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.
Misfolding of proinsulin variants in the pancreatic β-cell, a monogenic cause of permanent neonatal-onset diabetes mellitus, provides a model for a disease of protein toxicity. A hot spot for such clinical mutations is found at position B8, conserved as glycine within the vertebrate insulin superfamily. We set out to investigate the molecular basis of the aberrant properties of a proinsulin clinical mutant in which residue GlyB8 is replaced by SerB8. Modular total chemical synthesis was used to prepare the wild-type [GlyB8]proinsulin molecule and three analogs: [d-AlaB8]proinsulin, [l-AlaB8]proinsulin, and the clinical mutant [l-SerB8]proinsulin. The protein diastereomer [d-AlaB8]proinsulin produced higher folding yields at all pH values compared with the wild-type proinsulin and the other two analogs, but showed only very weak binding to the insulin receptor. The clinical mutant [l-SerB8]proinsulin impaired folding at pH 7.5 even in the presence of protein-disulfide isomerase. Surprisingly, although [l-SerB8]proinsulin did not fold well under the physiological conditions investigated, once folded the [l-SerB8]proinsulin protein molecule bound to the insulin receptor more effectively than wild-type proinsulin. Such paradoxical gain of function (not pertinent in vivo due to impaired secretion of the mutant insulin) presumably reflects induced fit in the native mechanism of hormone-receptor engagement. This work provides insight into the molecular mechanism of a clinical mutation in the insulin gene associated with diabetes mellitus. These results dramatically illustrate the power of total protein synthesis, as enabled by modern chemical ligation methods, for the investigation of protein folding and misfolding.  相似文献   

2.
The ACTH receptor, known as the melanocortin-2 receptor (MC2R), plays an important role in regulating and maintaining adrenocortical function. MC2R is a subtype of the melanocortin receptor (MCR) family and has unique characteristics among MCRs. Endogenous ACTH is the only endogenous agonist for MC2R, whereas the melanocortin peptides α-, β-, and γ-melanocyte-stimulating hormone and ACTH are full agonists for all other MCRs. In this study, we examined the molecular basis of MC2R responsible for ligand selectivity using ACTH analogs and MC2R mutagenesis. Our results indicate that substitution of Phe7 with d-Phe or d-naphthylalanine (d-Nal(2′)) in ACTH(1–24) caused a significant decrease in ligand binding affinity and potency. Substitution of Phe7 with d-Nal(2′) in ACTH(1–24) did not switch the ligand from agonist to antagonist at MC2R, which was observed in MC3R and MC4R. Substitution of Phe7 with d-Phe7 in ACTH(1–17) resulted in the loss of ligand binding and activity. Molecular analysis of MC2R indicated that only mutation of the third transmembrane domain of MC2R resulted in a decrease in d-Phe ACTH binding affinity and potency. Our results suggest that Phe7 in ACTH plays an important role in ligand selectivity and that the third transmembrane domain of MC2R is crucial for ACTH selectivity and potency.  相似文献   

3.
A Mg2+-dependent, alkaline phosphatase has been isolated from mature pollen of Lilium longiflorum Thunb., cv. Ace and partially purified. It hydrolyzes 1l- and 1d-myo-inositol 1-phosphate, myo-inositol 2-phosphate, and β-glycerophosphate at rates decreasing in the order named. The affinity of the enzyme for 1l- and 1d-myo-inositol 1-phosphate is approximately 10-fold greater than its affinity for myo-inositol 2-phosphate. Little or no activity is found with phytate, d-glucose 6-phosphate, d-glucose 1-phosphate, d-fructose 1-phosphate, d-fructose 6-phosphate, d-mannose 6-phosphate, or p-nitrophenyl phosphate. 3-Phosphosphoglycerate is a weak competitive inhibitor. myo-Inositol does not inhibit the reaction. Optimal activity is obtained at pH 8.5 and requires the presence of Mg2+. At 4 millimolar, Co2+, Fe2+ or Mn2+ are less effective. Substantial inhibition is obtained with 0.25 molar Li+. With β-glycerophosphate as substrate the Km is 0.06 millimolar and the reaction remains linear at least 2 hours. In 0.1 molar Tris, β-glycerophosphate yields equivalent amounts of glycerol and inorganic phosphate, evidence that transphosphorylation does not occur.  相似文献   

4.
1. When NAD+ was present, cell extracts of Pseudomonas (A) grown with d-glucarate or galactarate converted 1mol. of either substrate into 1mol. each of 2-oxoglutarate and carbon dioxide; 70–80% of the gas originated from C-1 of the hexarate. 2. The enzyme system that liberated carbon dioxide from galactarate was inactive in air and was stabilized by galactarate or Fe2+ ions; the system that acted on d-glucarate was more stable and was stimulated by Mg2+ ions. 3. When NAD+ was not added, 2-oxoglutarate semialdehyde accumulated from either substrate. This compound was isolated as its bis-2,4-dinitrophenylhydrazone, and several properties of the derivative were compared with those of the chemically synthesized material. Methods were developed for the determination of 2-oxoglutarate semialdehyde. 4. Synthetic 2-oxoglutarate semialdehyde was converted into 2-oxoglutarate by an enzyme that required NAD+; the reaction rate with NADP+ was about one-sixth of that with NAD+. 5. For extracts of Pseudomonas (A) grown with d-glucarate or galactarate, or for those of Pseudomonas fragi grown with l-arabinose or d-xylose, specific activities of 2-oxoglutarate semialdehyde–NAD oxidoreductase were much higher than for extracts of the organisms grown with (+)-tartrate and d-glucose respectively. 6. Extracts of Pseudomonas fragi grown with l-arabinose or d-xylose converted l-arabonate or d-xylonate into 2-oxoglutarate when NAD+ was added to reaction mixtures and into 2-oxoglutarate semialdehyde when NAD+ was omitted.  相似文献   

5.
Enzymatic processes are useful for industrially important sugar production, and in vitro two-step isomerization has proven to be an efficient process in utilizing readily available sugar sources. A hypothetical uncharacterized protein encoded by ydaE of Bacillus licheniformis was found to have broad substrate specificities and has shown high catalytic efficiency on d-lyxose, suggesting that the enzyme is d-lyxose isomerase. Escherichia coli BL21 expressing the recombinant protein, of 19.5 kDa, showed higher activity at 40 to 45°C and pH 7.5 to 8.0 in the presence of 1.0 mM Mn2+. The apparent Km values for d-lyxose and d-mannose were 30.4 ± 0.7 mM and 26 ± 0.8 mM, respectively. The catalytic efficiency (kcat/Km) for lyxose (3.2 ± 0.1 mM−1 s−1) was higher than that for d-mannose (1.6 mM−1 s−1). The purified protein was applied to the bioproduction of d-lyxose and d-glucose from d-xylose and d-mannose, respectively, along with the thermostable xylose isomerase of Thermus thermophilus HB08. From an initial concentration of 10 mM d-lyxose and d-mannose, 3.7 mM and 3.8 mM d-lyxose and d-glucose, respectively, were produced by two-step isomerization. This two-step isomerization is an easy method for in vitro catalysis and can be applied to industrial production.  相似文献   

6.
Peptidoglycan hydrolases (PGHs) are responsible for bacterial cell lysis. Most PGHs have a modular structure comprising a catalytic domain and a cell wall-binding domain (CWBD). PGHs of bacteriophage origin, called endolysins, are involved in bacterial lysis at the end of the infection cycle. We have characterized two endolysins, Lc-Lys and Lc-Lys-2, identified in prophages present in the genome of Lactobacillus casei BL23. These two enzymes have different catalytic domains but similar putative C-terminal CWBDs. By analyzing purified peptidoglycan (PG) degradation products, we showed that Lc-Lys is an N-acetylmuramoyl-l-alanine amidase, whereas Lc-Lys-2 is a γ-d-glutamyl-l-lysyl endopeptidase. Remarkably, both lysins were able to lyse only Gram-positive bacterial strains that possess PG with d-Ala4d-Asx-l-Lys3 in their cross-bridge, such as Lactococcus casei, Lactococcus lactis, and Enterococcus faecium. By testing a panel of L. lactis cell wall mutants, we observed that Lc-Lys and Lc-Lys-2 were not able to lyse mutants with a modified PG cross-bridge, constituting d-Ala4l-Ala-(l-Ala/l-Ser)-l-Lys3; moreover, they do not lyse the L. lactis mutant containing only the nonamidated d-Asp cross-bridge, i.e. d-Ala4d-Asp-l-Lys3. In contrast, Lc-Lys could lyse the ampicillin-resistant E. faecium mutant with 3→3 l-Lys3-d-Asn-l-Lys3 bridges replacing the wild-type 4→3 d-Ala4-d-Asn-l-Lys3 bridges. We showed that the C-terminal CWBD of Lc-Lys binds PG containing mainly d-Asn but not PG with only the nonamidated d-Asp-containing cross-bridge, indicating that the CWBD confers to Lc-Lys its narrow specificity. In conclusion, the CWBD characterized in this study is a novel type of PG-binding domain targeting specifically the d-Asn interpeptide bridge of PG.  相似文献   

7.
d-Serine is a physiological co-agonist of the N-methyl-d-aspartate receptor. It regulates excitatory neurotransmission, which is important for higher brain functions in vertebrates. In mammalian brains, d-amino acid oxidase degrades d-serine. However, we have found recently that in chicken brains the oxidase is not expressed and instead a d-serine dehydratase degrades d-serine. The primary structure of the enzyme shows significant similarities to those of metal-activated d-threonine aldolases, which are fold-type III pyridoxal 5′-phosphate (PLP)-dependent enzymes, suggesting that it is a novel class of d-serine dehydratase. In the present study, we characterized the chicken enzyme biochemically and also by x-ray crystallography. The enzyme activity on d-serine decreased 20-fold by EDTA treatment and recovered nearly completely by the addition of Zn2+. None of the reaction products that would be expected from side reactions of the PLP-d-serine Schiff base were detected during the >6000 catalytic cycles of dehydration, indicating high reaction specificity. We have determined the first crystal structure of the d-serine dehydratase at 1.9 Å resolution. In the active site pocket, a zinc ion that coordinates His347 and Cys349 is located near the PLP-Lys45 Schiff base. A theoretical model of the enzyme-d-serine complex suggested that the hydroxyl group of d-serine directly coordinates the zinc ion, and that the ϵ-NH2 group of Lys45 is a short distance from the substrate Cα atom. The α-proton abstraction from d-serine by Lys45 and the elimination of the hydroxyl group seem to occur with the assistance of the zinc ion, resulting in the strict reaction specificity.  相似文献   

8.
The glycolate/glycerate transporter of spinach (Spinacia oleracea L.) chloroplast inner envelope membranes was solubilized by treatment of the membranes with sodium cholate. Mixtures of the cholate extracts and soy asolectin were subjected to gel filtration to remove the detergent. The reconstituted vesicles were frozen, thawed, and sonicated in a buffer that contained 10 millimolar d-glycerate and, usually, [3H]sucrose as an internal space indicator. The dilution of the vesicles into a medium that contained 0.4 millimolar [14C]d-glycerate resulted in a rapid accumulation of labeled glycerate, followed by a much slower loss of [14C]d-glycerate from the vesicles. This behavior is characteristic of counterflow. The accumulation of [14C]d-glycerate was strongly inhibited by HgCl2, which blocks glycolate/glycerate transport in intact chloroplasts. In the absence of proton ionophores, the extent of [14C]glycolate accumulation under similar conditions was much greater than that of [14C]d-glycerate. External glycolate inhibited d-glycerate counterflow and external d-glycerate inhibited glycolate counterflow. The external pH dependence of the efflux of [14C]d-glycerate accumulated in vesicles by counterflow and its inhibition by external l-mandelate are characteristics displayed by glycolate transport in intact chloroplasts. Partial purification of the transporter was achieved by glycerol gradient centrifugation. The solubilized glycolate and glycerate counterflow activities, assayed by reconstitution into vesicles, were found to sediment similarly.  相似文献   

9.
Fry SC  Northcote DH 《Plant physiology》1983,73(4):1055-1061
Cultured spinach (Spinacia oleracea L. cv Monstrous Viroflay) cells incorporated exogenous l-[3H]arabinose sequentially into β-l-arabinopyranose-1-phosphate, uridine diphospho-β-l-arabinopyranose, uridine diphospho-α-d-xylopyranose and (in some experiments) α-d-xylopyranose-1-phosphate. The amount of 3H in each of these compounds reached a plateau after a few minutes, and could be rapidly chased with nonradioactive l-arabinose, demonstrating rapid turnover. After a few minutes' lag, incorporation of 3H into the arabinofuranosyl, arabinopyranosyl, and xylopyranosyl residues of polysaccharides was linear with respect to time. The kinetics of labeling were compatible with UDP-β-l-arabinopyranose and UDP-α-d-xylopyranose being the immediate precursors of arabians (both the pyranose and the furanose residues) and xylans, respectively. No other radioactive nucleotides were formed; in particular, UDP-arabinofuranose was absent. There was no evidence for conversion of arabinopyranose to arabinofuranose within the polysaccharides, suggesting that this conversion occurs during polymer synthesis. The glycolipids detected showed too slow a turnover to be intermediates of pentosan synthesis.  相似文献   

10.
Saito K  Nick JA  Loewus FA 《Plant physiology》1990,94(3):1496-1500
d-[6-14C]Glucosone that had been prepared enzymically from d-[6-14C]glucose was used to compare relative efficiencies of these two sugars for l-ascorbic acid (AA) biosynthesis in detached bean (Phaseolus vulgaris L., cv California small white) apices and 4-week-old spinach (Spinacia oleracea L., cv Giant Noble) leaves. At tracer concentration, 14C from glucosone was utilized by spinach leaves for AA biosynthesis much more effectively than glucose. Carbon-14 from [6-14C]glucose underwent considerable redistribution during AA formation, whereas 14C from [6-14C]glucosone remained almost totally in carbon 6 of AA. In other experiments with spinach leaves, l-[U-14C]sorbosone was found to be equivalent to [6-14C]glucose as a source of 14C for AA. In the presence of 0.1% d-glucosone, conversion of [6-14C] glucose into labeled AA was greatly repressed. In a comparable experiment with l-sorbosone replacing d-glucosone, the effect was much less. The experiments described here give substance to the proposal that d-glucosone and l-sorbosone are putative intermediates in the conversion of d-glucose to AA in higher plants.  相似文献   

11.
Characterization of the hexose transport system in maize root tips   总被引:1,自引:1,他引:0  
Xia JH  Saglio PH 《Plant physiology》1988,88(4):1015-1020
Sugar-depleted excised maize (Zea mays L.) root tips were used to study the kinetics and the specificity of hexose uptake. It was found that difficulties induced by bulk diffusion and penetration barriers did not exist with root tips. Several lines of evidence indicate the existence of a complex set of uptake systems for hexoses showing an overall biphasic dependence on external sugar concentrations. The results suggest that the high and the low affinity components might be located on the same carrier. One uptake system was specific for fructose, but the high affinity component was repressed by high concentrations of external glucose. A second system was specific for glucose and its analogs (2-deoxy-d-glucose and 3-O-methyl-d-glucose), and a third one, more complex, had a high affinity for glucose and its analogs but could transport fructose when glucose was not present in the external solution. A simple method is proposed to determine the inhibitor constants in competition experiments.  相似文献   

12.
Lectins extracted from corn (Zea mays L.) kernel with Tris-HCl buffer pH 7.5 were isolated from the crude extract by affinity chromatography on Sepharose 6B-N-acetyl-d-galactosamine and Sepharose 6B-methylα-d-mannoside, and also by lectin affinity chromatography using concanavalin A and Lens culinaris lectin as ligands. According to preferential monosaccharide specificity, salt-soluble lectins of corn seed comprise at least two distinct types: N-acetyl-d-galactosamine-interactive and mannose-interactive lectins. The extracted lectins are unstable, with a tendency to form aggregates during storage.  相似文献   

13.
The metabolism of myo-inositol-2-14C, d-glucuronate-1-14C, d-glucuronate-6-14C, and l-methionine-methyl-14C to cell wall polysaccharides was investigated in excised root-tips of 3 day old Zea mays seedlings. From myo-inositol, about one-half of incorporated label was recovered in ethanol insoluble residues. Of this label, about 90% was solubilized by treatment, first with a preparation of pectinase-EDTA, then with dilute hydrochloric acid. The only labeled constituents in these hydrolyzates were d-galacturonic acid, d-glucuronic acid, 4-O-methyl-d-glucuronic acid, d-xylose, and l-arabinose, or larger oligosaccharide fragments containing these units. Medium external to excised root-tips grown under sterile conditions in myo-inositol-2-14C contained labeled polysaccharide.  相似文献   

14.
The mechanism of hexose transport into plasma membrane vesicles isolated from mature sugarbeet leaves (Beta vulgaris L.) was investigated. The initial rate of glucose uptake into the vesicles was stimulated approximately fivefold by imposing a transmembrane pH gradient (ΔpH), alkaline inside, and approximately fourfold by a negative membrane potential (ΔΨ), generated as a K+-diffusion potential, negative inside. The -fold stimulation was directly related to the relative ΔpH or ΔΨ gradient imposed, which were determined by the uptake of acetate or tetraphenylphosphonium, respectively. ΔΨ- and ΔpH-dependent glucose uptake showed saturation kinetics with a Km of 286 micromolar for glucose. Other hexose molecules (e.g. 2-deoxy-d-glucose, 3-O-methyl-d-glucose, and d-mannose) were also accumulated into plasma membrane vesicles in a ΔpH-dependent manner. Inhibition constants of a number of compounds for glucose uptake were determined. Effective inhibitors of glucose uptake included: 3-O-methyl-d-glucose, 5-thio-d-glucose, d-fructose, d-galactose, and d-mannose, but not 1-O-methyl-d-glucose, d- and l-xylose, l-glucose, d-ribose, and l-sorbose. Under all conditions of proton motive force magnitude and glucose and sucrose concentration tested, there was no effect of sucrose on glucose uptake. Thus, hexose transport on the sugarbeet leaf plasma membrane was by a H+-hexose symporter, and the carrier and possibly the energy source were not shared by the plasma membrane H+-sucrose symporter.  相似文献   

15.
Vancomycin-resistant enterococci acquire high-level resistance to glycopeptide antibiotics through the synthesis of peptidoglycan terminating in d-alanyl-d-lactate. A key enzyme in this process is a d-alanyl-d-alanine ligase homologue, VanA or VanB, which preferentially catalyzes the synthesis of the depsipeptide d-alanyl-d-lactate. We report the overexpression, purification, and enzymatic characterization of DdlN, a VanA and VanB homologue encoded by a gene of the vancomycin-producing organism Amycolatopsis orientalis C329.2. Evaluation of kinetic parameters for the synthesis of peptides and depsipeptides revealed a close relationship between VanA and DdlN in that depsipeptide formation was kinetically preferred at physiologic pH; however, the DdlN enzyme demonstrated a narrower substrate specificity and commensurately increased affinity for d-lactate in the C-terminal position over VanA. The results of these functional experiments also reinforce the results of previous studies that demonstrated that glycopeptide resistance enzymes from glycopeptide-producing bacteria are potential sources of resistance enzymes in clinically relevant bacteria.The origin of antibiotic resistance determinants is of significant interest for several reasons, including the prediction of the emergence and spread of resistance patterns, the design of new antimicrobial agents, and the identification of potential reservoirs for resistance elements. Antibiotic resistance can occur either through spontaneous mutation in the target or by the acquisition of external genetic elements such as plasmids or transposons which carry resistance genes (7). The origins of these acquired genes are varied, but it has long been recognized that potential reservoirs are antibiotic-producing organisms which naturally harbor antibiotic resistance genes to protect themselves from the actions of toxic compounds (6).High-level resistance to glycopeptide antibiotics such as vancomycin and teicoplanin in vancomycin-resistant enterococci (VRE) is conferred by the presence of three genes, vanH, vanA (or vanB), and vanX, which, along with auxiliary genes necessary for inducible gene expression, are found on transposons integrated into plasmids or the bacterial genome (1, 20). These three genes are essential to resistance and serve to change the C-terminal peptide portion of the peptidoglycan layer from d-alanyl-d-alanine (d-Ala-d-Ala) to d-alanyl-d-lactate (d-Ala-d-Lac). This change results in the loss of a critical hydrogen bond between vancomycin and the d-Ala-d-Ala terminus and in a 1,000-fold decrease in binding affinity between the antibiotic and the peptidoglycan layer, which is the basis for the bactericidal action of this class of compounds (5). The vanH gene encodes a d-lactate dehydrogenase which provides the requisite d-Lac (3, 5), while the vanX gene encodes a highly specific dd-peptidase which cleaves only d-Ala-d-Ala produced endogenously while leaving d-Ala-d-Lac intact (19, 21). The final gene, vanA or vanB, encodes an ATP-dependent d-Ala-d-Lac ligase (4, 8, 10). This enzyme has sequence homology with the chromosomal d-Ala-d-Ala ligases, which are essential for peptidoglycan synthesis but which generally lack the ability to synthesize d-Ala-d-Lac (9).We have recently cloned vanH, vanA, and vanX homologues from two glycopeptide antibiotic-synthesizing organisms: Amycolatopsis orientalis C329.2, which produces vancomycin, and Streptomyces toyocaensis NRRL 15009, which produces A47934 (14). In addition, the vanH-vanA-vanX gene cluster was identified in several other glycopeptide producers. We have also demonstrated that the VanA homologue from S. toyocaensis NRRL 15009 can synthesize d-Ala-d-Lac in vitro and in the glycopeptide-sensitive host Streptomyces lividans (15, 16). We now report the expression of the A. orientalis C329.2 VanA homologue DdlN in Escherichia coli, its purification, and its enzymatic characterization. These data reinforce the striking similarity between vancomycin resistance elements in VRE and glycopeptide-producing organisms and support the possibility of a common origin for these enzymes.

Expression, purification, and specificity of DdlN.

DdlN was overexpressed in E. coli under the control of the bacteriophage T7 promoter. The construct gave good yields of highly purified enzyme following a four-step purification procedure (Table (Table1;1; Fig. Fig.1).1). Like other dd-ligases, DdlN behaved like a dimer in solution (not shown).

TABLE 1

Purification of DdlN from E. coli BL21 (DE3)/pETDdlN
SampleProtein (mg)Activity (nmol/min)Sp act (nmol/ min/mg)Recovery (%)Purification (fold)
Lysate1248436.82100
Ammonium sulfate (20–50% saturation)67.678011.5921.7
Sephacryl S20011.682571.49811
Q Sepharose2.87422658839
Phenyl Superose0.429974835110
Open in a separate windowOpen in a separate windowFIG. 1Purification of DdlN from E. coli BL21 (DE3)/pETDdlN. Proteins were separated on an SDS–11% polyacrylamide gel and stained with Coomassie blue. Lane 1, molecular mass markers (masses are noted at the left in kilodaltons); lane 2, whole-cell lysate; lane 3, ammonium sulfate fraction (20 to 50% saturation); lane 4, Sephacryl S200; lane 5, Q Sepharose; lane 6, phenyl Superose.The amino acid substrate specificity of DdlN was assessed by incubation of 14C-d-Ala with all 20 common amino acids in the d configuration. Purified DdlN catalyzed the synthesis of d-Ala-d-Ala in addition to that of several other mixed dipeptides, including d-Ala-d-Met and d-Ala-d-Phe (Fig. (Fig.2).2). Thus, DdlN exhibits a substrate specificity which is similar to that of VanA (4), with the capacity to synthesize not only d-Ala-d-Ala but also mixed dipeptides with bulky side chains in the C-terminal position.Open in a separate windowFIG. 2Substrate specificity of DdlN. Autoradiogram from thin-layer chromatography analysis of DdlN substrate specificity. All reaction mixtures contained 2.5 mM d-Ala and 1 mM ATP, and the radiolabel was 14C-d-Ala, except where noted. Lane 1, d-Ala; lane 2, d-Lac with 14C-d-Lac label; lane 3, d,l-methionine; lane 4, dl-phenylalanine; lane 5, d-Hbut; lane 6, d-hydroxyvalerate. Letters indicate the following: A, d-Ala-d-Lac; B, d-Lac; C, d-Ala-d-Met; D, d-Ala-d-Phe; E, d-Ala-d-Hbut; F, d-Ala-d-hydroxyvalerate.Importantly, DdlN is a depsipeptide synthase with the ability to synthesize d-Ala-d-Lac, d-Ala-d-hydroxybutyrate (Hbut), and d-Ala-d-hydroxyvalerate (Fig. (Fig.2).2). However, unlike VanA (5), d-hydroxycaproate and d-phenyllactate are not substrates (not shown). Thus, DdlN is a broad-spectrum d-Ala-d-X ligase with depsipeptide synthase activity.

Characterization of d-Ala-d-X ligase activity.

Following the initial assessment of the specificity of the enzyme, several substrates were selected for quantitative analysis by evaluation of their steady-state kinetic parameters (Table (Table2).2). DdlN has two amino acid (or hydroxy acid) Km values. Steady-state kinetic plots indicated that, like other dd-ligases, the N-terminal Km (Km1) was significantly lower (higher specificity) than the C-terminal Km (Km2). Since the former value is expected to be independent of the C-terminal substrate, only Km2 values were determined and are reported here.

TABLE 2

Characterization of steady-state parameters of DdlN and VanA
LigaseSubstrateKm2 (mM)kcat (min−1)kcat/Km2 (M−1 s−1)
DdlNd-Ala21 ± 2229 ± 71.8 × 102
d-Lac0.4 ± 0.0555 ± 12.3 × 103
d-Hbut2.5 ± 0.332 ± 22.1 × 102
ATPa1.2 ± 0.271 ± 50.98 × 102
DdlMbd-Ala166 ± 27
d-Lac1.08 ± 0.10
VanAcd-Ala382951.3 × 102
d-Lac7.1942.2 × 102
d-Hbut0.601083.0 × 103
Open in a separate windowa Determined in the presence of 10 mM d-Lac. b Data from reference 16c Data from reference 5. DdlN showed good d-Ala-d-Ala ligase activity but with a very high and physiologically questionable Km2 (21 mM). On the other hand, d-Ala-d-Lac synthesis was excellent, with a 4-fold decrease in kcat, compared to d-Ala-d-Ala synthesis, which was offset by a 52-fold drop in Km that resulted in a >12-fold increase in specificity (kcat/Km2). d-Hbut was also a good substrate, with a kcat/Km2 comparable to that of d-Ala.Steady-state kinetic parameters for d-Ala-d-X formation showed trends similar to those found with both VanA and DdlN. For example, the kcat values between VanA and DdlN were virtually the same for most substrates. There were significant differences, however. For instance, while the Km2 values for d-Ala were very high for all three enzymes, DdlN does have greater affinity for d-Ala, with a 1.8- and 7.9-fold lower Km2 than those of VanA and DdlM, respectively. Additionally, the Km2 for d-Lac was 17.8- and 2.7-fold lower than those for VanA and DdlM. Thus, DdlN has a more restrictive specificity for the C-terminal residue than VanA, which is compensated for by a higher affinity for the critical substrate d-Lac.

pH dependence of peptide versus that of depsipeptide synthesis activity.

The partitioning of the syntheses of d-Ala-d-Ala and d-Ala-d-Hbut in VanA and other depsipeptide-competent dd-ligases has been shown to be pH dependent (17). Determination of the pH dependence of DdlN in synthesizing peptide versus depsipeptide (Fig. (Fig.3)3) directly paralleled the results obtained with VanA in similar experiments. At lower pHs (<7), d-Ala-d-Hbut synthesis predominates and is exclusive at a pH of <6 (Fig. (Fig.3).3). At pH 7.5, levels of synthesis of d-Ala-d-Hbut and d-Ala-d-Ala are relatively equal, while at a pH greater than 8, the capacity to synthesize peptide overtakes the capacity to synthesize depsipeptide, although the latter is never abolished. Open in a separate windowFIG. 3pH dependence of partitioning of the syntheses of peptide and depsipeptide by DdlN. (A) Autoradiogram of a thin-layer chromatography separation of the products of reaction mixtures containing 14C-D-Ala, unlabeled D-Ala, and d-Hbut. (B) Quantification of reaction products following phosphorimage analysis. Filled circles, D-Ala-d-Hbut; open circles, D-Ala-D-Ala.The partitioning of the formation of peptide versus depsipeptide as a function of pH by DdlM is comparable to that by VanA and depsipeptide-competent mutants of DdlB (17), which show essentially exclusively depsipeptide formation at lower pHs and increasing peptide formation as the pH increases. This implies a potential role for the protonated ammonium group of d-Ala2 in second-substrate recognition and suggests a mechanism for the discrimination between d-Ala and d-Lac at physiologic pH. The structural basis for this distinction remains obscure for DdlB and VanA or DdlN.

Concluding remarks.

Resistance to vancomycin and other glycopeptides is mediated through the synthesis of a peptidoglycan which does not terminate with the canonical d-Ala-d-Ala dipeptide. Thus, enterococci which exhibit the VanC phenotype, which consists of low-level, noninducible resistance to vancomycin only, have peptidoglycan terminating in d-Ala-d-Ser (19). On the other hand, bacteria which are constitutively resistant to high concentrations of glycopeptides, such as lactic acid bacteria and VRE exhibiting the VanA or VanB phenotype (high-level inducible resistance to vancomycin), incorporate the depsipeptide d-Ala-d-Lac into their cell walls (2, 12, 13). The enzymes responsible for the intracellular synthesis of d-Ala-d-Lac not surprisingly have significant amino acid sequence similarity with d-Ala-d-Ala ligases, which are responsible for d-Ala-d-Ala synthesis in all bacteria with a cell wall (9).The d-Ala-d-Lac synthases can be subdivided into two groups based on sequence homology: those found in the constitutively resistant lactic acid bacteria and those found in glycopeptide-producing organisms and VanA or VanB VRE (9, 14). The former have more similarity with exclusive d-Ala-d-Ala ligases. Indeed, single point mutations in d-Ala-d-Ala ligases which yield sequences more similar to those of lactic acid bacterium d-Ala-d-Lac ligases are sufficient to induce significant depsipeptide synthase activity in these enzymes (17). Similarly, mutational studies of the d-Ala-d-Lac ligase from Leuconostoc mesenteroides have demonstrated that the converse also holds (18). On the other hand, the molecular basis for depsipeptide synthesis by the VanA or VanB ligases is unknown, in large part due to the lack of protein structural information on which to base mutational studies, unlike the situation with d-Ala-d-Ala ligases, where the E. coli DdlB structure serves as a template for mechanistic research (11).Significantly, a major difference in the VanA or VanB ligases and other dd-ligases lies in the amino acid sequence of the ω-loop region, which closes off the active site of DdlB (11) and has been shown to contribute amino acid residues with the capacity to control the syntheses of d-Ala-d-Ala and d-Ala-d-Lac, notably, Tyr216 (17, 18). Until recently, the VanA and VanB ligases were exceptional in amino acid structure and had no known homologues. The sequencing of resistance genes from glycopeptide-producing bacteria has uncovered enzymes with >60% homology to VanA or VanB and which are virtually superimposable in the critical ω-loop region (14, 15). One of these, DdlM from S. toyocaensis NRRL 15009, has been shown to have d-Ala-d-Lac ligase ability (15, 16), although no rigorous analysis of this activity has been performed. The results presented here demonstrate that DdlN from the vancomycin producer A. orientalis C329.2 not only is a d-Ala-d-Lac ligase but also has significant functional homology with VanA. It is not known at present if, like S. toyocaensis NRRL 15009 (16), A. orientalis C329.2 also possess a d-Ala-d-Ala-exclusive ligase, though the presence of a vanX gene (14) suggests that it may.These studies demonstrate that DdlN cloned from a vancomycin-producing bacterium is a d-Ala-d-Lac ligase which has not only amino acid sequence homology with the dd-ligases from VRE but also functional homology. Thus, VanA, VanB, DdlN, and DdlM have likely evolved from similar origins. The fact that a vanH-vanA-vanX gene cluster can be found in other glycopeptide producers as well (14) suggests that the genes now found in VRE may have originated in glycopeptide-producing bacteria. Our finding that overexpressed, purified, DdlN shows many enzymatic characteristics similar (though not identical) to those of VanA suggests that the genes from glycopeptide-producing bacteria can be important in elucidating biochemical and protein structural aspects of the VRE proteins.  相似文献   

16.
The serP1 and serP2 genes found adjacently on the chromosome of Lactococcus lactis strains encode two members of the amino acid-polyamine-organocation (APC) superfamily of secondary transporters that share 61% sequence identity. SerP1 transports l-serine, l-threonine, and l-cysteine with high affinity. Affinity constants (Km) are in the 20 to 40 μM range. SerP2 is a dl-alanine/dl-serine/glycine transporter. The preferred substrate appears to be dl-alanine for which the affinities were found to be 38 and 20 μM for the d and l isomers, respectively. The common substrate l-serine is a high-affinity substrate of SerP1 and a low-affinity substrate of SerP2 with affinity constants of 18 and 356 μM, respectively. Growth experiments demonstrate that SerP1 is the main l-serine transporter responsible for optimal growth in media containing free amino acids as the sole source of amino acids. SerP2 is able to replace SerP1 in this role only in medium lacking the high-affinity substrates l-alanine and glycine. SerP2 plays an adverse role for the cell by being solely responsible for the uptake of toxic d-serine. The main function of SerP2 is in cell wall biosynthesis through the uptake of d-alanine, an essential precursor in peptidoglycan synthesis. SerP2 has overlapping substrate specificity and shares 42% sequence identity with CycA of Escherichia coli, a transporter whose involvement in peptidoglycan synthesis is well established. No evidence was obtained for a role of SerP1 and SerP2 in the excretion of excess amino acids during growth of L. lactis on protein/peptide-rich media.  相似文献   

17.
Thermotoga maritima is a Gram-negative, hyperthermophilic bacterium whose peptidoglycan contains comparable amounts of l- and d-lysine. We have determined the fine structure of this cell-wall polymer. The muropeptides resulting from the digestion of peptidoglycan by mutanolysin were separated by high-performance liquid chromatography and identified by amino acid analysis after acid hydrolysis, dinitrophenylation, enzymatic determination of the configuration of the chiral amino acids, and mass spectrometry. The high-performance liquid chromatography profile contained four main peaks, two monomers, and two dimers, plus a few minor peaks corresponding to anhydro forms. The first monomer was the d-lysine-containing disaccharide-tripeptide in which the d-Glu-d-Lys bond had the unusual γ→ϵ arrangement (GlcNAc-MurNAc-l-Ala-γ-d-Glu-ϵ-d-Lys). The second monomer was the conventional disaccharide-tetrapeptide (GlcNAc-MurNAc-l-Ala-γ-d-Glu-l-Lys-d-Ala). The first dimer contained a disaccharide-l-Ala as the acyl donor cross-linked to the α-amine of d-Lys in a tripeptide acceptor stem with the sequence of the first monomer. In the second dimer, donor and acceptor stems with the sequences of the second and first monomers, respectively, were connected by a d-Ala4-α-d-Lys3 cross-link. The cross-linking index was 10 with an average chain length of 30 disaccharide units. The structure of the peptidoglycan of T. maritima revealed for the first time the key role of d-Lys in peptidoglycan synthesis, both as a surrogate of l-Lys or meso-diaminopimelic acid at the third position of peptide stems and in the formation of novel cross-links of the l-Ala1(α→α)d-Lys3 and d-Ala4(α→α)d-Lys3 types.Peptidoglycan (or murein) is a giant macromolecule whose main function is the protection of the cytoplasmic membrane against the internal osmotic pressure. It is composed of alternating residues of N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc)2 cross-linked by short peptides (1). The composition of the peptide stem in nascent peptidoglycan is l-Ala1-γ-d-Glu2-X3-d-Ala4-d-Ala5, where X is most often meso-diaminopimelic acid (meso-A2pm) or l-lysine in Gram-negative and Gram-positive species, respectively (2, 3). In the mature macromolecule, the last d-Ala residue is removed. Cross-linking of the glycan chains generally occurs between the carboxyl group of d-Ala at position 4 of a donor peptide stem and the side-chain amino group of the diamino acid at position 3 of an acceptor peptide stem (4→3 cross-links). Cross-linking is either direct or through a short peptide bridge such as pentaglycine in Staphylococcus aureus (2, 3). The enzymes for the formation of the 4→3 cross-links are active-site serine dd- transpeptidases that belong to the penicillin-binding protein (PBP) family and are the essential targets of β-lactam antibiotics in pathogenic bacteria (4). Catalysis involves the cleavage of the d-Ala4-d-Ala5 bond of a donor peptide stem and the formation of an amide bond between the carboxyl of d-Ala4 and the side chain amine at the third position of an acceptor stem. Transpeptidases of the ld specificity are active-site cysteine enzymes that were shown to act as surrogates of the PBPs in mutants of Enterococcus faecium resistant to β-lactam antibiotics (5). They cleave the X3-d-Ala4 bond of a donor stem peptide to form 3→3 cross-links. This alternate mode of cross-linking is usually marginal, although it has recently been shown to predominate in non-replicative “dormant” forms of Mycobacterium tuberculosis (6).Thermotoga maritima is a Gram-negative, extremely thermophilic bacterium isolated from geothermally heated sea floors by Huber et al. (7). A morphological characteristic is the presence of an outer sheath-like envelope called “toga.” Although the organism has received considerable attention for its biotechnological potential, studies about its peptidoglycan are scarce (811), and in particular the fine structure of the macromolecule is still unknown. In their initial work, Huber et al. (7) showed that the composition of its peptidoglycan was unusual for a Gram-negative species, because it contained both isomers of lysine and no A2pm. Recently, we purified and studied the properties of T. maritima MurE (12); this enzyme is responsible for the addition of the amino acid residue at position 3 of the peptide stem (13, 14). We demonstrated that T. maritima MurE added in vitro l- and d-Lys to UDP-MurNAc-l-Ala-d-Glu. Although l-Lys was added in the usual way, yielding the conventional nucleotide UDP-MurNAc-l-Ala-γ-d-Glu-l-Lys containing a d-Glu(γ→α)l-Lys amide bond, the d-isomer was added in an “upside-down” manner, yielding the novel nucleotide UDP-MurNAc-l-Ala-d-Glu(γ→ϵ)d-Lys. We also showed that the d-Lys-containing nucleotide was not a substrate for T. maritima MurF, the subsequent enzyme in the biosynthetic pathway, whereas this ligase catalyzed the addition of dipeptide d-Ala-d-Ala to the l-Lys-containing tripeptide, yielding the conventional UDP-MurNAc-pentapeptide (12).However, both the l-Lys-containing UDP-MurNAc-pentapeptide and d-Lys-containing UDP-MurNAc-tripeptide were used as substrates by T. maritima MraY with comparable efficiencies in vitro (12). This observation implies that the unusual d-Lys-containing peptide stems are likely to be translocated to the periplasmic face of the cytoplasmic membrane and to participate in peptidoglycan polymerization. Therefore, we have determined here the fine structure of T. maritima peptidoglycan and we have shown that l-Lys- and d-Lys-containing peptide stems are both present in the polymer, the latter being involved in the formation of two novel types of peptidoglycan cross-link.  相似文献   

18.
1. Human uterine cervical stroma was found to contain a Ca2+-independent neutral proteinase against casein and N-benzoyl-dl-arginine p-nitroanilide (Bz-dl-Arg-Nan). This enzyme was tightly bound to an insoluble material (20000g pellet) and was solubilized by high concentrations of NaCl or KCl. High concentrations of them in the reaction system, however, inhibited reversibly the activity of this enzyme. 2. The neutral proteinase was partially purified by extraction with NaCl, gel filtration on Sephadex G-200 and affinity chromatography on casein–Sepharose. 3. The optimal pH of this partially purified enzyme was 7.4–8.0 against casein and Bz-dl-Arg-Nan. The molecular weight of the enzyme was found to be about 1.4×105 by gel filtration on Sephadex G-200. 4. The enzyme was significantly inhibited by di-isopropyl phosphorofluoridate (0.1mm). High concentration of phenylmethanesulphonyl fluoride (5mm), 7-amino-1-chloro-3-l-tosylamidoheptan-2-one (0.5mm), antipain (10μm) or leupeptin (10μm) was also found to be inhibitory, but chymostatin (40μg/ml), soya-bean trypsin inhibitor (2.5mg/ml), human plasma (10%, v/v), p-chloromercuribenzoate (1mm), EDTA (10mm) and 1-chloro-4-phenyl-3-l-tosylamidobutan-2-one (1mm) had no effect on the enzyme. 5. The neutral proteinase hydrolysed casein, Bz-dl-Arg-Nan and heat-denatured collagen, but was inactive towards native collagen and several synthetic substrates, such as 4-phenylazobenzyloxycarbonyl-Pro-Leu-Gly-Pro-d-Arg, 3-carboxypropionyl-Ala-Ala-Ala p-nitroanilide and 2,4-dinitrophenyl-Pro-Gln-Gly-Ile-Ala-Gly-Gln-d-Arg, and also proteoglycan. The enzyme did not act as a plasminogen activator. 6. These properties suggested that a neutral proteinase in the human uterine cervix was different from enzymes previously reported.  相似文献   

19.
In pancreatic β-cells, uptake of Ca2+ into mitochondria facilitates metabolism-secretion coupling by activation of various matrix enzymes, thus facilitating ATP generation by oxidative phosphorylation and, in turn, augmenting insulin release. We employed an siRNA-based approach to evaluate the individual contribution of four proteins that were recently described to be engaged in mitochondrial Ca2+ sequestration in clonal INS-1 832/13 pancreatic β-cells: the mitochondrial Ca2+ uptake 1 (MICU1), mitochondrial Ca2+ uniporter (MCU), uncoupling protein 2 (UCP2), and leucine zipper EF-hand-containing transmembrane protein 1 (LETM1). Using a FRET-based genetically encoded Ca2+ sensor targeted to mitochondria, we show that a transient knockdown of MICU1 or MCU diminished mitochondrial Ca2+ uptake upon both intracellular Ca2+ release and Ca2+ entry via L-type channels. In contrast, knockdown of UCP2 and LETM1 exclusively reduced mitochondrial Ca2+ uptake in response to either intracellular Ca2+ release or Ca2+ entry, respectively. Therefore, we further investigated the role of MICU1 and MCU in metabolism-secretion coupling. Diminution of MICU1 or MCU reduced mitochondrial Ca2+ uptake in response to d-glucose, whereas d-glucose-triggered cytosolic Ca2+ oscillations remained unaffected. Moreover, d-glucose-evoked increases in cytosolic ATP and d-glucose-stimulated insulin secretion were diminished in MICU1- or MCU-silenced cells. Our data highlight the crucial role of MICU1 and MCU in mitochondrial Ca2+ uptake in pancreatic β-cells and their involvement in the positive feedback required for sustained insulin secretion.  相似文献   

20.
Renal transport of four different categories of organic solutes, namely sugars, neutral amino acids, monocarboxylic acids and dicarboxylic acids, was studied by using the potential-sensitive dye 3,3′-diethyloxadicarbocyanine iodide in purified luminal-membrane and basolateral-membrane vesicles isolated from rabbit kidney cortex. Valinomycin-induced K+ diffusion potentials resulted in concomitant changes in dye–membrane-vesicle absorption spectra. Linear relationships were obtained between these changes and depolarization and hyperpolarization of the vesicles. Addition of d-glucose, l-phenylalanine, succinate or l-lactate to luminal-membrane vesicles, in the presence of an extravesicular>intravesicular Na+ gradient, resulted in rapid transient depolarization. With basolateral-membrane vesicles no electrogenic transport of d-glucose or l-phenylalanine was observed. Spectrophotometric competition studies revealed that d-galactose is electrogenically taken up by the same transport system as that for d-glucose, whereas l-phenylalanine, succinate and l-lactate are transported by different systems in luminal-membrane vesicles. The absorbance changes associated with simultaneous addition of d-glucose and l-phenylalanine were additive. The uptake of these solutes was influenced by the presence of Na+-salt anions of different permeabilities in the order: Cl>SO42−>gluconate. Addition of valinomycin to K+-loaded vesicles enhanced uptake of d-glucose and l-phenylalanine in the presence of an extravesicular>intravesicular Na+ gradient. Gramicidin or valinomycin plus nigericin diminished/abolished electrogenic solute uptake by Na+- or Na++K+-loaded vesicles respectively. These results strongly support the presence of Na+-dependent renal electrogenic transport of d-glucose, l-phenylalanine, succinate and l-lactate in luminal-membrane vesicles.  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