS associated strains, might enhance submucosal localization, thereby facilitating entry of toxins into the blood and establishment of TSS.  相似文献   

4.
Toxic shock syndrome toxin-1 (TSST-1) producing strains of Staphylococcus aureus isolated from 18 patients with toxic shock syndrome (TSS) and from 56 patients with other diagnoses were compared for capacity to interact with various serum and connective tissue proteins. TSS associated isolates showed significantly stronger binding of Type-I collagen (Cn-I) and Cn-II than non-TSS strains, in a particle agglutination assay (PAA) as well as in 125I labelled Cn uptake experiments. 125I Cn-IV binding, was similar between the two groups, whereas in PAA, a stronger interaction was observed for non-TSS than TSS associated strains. The median binding of 125I Cn to TSS-associated strains were 52.2 (Cn-I), 30.6 (Cn-II) and 20.0 (Cn-IV) compared to 20.0 (Cn-I), 14.4 (Cn-II) and 24.4 (Cn-IV) values of non-TSS strains. A saturation with 125I Cn-I and Cn-II binding was established for TSS (30 min) and non-TSS (15 min) strains. 125I Cn-IV binding reached a saturation in 10 min and 90 min with TSS and non-TSS strains respectively. Finally, the binding profiles of TSS associated and non-TSS strains to fibronectin, fibrinogen, laminin and IgG did not differ in both PAA and radioisotope assays. In scanning electron microscopy, cells of TSS associated strains bound to the reprecipitated native Cn-I fibrils. In contrast, most cells of non-TSS strains were localized to the distal end or were trapped between the Cn fibrils.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Summary The specific binding of [3H]corticosterone to hepatocytes is a nonsaturable, reversible and temperature-dependent process. The binding to liver purified plasma membrane fraction is also specific, reversible and temperature dependent but it is saturable. Two types of independent and equivalent binding sites have been determined from hepatocytes. One of them has high affinity and low binding capacity (K D=8.8nm andB max=1477 fmol/mg protein) and the other one has low affinity and high binding capacity (K D=91nm andB max=9015 fmol/mg). In plasma membrane only one type of binding site has been characterized (K D=11.2nm andB max=1982 fmol/mg). As it can be deduced from displacement data obtained in hepatocytes and plasma membrane the high affinity binding sites are different from the glucocorticoid, progesterone nuclear receptors and the Na+,K+-ATPase digitalis receptor. Probably it is of the same nature that the one determinate for [3H]cortisol and [3H]corticosterone in mouse liver plasma membrane. Beta-and alpha-adrenergic antagonists as propranolol and phentolamine did not affect [3H]corticosterone binding to hepatocytes and plasma membranes; therefore, these binding sites are independent of adrenergic receptors. The binding sites in hepatocytes and plasma membranes are not exclusive for corticosterone but other steroids are also bound with very different affinities.  相似文献   

6.
Summary. To mutant ddY/DAO mice lacking D-amino-acid oxidase activity and normal ddY/DAO+ mice, five D-amino acids (D-Asp, D-Ser, D-Ala, D-Leu and D-Pro) were orally administered for two weeks, and the D-amino acid levels were examined in seven brain regions. The levels of D-Asp markedly increased in the pituitary and pineal glands in both strains. In the ddY/DAO+ mice, the levels of the other D-amino acids did not significantly change in most of the brain regions. While in the ddY/DAO mice the levels of D-Ser significantly increased in most of the brain regions except for the cerebrum and hippocampus. The levels of D-Ala and D-Leu increased in all regions but the levels of D-Pro did not significantly change. The same five D-amino acids were intravenously injected into Wistar rats and the D-amino acid levels in their brains were examined for 60 min after the administration. The levels of D-Asp markedly increased in the pineal gland 3 min after the administration, while the levels of D-Ser, D-Ala, and D-Pro increased both in the pineal and pituitary glands, the levels of D-Leu increased in all brain regions. These results are useful for the elucidation of the origins and regulation of D-amino acids in the mammalian body.  相似文献   

7.
Brush border membrane vesicles (BBMV) enriched in sucrase, maltase and alkaline phosphatase, and impoverished in Na+-K+-ATPase, were isolated from proximal and distal intestine of the gilthead sea bream (Sparus aurata) by a MgCl2 precipitation method. Vesicles were suitable for the study of the characteristics of D-glucose apical transport. Only one D-glucose carrier was found in vesicles from each intestinal segment. In both cases, the D-glucose transport system was sodium-dependent, phlorizin-sensitive, significantly inhibited by D-glucose, D-galactose, α-methyl-D-glucose, 3-O-methyl-D-glucose and 2-deoxy-D-glucose, and showed stereospecificity. Apparent affinity constants of D-glucose transport (Kt) were 0.24 ± 0.03 mM in proximal and 0.18 ± 0.03 mM in distal intestine. Maximal rate of influx (Jmax) was 47.3 ± 2.2 pmols. mg−1 protein for proximal and 27.3 ± 3.6 pmols. mg−1 protein for distal intestine. Specific phlorizin binding and relative abundance of an anti-SGLT1 reactive protein were significantly higher in proximal than in distal BBMV. These results suggest the presence of the same D-glucose transporter along the intestine, with a higher density in the proximal portion. This transporter is compatible with the sodium-dependent D-glucose carrier described for other fish and with the SGLT1 of higher vertebrates.This revised version was published online in June 2005 with a corrected cover date.  相似文献   

8.
We studied the effects of H2O/D2O substitution on the permeation and gating of the large conductance Ca2+-activated K+ channels inChara gymnophylla droplet membrane using the patchclamp technique. The selectivity sequence of the channel was: K+>Rb+≫Li+, Na+, Cs+ and Cl. The conductance of this channel in symmetric 100mm KCl was found to be 130 pS. The single channel conductance was decreased by 15% in D2O as compared to H2O. The blockade of channel conductance by cytosolic Ca2+ weakened in D2O as a result of a decrease in zero voltage Ca2+ binding affinity by a factor of 1.4. Voltage-dependent channel gating was affected by D2O primarily due to the change in Ca2+ binding to the channel during the activation step. The Hill coefficient for Ca2+ binding was 3 in D2O and around 1 in H2O. The values of the Ca2+ binding constant in the open channel conformation were 0.6 and 6 μm in H2O and D2O, respectively, while the binding in the closed conformation was much less affected by D2O. The H2O/D2O substitution did not produce a significant change in the slope of channel voltage dependence but caused a shift as large as 60 mV with 1mm internal Ca2+.  相似文献   

9.
The peptide subunit pentapeptide H-L-Ala-D-Glu(L-Lys-D-Ala-D-Ala-OH)-NH2 of peptidoglycan was localized in the cell walls of several Gram-positive bacteria employing the indirect immunoferritin technique. Specific antibodies to the D-alanyl-D-alanine moiety of non-crosslinked peptide subunit pentapeptide were raised in rabbits by immunization with synthetic immunogen albumin-(CH2CO-Gly-L-Ala-L-Ala-D-Ala-D-Ala-OH)39. Specificity of these antibodies for the peptide subunit pentapeptide and not for the peptide subunit tetrapeptide was corroborated in a Farr-type radio-active hapten binding assay. Specificity of labelling with ferritin was established by immunoelectron microscopic controls, and by the excellent correlation between specific labelling of cells with ferritin and the particular peptidoglycan primary structure of bacterial strains investigated. Cells of Lactobacillus gasseri, Streptococcus pyogenes and Staphylococcus aureus revealing non-crosslinked peptide subunit pentapeptides in their peptidoglycans could specifically be labelled. Lactobacillus acidophilus and Bacillus subtilis, on the contrary, missing such pentapeptides, failed in labelling.The implication of this method to possibly localize the points of attack of penicillin or cycloserine is discussed.Abbreviations used meso-A2pm meso-diaminopimelic acid - DSM Deutsche Sammlung für Mikroorganismen, Göttingen, FRG This paper is dedicated to Professor Gerhart Drews on the occasion of his 60th birthday  相似文献   

10.
Thirteen bacteria were isolated on D-4-hydroxyphenylglycine as sole carbon and energy source. Seven strains transaminated only the D-enantiomer while the other six isolates transaminated both enantiomers of 4-hydroxyphenylglycine. One of the six strains utilizing both enantiomers was characterized as a Pseudomonas putida. This strain, MW27, employed two enantioselective transaminases, to catalyze the initial step in the metabolism of DL-4-hydroxyphenylglycine. The product of the transamination, 4-hydroxyphenylglyoxylate, was further metabolized via 4-hydroxybenzaldehyde and 4-hydroxybenzoate to protocatechuate. Preliminary results indicate that both transaminases are co-ordinately synthesized together with the 4-hydroxyphenylglyoxylate decarboxylase and the NADP+-dependent 4-hydroxybenzaldehyde dehydrogenase.  相似文献   

11.
The activity of β-D-galactosidase was studied in 13 strains of lactobacilli (groupsStreptobacterium, Thermobacterium andBetabacterium). Using 2-nitrophenyl galactopyranoside as substrate, the enzyme activity varied with the strain. The values found in theThermobacterium group were superior to those in theStreptobacterium group. The optimum pH for the species belonging to theThermobacterium group was uniform, in contrast to the ph for those from theStreptobacterium which varied according to the species. The optimum temperature was quite uniform within each group and higher in theStreptobacterium. Lactose acted as a competitive inhibitor. MgCl2 protected the enzyme from thermal denaturation. The calcium ions inhibited the activity in all cases. The behaviour of the protectors of the SH groups varied according to the strain. 6-Phospho-β-D-galactosidase activity was also determined, levels lower than β-D-galactosidase were found, except inLactobacillus plantarum ATCC 8014 and 14917.  相似文献   

12.
α-Chymotrypsin-catalyzed peptide synthesis was carried out between an N-protected D-amino acid ester and an L-amino acid amide (acyl donor, 10 mM; acyl acceptor, 50 mM; enzyme, 2 mg ml−1; pH 8). By using a highly reactive carbamoylmethyl (Cam) ester as acyl donor, the D-amino acid was incorporated into the N-terminus of the resulting dipeptide amide. N-Protected dipeptide amides bearing D-amino acids such as D-Phe, D-Leu and D-Ala at their N-terminus were synthesized in high yields (up to 80%) in 1–3 h.  相似文献   

13.
The xylitol dehydrogenase-encoding Arxula adeninivorans AXDH gene was isolated and characterized. The gene includes a coding sequence of 1107 bp encoding a putative 368 amino acid protein of 40.3 kDa. The identity of the gene was confirmed by a high degree of homology of the derived amino acid sequence to that of xylitol dehydrogenases from different sources. The gene activity was regulated by carbon source. In media supplemented with xylitol, D-sorbitol and D-xylose induction of the AXDH gene and intracellular accumulation of the encoded xylitol dehydrogenase was observed. This activation pattern was confirmed by analysis of AXDH promoter – GFP gene fusions. The enzyme characteristics were analysed from isolates of native strains as well as from those of recombinant strains expressing the AXDH gene under control of the strong A. adeninivorans-derived TEF1 promoter. For both proteins, a molecular mass of ca. 80 kDa was determined corresponding to a dimeric structure, an optimum pH at 7.5 and a temperature optimum at 35 °C. The enzyme oxidizes polyols like xylitol and D-sorbitol whereas the reduction reaction is preferred when providing D-xylulose, D-ribulose and L-sorbose as substrates. Enzyme activity exclusively depends on NAD+ or NADH as coenzymes.  相似文献   

14.
Candida shehatae cells pre-grown on D-xylose simultaneously consumed mixtures of D-xylose and D-glucose, under both non-growing (anoxic) and actively growing conditions (aerobic), to produce ethanol. The rate of D-glucose consumption was independent of the D-xylose concentration for cells induced on D-xylose. However, the D-xylose consumption rate was approximately three times lower than the D-glucose consumption rate at a 50% D-glucose: 50% D-xylose mixture. Repression was not observed (substrate utilization rates were approximately equal) when the percentage of D-glucose and D-xylose was changed to 22% and 78%, respectively. In fermentations with actively growing cells (50% glucose and D-xylose), ethanol yields from D-xylose increased, the % D-xylose utilized increased, and the xylitol yield was significantly reduced in the presence of D-glucose, compared to anoxic fermentations (YETOH,xylose = 0.2–0.40 g g−1, 75–100%, and Yxylitol = 0–0.2 g g−1 compared to YETOH,xylose = 0.15 g g−1, 56%, Yxylitol = 0.51 g g−1, respectively). To increase ethanol levels and reduce process time, fed-batch fermentations were performed in a single stage reactor employing two phases: (1) rapid aerobic growth on D-xylose (μ = 0.32 h−1) to high cell densities; (2) D-glucose addition and anaerobic conditions to produce ethanol (YETOH,xylose = 0.23 g g−1). The process generated high cell densities, 2 × 109 cells ml−1, and produced 45–50 g L−1 ethanol within 50 h from a mixture of D-glucose and D-xylose (compared to 30 g L−1 in 80 h in the best batch process). The two-phase process minimized loss of cell viability, increased D-xylose utilization, reduced process time, and increased final ethanol levels compared to the batch process. Received 23 February 1998/ Accepted in revised form 15 July 1998  相似文献   

15.
N-Acyl-D-glutamate amidohydrolase (D-AGase) was inhibited by 94 % when 1 mol/l N-acetyl-DL- glutamate was used as a substrate. The addition of 1 mM Co2+ stabilized D-AGase. Moreover, the substrate inhibition was weakened to 88% with the addition of 0.4 mM Co2+ to the reaction mixture. Although D-AGase is a zinc-metalloenzyme, the addition of Zn2+ from 0.01 to 10 mM did not increase the D-glutamic acid production in the saturated substrate. Under optimal conditions, 0.38 M D-glutamic acid was obtained from N-acyl-DL-glutamate with 100% of the theoretical yield after 48 h.  相似文献   

16.
Mutant strains of Anabaena variabilis which are resistant to the tryptophan analogue, 6-fluorotryptophan, liberated a wide range of amino acids although none liberated tryptophan in detectable quantities. Four strains (FT-7, FT-8, FT-9, FT-10) produced predominantly alanine together with small amounts of phenylalamine and tyrosine, strain FT-2 liberated mainly phenylalanine and tyrosine and strain FT-6 liberated mainly glutamate, NH 4 + and several unidentified ninhydrin-positive compounds. Two forms of 3-deoxy-D-arbinoheptulosonate 7-phosphate (DAHP) synthase were identified in the parent strain, a tyrosine-sensitive form and a phenylalanine-sensitive form. In strains FT-2 and FT-6 the phenylalanine-sensitive enzyme was not detected and in strain FT-7 it was apparently deregulated with respect to inhibition by phenylalanine. No deregulation of anthranilate synthase was observed but mutant strains were found to have higher specific activities of this enzyme than the parent strain.Abbreviations chla chlorophyll a - 6-FT 6-fluorotryptophan - DAHP 3-deoxy-D-arabinoheptulosonate 7-phosphate - PEP phosphoenolpyruvate  相似文献   

17.
Summary R. meliloti strains 107-1, 111 and 152 were adapted to D-methionine in three ways: a) consecutive transfer in the presence of increasing amounts of D-methionine, b) alternate transfer between D- and L-methionine-containing media followed by final cultivation in the presence of each isomer, c) alternate transfer between D-methionine and medium 79 followed by cultivation in medium 79 or in D-methionine-medium. At the end of the experiment efficiency of the strains was ascertained by a plant test.Strain 111 lost efficiency when it was adapted consecutively to 0.125% D-methionine or alternated between D-methionine and either L-methionine or medium 79-Strain 107-1 sucessively adapted to D-methionine lost efficiency within 16 weeks. On adaptation to D-methionine alternated with L-methionine, efficiency was retained in L-methionine medium and lost in D-methionine medium. On alternate adaptation between D-methionine and medium 79, strain 107-1 lost efficiency in the D-methionine-medium but not in medium 79. Efficiency of strain 152 was lost by adaptation to 0.125% D-methionine, but it was maintained on the alternate adaptation between D-methionine and L-methionine or medium 79.  相似文献   

18.
In mammals, D-fructose transport takes place across the brush-border membrane of the small intestine through GLUT5, a member of the facilitative glucose transporter family. In the present paper, we describe and characterize for the first time the apical transport of D-fructose in chicken intestine. Brush-border membrane vesicles (BBMV) were obtained from jejunum of 5- to 6-wk-old chickens. D-Fructose uptake by BBMV from chicken jejunum comprises a saturable component and a simple diffusion process. The maximal rate of transport (Vmax) for D-fructose was 2.49 nmol·(mg prot)–1·s–1, the Michaelis constant (Km) was 29 mM, and the diffusion constant (Kd) was 25 nl·(mg prot)–1·s–1. The apical transport of D-fructose was Na+-independent, phlorizin-, phloretin-, and cytochalasin B-insensitive, and did not show cis-inhibition by D-glucose or D-galactose. These properties, together with the detection of specific GLUT5 mRNA, indicate the presence of a low-affinity high-capacity GLUT5-type carrier in the chicken jejunum, responsible for the entry of D-fructose across the brush-border membrane of enterocytes.  相似文献   

19.
Ascitic fluids from patients with various types of cancer were screened for the CA 19-9 and CA 125 tumor-associated antigenic activities. Two fluids exhibiting the highest activities were tested for their binding to various lectin-Sepharose columns resulting in both being bound best to wheat germ agglutinin (WGA) Sepharose. The WGA column eluate of one fluid was further chromatographed by HPLC and three peaks were obtained with approximate molecular weights of 3.65 MDa, 664 kDa and 330 kDa, of which only the largest fraction contained the CA 19-9 activity. The fluids were also fractionated on a Sephacryl S-400 column with most of the activity being present in or near the void volume.Monoclonal antibodies were used to demonstrate that the purified glycoproteins also contained the blood group A determinant, the four Lewis determinants Lea, Leb, Lex and Ley, and the sialylated-Lex determinant, while other antibody analyses failed to detect other blood group and/or carbohydrate sequence determinants. Some of the blood group expressions could be separated from the CA 19-9 and CA 125 active glycoproteins by adsorption with various lectins other than the WGA.Abbreviations used NeuAc N-acetyl-D-neuraminic acid - Gal galactose,D-galactopyranose - Fuc fucose,L-fucopyranose - GlcNAc N-acetyl-D-glucosamine - GalNAc N-acetyl-D-galactosamine - WGA wheat germ agglutinin - PBS phosphate buffered saline  相似文献   

20.
ETB, ether treated bacteria, from E. coli and other Gram-negative strains, contain in a cell-free system all enzymes necessary for murein biosynthesis. Starting with a variety of combinations of peptidoglycan precursors, high yields of sodium dodecylsulfate (SDS, 4%) insoluble murein or murein like material were synthesized. The amount of newly synthesized SDS insoluble material (NSM) was dependent upon the growing phase at which cells had been harvested for preparation of ETB. This data may provide some insight into the regulation of peptidoglycan biosynthesis.Starting from early peptidoglycan precursors, the cell-free synthesis of NSM was inhibited by specific inhibitors of murein synthesis, such as D-cycloserine, D-fluoroalanine, 2-amino-ethylphosphonate, analogues of D-alanyl-D-alanine and -lactam antibiotics at appropriate concentrations. Some D-alanyl-D-alanine analogues and 4-chlorodiaminopimelic acid were incorporated into NSM in place of their corresponding natural substrates.Abbreviations ETB ether treated bacteria (E. coli) - NSM newly synthesized SDS insoluble material - SDS sodium dodecylsulfate - UDP-MAG UDP-MurNAc-dipeptide, UDP-N-acetylmuramoyl-L-alanyl-D-glutamate - UDP-MAGD UDP-MurNAc-tripeptide, UDP-N-acetylmuramoyl-L-alanyl-D-glutamyl-meso-2,6-diaminopimelate - UDP-MAGDAA UDP-MurNAc-pentapeptide, UDP-N-acetylmuramoyl-L-alanyl-D-glutamyl-meso-2,6-diaminopimeloyl-D-alanyl-D-alanine - GINAc N-Acetylglucosamine Definitions Murein highly cross-linked bagshaped peptidoglycan (Weidel and Pelzer 1964)  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