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 312 毫秒
1.
Poly[(R)-3-hydroxybutyrate] (PHB) depolymerase from Ralstonia pickettii T1 (PhaZRpiT1) consists of three functional domains to effectively degrade solid PHB materials, and its catalytic domain catalyzes the ester bond cleavage of the substrate. We performed the directed evolution of PhaZRpiT1 targeted at the catalytic domain in combination with the cell surface display method to effectively screen for mutants with improved p-nitrophenyl butyrate (pNPC4) activity. Mutated PhaZRpiT1 genes generated by error-prone PCR were fused to the oprI gene to display them as fusion proteins on Escherichia coli cell surface. Some cells displaying the mutant enzymes showed a two- to fourfold increase in pNPC4 hydrolysis activity relative to cells displaying wild-type enzyme. These mutant genes were recombined by a staggered extension process and the recombined enzymes were displayed to result in a five- to eightfold higher pNPC4 hydrolysis activity than the wild type. To further evaluate the mutation effects, unfused and undisplayed enzymes were prepared and applied to the hydrolysis of p-nitrophenyl esters having different chain lengths (pNPCn; n?=?2–6) and PHB degradation. One specific second-generation mutant showed an approximately tenfold increase in maximum rate for pNPC3 hydrolysis, although its PHB degradation efficiency at 1 μg/mL of enzyme concentration was approximately 3.5-fold lower than that of the wild type. Gene analysis showed that N285D or N285Y mutations were found in six of the seven improved second-generation mutants, indicating that Asn285 probably participates in the regulation of substrate recognition and may be more favorable for PHB degradation process than other amino acid residues.  相似文献   

2.
The cathepsin B-like proteinase from Helicoverpa armigera (HCB) is involved in the degradation of yolk proteins during embryonic development. In order to gain insight into the substrate specificity of this proteinase, various proteins from animals and plants were tested as substrates. The specific cleavage sites of this enzyme on endopeptide bonds were assayed using bovine serum albumin (BSA) as a substrate. Results showed that BSA was degraded into several fragments, which suggests that HCB cleaves BSA at specific endopeptidyl sites. The amino acid sequences of the BSA derived peptides were determined, revealing cleavage of the bonds between residues Arg81–Glu82, Val423–Glu424 and Gly430–Lys431. This suggests that the minimum requirement for a scissile bond to be recognized by HCB is the presence of an ionic amino acid at the P1 position and the P1 position can vary. These observations suggest that HCB cleaves bonds at the N-terminal side of ionic amino acid residues giving HCB a wide range of substrates, though other factors dictating the substrate specificity of this enzyme remains to be clarified. Our results provide new evidence that HCB functions as an endopeptidase on some proteins.  相似文献   

3.
The chromatographic conditions for separation of N,N′-bislysyl(?-N)adipamidine and N-lysyl(?-N)adipamidinic acid, which were the products of acid hydrolysis of proteins treated with adipimidate esters, from other amino acids on an amino acid analyzer were established including their ninhydrin color values. Kinetics of decomposition of these lysine derivatives under the conditions of total acid hydrolysis of protein are also reported.  相似文献   

4.
The in vitro digestibility of rice glutelin and wheat glutenin was investigated with a view to assessing their nutritional qualities, using casein and bovine serum albumin (BSA) as references. The following hydrolytic processes were adopted: pepsin-pancreation digestion (a model system before intestinal absorption) and aminopeptidase-prolidase hydrolysis [a model system for the intestinal mucosa (membrane digestion) and after intestinal absorption (intracellular hydrolysis)]. The pepsin-pancreatin digests were first examined. The degree of amino acid released from the proteins was 30% (glutelin), 23% (glutenin), 24% (casein) and 30% (BSA). A similar release pattern of individual amino acids was observed for all the proteins. The amounts of large peptide fractions increased in the order: glutelin < glutenin < casein < BSA. Glutelin was highly digestible. Apart from containing high amounts of glutamic acid (glutamine), cystine and proline, the large peptide fractions of glutelin were also rich in threonine, glycine and isoleucine while those of glutenin were only rich in glycine. The aminopeptidase-prolidase digests were examined next. Glutelin was almost completely hydrolyzed to amino acid, except for a low release of cystine, suggesting that the amino acid residues constituting glutelin could be easily utilized as nutrients in the living tissues. The degree of amino acid released from the proteins was 97% (glutelin), 93% (glutenin), 90% (casein) and 79% (BSA).

The convenient application of these model systems for the assessment of the in vitro digestibility of food proteins have been discussed.  相似文献   

5.
Beef liver esterase. II. Kinetic properties   总被引:1,自引:0,他引:1  
The kinetic parameters, kcat and KM, in beef liver esterase-catalyzed hydrolysis were determined for about 100 substrates, which can be classified in several groups: (1) In the ethyl ester series of fatty acids KM decreases with elongation of the acid, while kcat has a maximum value with pentanoate. (2) Alkyl acetates are better substrates as the alkyl moiety is longer, whereas esters with branched alkyl groups become worse substrates. (3) Aryl esters are very good substrates. (4) Esters of dicarboxylic acids are good substrates, but only one ester group is cleaved by the enzyme. Fumarate diester is susceptible to esterase hydrolysis, while maleate is not. (5) Esters of hydrophobic amino acids are very good substrates; the enzyme is not stereoselective and both the l and d stereoisomers are readily hydrolyzed. Branching at the β-carbon atom leads to loss of activity, and blocking of the amino group abolishes it. Fluoride ion and dl-malate esters are potent competitive inhibitors of the enzymic reaction. The optimal pH was found to lie between 8 and 8.5. The reaction rate increased between 5 and 40 °C then dropped sharply. The activity decreased at high salt concentration.  相似文献   

6.
Chlamydia, like other intracellular bacteria, are auxotrophic for a variety of essential metabolites and obtain cholesterol and fatty acids from their eukaryotic host cell, however not many Chlamydia-specific enzymes have been identified that are involved in lipid metabolism. In silico analysis of one candidate Chlamydia trachomatis enzyme, annotated as a conserved putative hydrolase (CT149), identified two lipase/esterase GXSXG motifs, and a potential cholesterol recognition/interaction amino acid consensus (CRAC) sequence. His-tag purified recombinant CT149 exhibited ester hydrolysis activity in a nitrophenyl acetate-based cell-free assay system. When cholesteryl linoleate was used as substrate, ester hydrolysis occurred and production of cholesterol was detected by high performance liquid chromatography. Exogenous expression of transfected CT149 in HeLa cells resulted in a significant decrease of cytoplasmic cholesteryl esters within 48 h. These results demonstrate that CT149 has cholesterol esterase activity and is likely to contribute to the hydrolysis of eukaryotic cholesteryl esters during intracellular chlamydial growth.  相似文献   

7.
We have purified the MutL protein from Rhodobacter sphaeroides mismatch repair system (rsMutL) for the first time. rsMutL demonstrated endonuclease activity in vitro, as predicted by bioinformatics analysis. Based on the alignment of 1483 sequences of bacterial MutL homologs with presumed endonuclease activity, conserved functional motifs and amino acid residues in the rsMutL sequence were identified: five motifs comprising the catalytic site responsible for DNA cleavage were found in the C–terminal domain; seven conserved motifs involved in ATP binding and hydrolysis and specific to the GHKL family of ATPases were found in the N–terminal domain. rsMutL demonstrated the highest activity in the presence of Mn2+. The extent of plasmid DNA hydrolysis declined in the row Mn2+ > Co2+ > Mg2+ > Cd2+; Ni2+ and Ca2+ did not activate rsMutL. Divalent zinc ions inhibited rsMutL endonuclease activity in the presence of Mn2+ excess. ATP also suppressed plasmid DNA hydrolysis by rsMutL. Analysis of amino acid sequences and biochemical properties of five studied bacterial MutL homologs with endonuclease activity revealed that rsMutL resembles the MutL proteins from Neisseria gonorrhoeae and Pseudomonas aeruginosa.  相似文献   

8.
The spontaneous hydrolysis of glycosylamines, where the aglycone is either a primary amine or ammonia, is over a hundred million-times faster than that of O- or S-glycosides. The reason for this (as pointed out by Capon and Connett in 1965) is that, in contrast to the mechanism for O- or S-glycoside hydrolysis, hydrolysis of these N-glycosides (e.g., glc-NHR) involves an endocyclic C-O bond cleavage resulting in formation of an imine (iminium ion) which then reacts with water. Since ring-opening is kinetically favored with glycosylamines, compounds such as phenylglucosylamine can be a useful probes of enzymes that have been suggested to possibly follow this mechanism. With β-glucosidase from sweet almonds, the enzyme is highly efficient in catalyzing the hydrolysis of phenyl glucoside (kcat/knon ∼ 1014) and phenyl thioglucoside (kcat/knon ∼ 1010) while with either the almond or the Aspergillus niger enzyme or with yeast α-glucosidase, there is no detectable catalysis of phenylglucosylamine hydrolysis (kcat/knon < 20). These results are consistent with the generally accepted mechanism involving exocyclic bond cleavage by these enzymes.  相似文献   

9.
The feasibility of monitoring the peptic and tryptic digestion of peptides and proteins with capillary electrophoresis using contactless conductivity detection was investigated. The peptide minigastrin I and the proteins cytochrome c from bovine heart, human serum albumin (HSA), myoglobin, and bovine serum albumin (BSA) were digested off-line with pepsin, and the resulting peptide and amino acid fragments were successfully separated and detected by conductivity measurement. Cytochrome c and myoglobin were also subjected to off-line cleavage with trypsin. On-line digestion using the electrophoretically mediated microanalysis (EMMA) approach was demonstrated with cytochrome c and apomyoglobin using trypsin.  相似文献   

10.
In this study, we report the characterization of a protein from Aspergillus oryzae, exhibiting sequence identity with paraben esterase from the genus Aspergillus. The coding region of 1,586 bp, including a 77-bp intron, encoded a protein of 502 amino acids. The gene without the signal peptide of 19 amino acids was cloned into a vector, pPICZαC, and expressed successfully in Pichia pastoris as an active extracellular protein. The purified recombinant protein had pH and temperature optima of 7.0–8.0 and 30 °C, respectively, and was stable at the pH range of 7.0–10.0 and up to 40 °C. The optimal substrate for hydrolysis by the purified recombinant protein, among a panel of α-naphthyl esters (C2–C16), was α-naphthyl butyrate (C4), with activity of 0.16 units/mg protein. The considerable hydrolytic activity of the purified recombinant enzyme toward tributyrin was determined. However, no paraben esterase activity was detected toward the ethyl, propyl, and butyl esters of 4-hydroxybenzoic acid. In addition, no activity was detected toward the methyl esters of ferulic, p-coumaric, caffeic, and sinapic acids that would indicate feruloyl esterase activity.  相似文献   

11.
《Inorganica chimica acta》1986,123(3):175-179
The palladium(II)-promoted hydrolysis of the methyl esters of glycyl-L-leucine, glycyl-L-alanine and L-alanylglycine have been studied at 25 °C and I=0.1 M in the pH range 4–5. At a 1:1 metal to ligand ratio the peptide esters act as tridentate ligands, donation occurring via the terminal amino group, the deprotonated amide nitrogen, and the carbonyl group of the ester. Due to the high Lewis acidity of Pd(II) rapid hydrolysis of the ester function by water and hydroxide ion occurs. Rate constants kOH and kH2O have been obtained for base hydrolysis and water hydrolysis of the coordinated peptide esters at 25 °C. The rate constants for base hydrolysis are 3.4 X 106 M−1 s−1 (L-alaglyOMe), 6.4 X 106 M−1 s−1 (gly-L-alaOMe) and 2.3 X 107 M−1 s−1 (gly-L-leuOMe). Base hydrolysis of the coordinated peptide esters is at least 106 times that of the free unprotonated ligand. Activation parameters have been obtained for both water and base hydrolysis of the Pd(II) complex of methyl L-alanylglycinate and possible mechanisms for the hydrolyses are considered.  相似文献   

12.
《Insect Biochemistry》1991,21(2):197-203
The induction of trypsin activity in the midgut of the mosquito, Aedes aegypti, was studied following meals of chicken blood, and several protein and peptide diets. Various concentrations of bovine serum albumin (BSA) in 0.15 M NaCl stimulated trypsin activity, in a similar fashion to the initial increase observed after a normal blood meal. Trypsin synthesis was also initiated when Ae. aegypti were fed on glutaraldehyde cross-linked BSA and on BSA fragments prepared by both pepsin and cyanogen bromide cleavage. Non-soluble proteins, in the form of glutaraldehyde-fixed erythrocyte ghosts, induced a delayed and reduced trypsin response, whilst small peptides from neutralized liver digests did not induce trypsin activity until 8–10 h after feeding. Metabolic inhibitors had varying effects on the post-feeding activity of trypsin stimulated by BSA feeding. Cycloheximide, a peptidyl transferase inhibitor prevented expression of all activity in vivo, whereas α-amanitin (RNA-polymerase inhibitor) did not affect trypsin activity in the first 10 h after feeding. At 20 μg/ml concentration in the diet, actinomycin D (RNA synthesis inhibitor) caused temporary superinduction followed by inhibition of trypsin activity, but at lower concentrations, the later phase of trypsin activity was inhibited. The results suggest that post-feeding induction of trypsin activity in Ae. aegypti is a two-phase process regulated at the midgut cellular level. The first phase of trypsin synthesis is stimulated by soluble proteins of variable molecular weights, and only involves translation of messenger RNA already available within the midgut cells. The second phase is stimulated by small peptides and requires complete synthesis of new mRNA from DNA.  相似文献   

13.
The ability of Escherichia coli to grow on a series of acetylated and glycosylated compounds has been investigated. It is surmised that E. coli maintains low levels of nonspecific esterase activity. This observation may have ramifications for previous reports that relied on nonspecific esterases from E. coli to genetically encode nonnatural amino acids. It had been reported that nonspecific esterases from E. coli deacetylate tri-acetyl O-linked glycosylated serine and threonine in vivo. The glycosylated amino acids were reported to have been genetically encoded into proteins in response to the amber stop codon. However, it is our contention that such amino acids are not utilized in this manner within E. coli. The current results report in vitro analysis of the original enzyme and an in vivo analysis of a glycosylated amino acid. It is concluded that the amber suppression method with nonnatural amino acids may require a caveat for use in certain instances.The central question addressed in this paper is whether the glycosylated amino acids GlcNAc-Ser and GalNAc-Thr have been genetically encoded into proteins in vivo (1, 2). The reports for the incorporation of these two amino acids are unique from all other reports (3) that have incorporated unnatural amino acids using the recoded UAG codon and Methanococcus jannaschii orthogonal pairs in that these two amino acids required further processing by the host organism before incorporation (see Fig. 1). Here we posit that the primary barrier to their incorporation would appear to be the fact that the host organism used in the original reports, Escherichia coli, maintains very low levels of nonspecific esterase activity. In fact, the original reports used citations from mammalian biology to substantiate the nonspecific esterase mechanism (see below).Open in a separate windowFIGURE 1.Proposed product of an esterase with GlcNAc-Ser and other esterase substrates discussed in this study.E. coli is likely the most thoroughly studied microorganism. This is especially true in regard to carbohydrate and amino acid uptake and utilization (4). Therefore, it should not be surprising that it has long been known that esterified carbon sources are not metabolized by E. coli in standard assays used to probe for microorganism lipase and esterase activity (5). Such results and our current analysis underscore the limitations of the reports that triacetyl O-linked glycosylated amino acids (GlcNAc-Ser and GalNAc-Thr) were deacetylated in E. coli by endogenous “nonspecific” esterases. The deacetylated amino acids were then believed to have been genetically encoded into full-length proteins in vivo (1, 2).In these previous studies the glycosylated amino acids were provided to the growth media as their tetraacetate analogs, and it was construed from the mass spectra and lectin binding assays that the ester groups of the saccharide had all been hydrolyzed. The notion that E. coli rapidly hydrolyzes a simple ester is not easily reconciled with what is commonly observed when the ester functional group is introduced into cultures of E. coli. For example, we were prompted by reports that claimed to have harvested β-hydroxy esters from E. coli (6). There was nothing in such a report to indicate that the E. coli strain used had undergone a drastic genetic modification beyond the introduction of one enzyme derived from yeast. The enzyme from yeast was expressed in E. coli to asymmetrically reduce β-keto esters to the corresponding β-hydroxy esters. The reduction was accomplished in 87% yield and was performed in whole cells. It stands to reason that such a report having claimed to extract significant amounts of an esterified product would not be possible if E. coli maintained even moderate levels of nonspecific esterase activity. The fact that E. coli maintains low levels of endogenous esterases and lipases has been quite pivotal for a number of studies that have used this organism as the host to express esterase genes in vivo (see below).Nonspecific esterase activity is common in eukaryotic organisms, for example, our ability to hydrolyze triacylglycerides to access an important energy source, but this stands in stark contrast to E. coli where it is possible to directly extract O-acetylated oligosaccharides (7) and other simple esters (6) in high yields. These reports are consistent with the observation that UDP-2,3-diacylglucosamine accumulates in E. coli when genes from lipid biosynthesis are deleted (8). E. coli is also the preferred host for evaluating esterase and lipase activity when screening genes from cultured and uncultured organisms (9, 10). Screening for lipase activity from various microorganisms is often performed on tributyrin agar plates (11). The results are typically the same as for triacetin, and it is repeatedly observed that E. coli does not naturally grow on triesters of glycerol (12, 13). These and many other similar esterase screens (14) would not have been feasible if E. coli produced even moderate levels of a lipase or nonspecific esterase.In the present article we use a combination of our current findings and a thorough review of the relevant literature to conclude that E. coli may not maintain sufficient levels of nonspecific esterase activity to permit the in vivo incorporation of the glycosylated amino acids by the mechanism reported (Fig. 1). Our conclusion is further supported by isothermal calorimetry measurements of Zhang et al. (1) original enzyme showing it retains considerable wild-type activity. We also show that the amino acid GlcNAc-Ser appears to be metabolized in E. coli.  相似文献   

14.
Acetyl esterases from carbohydrate esterase family 7 exhibit unusual substrate specificity. These proteins catalyze the cleavage of disparate acetate esters with high efficiency, but are unreactive to larger acyl groups. The structural basis for this distinct selectivity profile is unknown. Here, we investigate a thermostable acetyl esterase (TM0077) from Thermotoga maritima using evolutionary relationships, structural information, fluorescent kinetic measurements, and site directed mutagenesis. We measured the kinetic and structural determinants for this specificity using a diverse series of small molecule enzyme substrates, including novel fluorogenic esters. These experiments identified two hydrophobic plasticity residues (Pro228, and Ile276) surrounding the nucleophilic serine that impart this specificity of TM0077 for small, straight-chain esters. Substitution of these residues with alanine imparts broader specificity to TM0077 for the hydrolysis of longer and bulkier esters. Our results suggest the specificity of acetyl esterases have been finely tuned by evolution to catalyze the removal of acetate groups from diverse substrates, but can be modified by focused amino acid substitutions to yield enzymes capable of cleaving larger ester functionalities.  相似文献   

15.
Naproxen (nap) is belonging to Non-steriodal anti-inflammatory drugs (NSAIDs) group of drugs that characterized by their free carboxylic group. The therapeutic activity of nap is usually accompanied by GI untoward side effects. Recently synthesized naproxen amides of some amino acid esters prodrugs to mask the free carboxylic group were reported. Those prodrugs showed a promising colorectal cancer chemopreventive activity. The current study aims to investigate the fate and hydrolysis of the prodrugs kinetically in different pH conditions, simulated gastric and intestinal fluids with pHs of 1.2, 5.5 and 7.4 in vitro at 37 °C. The effect of enzymes on the hydrolysis of prodrugs was also studied through incubation of these prodrugs at 37 °C in human plasma and rat liver homogenates. The pharmacokinetic parameters of selected prodrugs and the liberated nap were studied after oral and intraperitoneal administration in male wistar rats. The results showed the hydrolysis of naproxen amides of amino acid esters to nap through two steps first by degradation of the ester moiety to form the amide of nap with amino acid and the second was through the degradation of the amide link to liberate nap. The two reactions were followed and studied kinetically where K1 and K2 (rate constants of degradation) is reported. The hydrolysis of prodrugs was faster in liver homogenates than in plasma. The relative bioavailability of the liberated nap in vivo was higher in case of prodrug containing ethyl glycinate moiety than that occupied l-valine ethyl ester moiety. Each of nap. prodrugs containing ethyl glycinate and l-valine ethyl ester moieties appears promising in liberating nap, decreasing direct GI side effect and consequently their colorectal cancer chemopreventive activity.  相似文献   

16.
Four proteases, crude acid protease from Aspergillus, pronase, amino-peptidase M, and prolidase, have been covalently attached to activated agarose and to amino propyl glass beads. The matrix-bound enzymes have been tested as catalysts for the complete hydrolysis of protein substrates, with the primary goal to isolate unstable amino acid derivatives present in the substrate protein. Under conditions used in the present work, the total amino acid release from the protease-catalyzed hydrolysis of four substrate proteins (pancreatic ribonuclease, egg white lysozyme, yeast enolase, and bovine insulin) was 95–103% of that observed in standard acid hydrolysis. Recovery of individual amino acids showed greater deviation from the theoretical values, but cystine was the only amino acid recovered in low yields (42–77%) from all four proteins. Derivatized amino acids, such as methionine sulfoxide, O-(butylcarbamoyl)-serine, and N-glycosyl asparagine have been obtained from chemically modified proteins or from unmodified glycoprotein in good yield, and normal amino acid constituents of proteins which cannot be quantified after acid hydrolysis (tryptophan, asparagine, and glutamine) have also been determined either directly after proteolysis or after proteolysis in conjunction with acid hydrolysis.  相似文献   

17.
In this paper, we present a method for the determination of low‐ and ultralow copy‐number proteins in biomaterials based on a combination of concentrating the protein from the sample onto cyanogen bromide‐activated Sepharose 4B (via nonspecific binding of free amino groups) and MRM. The detection limit and the dependence of the MRM peak areas on the concentration of protein in the sample were determined using the proteins CYP102 and BSA, as a model system, both in solution and after their addition to human plasma. Nonspecific protein enrichment of proteins from diluted sample volumes of 10–50 mL was found to increase the range of linear dependence of the chromatographic peak area on concentration by more than three orders of magnitude, allowing a lower LOD limit (LLOD) of as low as 10?18 M. At this LLOD, at least two tryptic peptides of CYP102 and BSA could be detected with S/N of ≥7.0. The results were equally good for samples containing pure protein mixtures and proteins spiked into diluted depleted human blood plasma.  相似文献   

18.
The N-terminal segment of the Semliki Forest virus polyprotein is an intramolecular serine protease that cleaves itself off after the invariant Trp267 from a viral polyprotein and generates the mature capsid protein. After this autoproteolytic cleavage, the free carboxylic group of Trp267 interacts with the catalytic triad (His145, Asp167 and Ser219) and inactivates the enzyme. We have deleted the last 1-7 C-terminal residues of the mature capsid protease to investigate whether removal of Trp267 regenerates enzymatic activity. Although the C-terminally truncated polypeptides do not adopt a defined three-dimensional structure and show biophysical properties observed in natively unfolded proteins, they efficiently catalyse the hydrolysis of aromatic amino acid esters, with higher catalytic efficiency for tryptophan compared to tyrosine esters and kcat/KM values up to 5 × 105 s−1 M−1. The enzymatic mechanism of these deletion variants is typical of serine proteases. The pH enzyme activity profile shows a pKa1 = 6.9, and the Ser219Ala substitution destroys the enzymatic activity. In addition, the fast release of the first product of the enzymatic reaction is followed by a steady-state second phase, indicative of formation and breakdown of a covalent acyl-enzyme intermediate. The rates of acylation and deacylation are k2 = 4.4±0.6 s−1 and k3 = 1.6±0.5 s−1, respectively, for a tyrosine derivative ester substrate, and the amplitude of the burst phase indicates that 95% of the enzyme molecules are active. In summary, our data provide further evidence for the potential catalytic activity of natively unfolded proteins, and provide the basis for engineering of alphavirus capsid proteins towards hydrolytic enzymes with novel specificities.  相似文献   

19.
Porcine pancreatic lipase (PPL)-catalyzed enantioselective hydrolysis of N-benzyloxycarbonyl-dl-amino acid esters (Z-dl-AA-ORs) was studied for the optical resolution of a variety of non-protein amino acids. The ester moiety (R) of the substrate affected the rate of hydrolysis significantly. The glyceryl (Gl) and carbamoylmethyl (Cam) esters were found to be highly reactive substrates. The hydrolysis of the Gl esters (Z-dl-AA-OGls) of both aliphatic and aromatic amino acids was examined in acetonitrile containing 70% (v/v) of 0.02 M phosphate buffer (pH 7.0) at 30°C. With all amino acids tested, the corresponding l-enantiomers were hydrolyzed preferentially. PPL favored aromatic amino acids, such as phenylalanine and p-chlorophenylalanine, leading to completion of the hydrolysis within 20 min with excellent enantioselectivities (E>100). The PPL-catalyzed hydrolysis of the corresponding Cam esters (Z-dl-AA-OCams) was also examined under the same reaction conditions. Although the hydrolysis of the Cam esters was rapid, the l-enantioselectivities were rather poor with aromatic amino acids, such as 2-phenylglycine and homophenylalanine.  相似文献   

20.
A C-N bond in one arm of the mixed-valence VIII-VIV complex bpbp(VOCl2)(VCl2), bpbpH = 2,6-bis((N,N-bis-(2-picolyl)amino)methyl)-4-tertbutylphenol, is cleaved in wet acetonitrile solution to give bpa(VOCl2), bpa = bis(2-methypyridyl)amine, and 2-((N,N-bis-(2-picolyl)amino)methyl)-6-hydroxymethyl-4-tertbutylphenol. The reaction corresponds overall to hydrolysis of a tertiary amine to form a secondary amine and a primary alcohol. The structure of bpa(VOCl2) was established by X-ray diffraction while 2-((N,N-bis-(2-picolyl)amino)methyl)-6-hydroxymethyl-4-tertbutylphenol was detected by ESI mass spectrometry. The phenol oxygen atom in bpbp(VOCl2)(VCl2) is proposed to be non-bridging and this asymmetry is likely to be important for the C-N bond cleavage reaction. A related asymmetrical VIV complex, [bpbpH(VO)(H2O)](ClO4)2 ·  H2O, containing bpbp bound to only one metal ion, has also been characterized by X-ray diffraction. In slightly more basic solution, bpbp(VOCl2)(VCl2) is oxidized to the VIV-VIV complex [bpbp(VOCl)2]+ and C-N bond cleavage is suppressed.  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