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 46 毫秒
1.
The title compounds were synthesized by 1,3‐dipolar cycloaddition of 3,3,3‐trifluoropropinyl benzene ( 2 ) to the azido sugars 2,3,4,6‐tetra‐O‐acetyl‐β‐D ‐galactopyranosyl azide ( 1 ), 6‐O‐acetyl‐4‐O‐cyclohexylcarbamoyl‐2,3‐O‐(2,2,2‐trichloroethylidene)‐β‐D ‐gulopyranosyl azide ( 6 ), 6‐azido‐6‐deoxy‐1,2:3,4‐di‐O‐isopropylidene‐α‐D ‐galactopyranose ( 12 ), and methyl 6‐azido‐4‐O‐cyclohexylcarbamoyl‐6‐deoxy‐2,3‐O‐(2,2,2‐trichloroethylidene)‐β‐D ‐gulopyranoside ( 16 ), respectively. Because of the dissymmetry of the dipolarophile 2 , always two regioisomeric products were obtained, the nucleoside‐analogous compounds 3/4 (from 1 ) and 7/8 (from 6 ), respectively, and the reversed nucleosides 13/14 (from 12 ) and 17/18 (from 16 ), respectively. Protecting group chemistry like transesterification, deacetalation, hydrodechlorination is demonstrated in some cases. Thus, the trichloroethylidene derivatives 7, 8, 17, and 18 were converted into the corresponding ethylidene derivatives ( 9, 10, 19, 20 ) by treatment with tributylstannane/AIBN. An X‐ray analysis is given for the 1‐(2,3,4,6‐tetra‐O‐acetyl‐β‐D ‐galactopyranosyl)‐4‐trifluoromethyl‐5‐phenyl‐1,2,3‐triazole ( 4 ) and for the 1‐[6‐O‐acetyl‐4‐O‐cyclohexylcarbamoyl‐2,3‐O‐(2,2,2‐trichloroethylidene)‐β‐D ‐gulopyranosyl]‐4‐trifluoromethyl‐5‐phenyl‐1,2,3‐triazole ( 7 ).  相似文献   

2.
The biocatalytic synthesis and purification of O‐β‐D ‐monoglucuronide conjugates of hydroxytyrosol, tyrosol, homovanillic alcohol, and 3‐(4′‐hydroxyphenyl)propanol, using porcine liver microsomes, are described here. The glucuronides were synthesized, analyzed and separated by HPLC‐UV, identified by HPLC‐MS, and their structures unequivocally established by NMR techniques. The outcome of the glucuronidation reaction depends on the structure of the phenolic compounds. Thus, the glucuronidation of hydroxytyrosol, biocatalyzed with porcine liver microsomes, proceeded exclusively on the phenolic hydroxy groups. The regioselectivity was similar to that observed for human and rat liver microsomes, the 4′‐hydroxy position being more favorable than the 3′‐hydroxy one. In the case of tyrosol, homovanillic alcohol, and hydroxyphenylpropanol, two products were formed during microsomal glucuronidation: a major one, the phenolic O‐β‐D ‐glucuronidated derivative and, a minor one, the O‐β‐D ‐glucuronidated aliphatic alcohol.  相似文献   

3.
A graft copolymer of poly(vinylidene fluoride) (PVDF) with a glucose‐carrying methacrylate, 3‐O‐methacryloyl‐1,2:5,6‐di‐O‐isopropylidene‐D ‐glucofuranose, was synthesized via the atom transfer radical polymerization technique with commercial PVDF as the macroinitiator. After a treatment with 88% formic acid, the isopropylidenyl groups of the precursor graft copolymer [poly(vinylidene fluoride)‐g‐poly(3‐O‐methacryloyl‐1,2:5,6‐di‐O‐isopropylidene‐ D ‐glucofuranose)] were converted into hydroxyl groups, and this produced an amphiphilic graft copolymer (PVDF‐g‐PMAG) [poly(vinylidene fluoride)‐g‐poly(3‐O‐methacryloyl‐α,β‐D‐glucopyranose)] with glycopolymer side chains and a narrow molecular weight distribution (weight‐average molecular weight/number‐average molecular weight < 1.29). This glucose‐carrying graft copolymer was characterized with Fourier transform infrared, proton nuclear magnetic resonance, gel permeation chromatography, and thermogravimetric analysis. A novel porous membrane prepared from blends of PVDF with PVDF‐g‐PMAG via an immersion–precipitation technique exhibited significantly enhanced hydrophilicity and an anti‐protein‐adsorption property. The surface chemical composition and morphology of the membrane were studied with X‐ray photoelectron spectroscopy and scanning electron microscopy, respectively. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

4.
Functional cationic polymers carrying mannose moieties were synthesized in a facile manner by employing RAFT polymerization. Initially, a protected carbohydrate based monomer, [2‐(2,3,4,6‐tetra‐O‐acetyl‐α‐d ‐mannopyranosyloxy)ethyl methacrylate (AcManEMA)], was prepared by the O‐glycosylation of 2‐hydroxyethyl methacrylate (HEMA). Subsequently, a macroRAFT agent of poly[2‐(dimethyl)amino ethyl methacrylate] (PDMAEMA) was generated, and a further chain extension polymerization with AcManEMA was carried out in dioxane to form a acetylated mannose cationic diblock copolymer, PDMAEMA‐b‐PAcManEMA. It was attained in high yields and displayed low dispersity (Ð). Acetylated mannose moieties on the polymer were deprotected with sodium methoxide and the amines from the DMAEMA block were protonated to yield a cationic diblock glycopolymer, PDMAEMA‐b‐PManEMA. The cationic property of polymers were characterized by mixing with a negatively charged siRNA duplex and a pDNA, and aggregates of 102 and 233 nm were obtained, respectively. Agarose gel shift assay revealed that the polymers were able to retain the nucleic acids as large polymer complexes. Lectin binding assay proved that the mannose residue on the polymers were only able to bind specifically with ConA. PNA lectin was employed as a control and did not show specific binding. The cationic glycopolymer could be advantageous in targeted nucleic acids delivery in specific cells. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134, 44947.  相似文献   

5.
A collection of 4‐(C‐galactosyl)‐ and 4‐(C‐ribosyl)‐β‐lactams featuring different substituents at C‐3 and N‐1 was prepared by combining in a one‐pot procedure a formyl C‐glycoside, a primary amine, and a substituted acetyl chloride in the presence of base (Staudinger‐type reaction). Sulfonyl chloride and aminomethylated resins were used in sequence to remove excess of components and by‐products. Two pure C‐glycosyl‐β‐lactams were effectively transformed into C‐glycosyl‐N‐Boc‐β‐amino‐α‐hydroxy esters (C‐glycosyl isoserines) and a C‐ribosyl dipeptide via base‐promoted heterocycle ring opening by methanol and L ‐phenylalanine methyl ester, respectively.  相似文献   

6.
The enzymatic epimerization of uridine 5′‐diphospho‐α‐D ‐glucose (UDP‐Glc, 1 ) and uridine 5′‐diphospho‐N‐acetyl‐α‐D ‐glucosamine (UDP‐GlcNAc, 2 ) and the subsequent oxidation of uridine 5′‐diphospho‐α‐D ‐galactose (UDP‐Gal, 3 ) and uridine 5′‐diphospho‐N‐acetyl‐α‐D ‐galactosamine (UDP‐GalNAc, 4 ) were combined with chemical biotinylation with biotin‐ε‐amidocaproylhydrazide in a one‐pot synthesis. Analysis by CE and NMR revealed a mixture (1.0:1.4) of the biotinylated nucleotide sugars uridine 5′‐diphospho‐6‐biotin‐ε‐amidocaproylhydrazino‐α‐D ‐galactose (UDP‐6‐biotinyl‐Gal, 7) and uridine 5′‐diphospho‐6‐biotin‐ε‐amidocaproylhydrazino‐α‐D ‐glucose (UDP‐6‐biotinyl‐Glc, 9 ), respectively, in a reaction started with 1 . One product, uridine 5′‐diphospho‐6‐biotin‐ε‐amidocaproylhydrazino‐N‐acetyl‐α‐D ‐galactosamine (UDP‐6‐biotinyl‐GalNAc, 8) was formed when the reaction was initiated with 2 . It could be demonstrated for the first time that a UDP‐Glc(NAc) 4′‐epimerase (Gne from Campylobacter jejuni) and galactose oxidase from Dactylium dendroides can be used simultaneously in enzymatic catalysis. This is of particular interest since the coaction of an enzyme demanding reductive conditions and an oxygen‐dependent oxidase is unexpected.  相似文献   

7.
A fusion protein composed of β1,3‐N‐acetyl‐D ‐glucosaminyltransferase (β1,3‐GlcNAcT) from Streptococcus agalactiae type Ia and maltose‐binding protein (MBP) was produced in Escherichia coli as a soluble and highly active form. Although this fusion protein (MBP‐β1,3‐GlcNAcT) did not show any sugar‐elongation activity to some simple low‐molecular weight acceptor substrates such as galactose, Galβ(1→4)Glc (lactose), Galβ(1→4)GlcNAc (N‐acetyllactosamine), Galβ(1→4)GlcNAcβ(1→3)Galβ(1→4)Glc (lacto‐N‐tetraose), and Galβ(1→4)GlcβCer (lactosylceramide, LacCer), the multivalent glycopolymer having LacCer‐mimic branches (LacCer mimic polymer, LacCer primer) was found to be an excellent acceptor substrate for the introduction of a β‐GlcNAc residue at the O‐3 position of the non‐reducing galactose moiety by this engineered enzyme. Subsequently, the polymer having GlcNAcβ(1→3)Galβ(1→4)Glc was subjected to further enzymatic modifications by using recombinant β1,4‐D ‐galactosyltransferase (β1,4‐GalT), α2,3‐sialyltransferase (α2,3‐SiaT), α1,3‐L ‐fucosyltransferase (α1,3‐FucT), and ceramide glycanase (CGase) to afford a biologically important ganglioside; Neu5Aα(2→3)Galβ(1→4)[Fucα(1→3)]GlcNAcβ(1→3)Galβ(1→4)GlcCerα(IV3Neu5Acα,III3Fucα‐nLc4Cer) in 40% yield (4 steps). Interestingly, it was suggested that MBP‐β1,3‐GlcNAcT could also catalyze a glycosylation reaction of the LacCer mimic polymer with N‐acetyl‐D ‐galactosamine served from UDP‐GalNAc to afford a polymer carrying trisaccharide branches, GalNAcβ(1→3)Galβ(1→4)Glc. The versatility of the MBP‐β1,3‐GlcNAcT in the practical synthesis was preliminarily demonstrated by applying this fusion protein as an immobilized biocatalyst displayed on the amylose resin which is known as a solid support showing potent binding‐affinity with MBP.  相似文献   

8.
N‐Acetyl‐D ‐neuraminic acid (Neu5Ac) was efficiently synthesized from lactate and a mixture of N‐acetyl‐D ‐glucosamine (GlcNAc) and N‐acetyl‐D ‐mannosamine (ManNAc) by whole cells. The biotransformation utilized Escherichia coli cells (Neu5Ac aldolase), Pseudomonas stutzeri cells (lactate oxidase components), GlcNAc/ManNAc and lactate. By this process, 18.32±0.56 g/liter Neu5Ac were obtained from 65.61±2.70 g/liter lactate as an initial substrate input. Neu5Ac (98.4±0.4 % purity, 80.87±0.79 % recovery yield) was purified by anionic exchange chromatography. Our results demonstrate that the reported Neu5Ac biosynthetic process can compare favorably with natural product extraction or chemical synthesis processes.  相似文献   

9.
The phytochemical resveratrol (trans‐3,5,4′‐trihydroxystilbene) has drawn great interest as health‐promoting food ingredient and potential therapeutic agent. However, resveratrol shows vanishingly low water solubility; this limits its uptake and complicates the development of effective therapeutic forms. Glycosylation should be useful to enhance resveratrol solubility, with the caveat that unselective attachment of sugars could destroy the molecule's antioxidant activity. UGT71A15 (a uridine 5′‐diphosphate α‐D ‐glucose‐dependent glucosyltransferase from apple) was used to synthesize resveratrol 3,5‐β‐D ‐diglucoside; this was about 1700‐fold more water‐soluble than the unglucosylated molecule (~0.18 mM ), yet retained most of the antioxidant activity. Resveratrol 3‐β‐D ‐glucoside, which is the naturally abundant form of resveratrol, was a practical substrate for perfect site‐selective conversion into the target diglucoside in quantitative yield (g L?1 concentration).  相似文献   

10.
Matrix metalloproteinase‐12 (MMP‐12) can be considered an attractive target to study selective inhibitors useful in the development of new therapies for lung and cardiovascular diseases. In this study, a new series of arylsulfonamide carboxylates, with increased hydrophilicity resulting from conjugation with a β‐N‐acetyl‐d ‐glucosamine moiety, were designed and synthesized as MMP‐12 selective inhibitors. Their inhibitory activity was evaluated on human MMPs by using the fluorimetric assay, and a crystallographic analysis was performed to characterize their binding mode. Among these glycoconjugates, a nanomolar MMP‐12 inhibitor with improved water solubility, compound 3 [(R)‐2‐(N‐(2‐(3‐(2‐acetamido‐2‐deoxy‐β‐d ‐glucopyranosyl)thioureido)ethyl)biphenyl‐4‐ylsulfonamido)‐3‐methylbutanoic acid], was identified.  相似文献   

11.
The mycobacterial cell wall is a complex architecture, which has, as its major structural component, a lipidated polysaccharide covalently bound to peptidoglycan. This structure, termed the mycolyl–arabinogalactan–peptidoglycan complex, possesses a core galactan moiety composed of approximately 30 galactofuranosyl (Galf) resides attached via alternating β‐(1→6) and β‐(1→5) linkages. Recent studies have shown that the entire galactan is synthesized by the action of only two bifunctional galactofuranosyltransferases, GlfT1 and GlfT2. We report here saturation‐transfer difference (STD) NMR spectroscopy studies with GlfT2 using two trisaccharide acceptor substrates, β‐D ‐Galf‐(1→6)‐β‐D ‐Galf‐(1→5)‐β‐D ‐Galf‐O(CH2)7CH3 ( 2 ) and β‐D ‐Galf‐(1→5)‐β‐D ‐Galf‐(1→6)‐β‐D ‐Galf‐O(CH2)7CH3 ( 3 ), as well as the donor substrate for the enzyme, UDP‐Galf. Competition STD‐NMR titration experiments and saturation transfer double difference (STDD) experiments with 2 and 3 were undertaken to explore the bifunctionality of this enzyme, in particular to answer whether one or two active sites are responsible for the formation of both β‐(1→5)‐ and β‐(1→6)‐Galf linkages. It was demonstrated that 2 and 3 bind competitively at the same site; this suggests that GlfT2 has one active site pocket capable of catalyzing both β‐(1→5) and β‐(1→6) galactofuranosyl transfer reactions. The addition of UDP‐Galf to GlfT2 in the presence of either 2 or 3 generated a tetrasaccharide product; this indicates that the enzyme was catalytically active under the conditions at which the STD‐NMR experiments were carried out.  相似文献   

12.
The unique five‐membered aminocyclitol core of the antitumor antibiotic pactamycin originates from d ‐glucose, so unprecedented enzymatic modifications of the sugar intermediate are involved in the biosynthesis. However, the order of the modification reactions remains elusive. Herein, we examined the timing of introduction of an amino group into certain sugar‐derived intermediates by using recombinant enzymes that were encoded in the pactamycin biosynthesis gene cluster. We found that the NAD+‐dependent alcohol dehydrogenase PctP and pyridoxal 5′‐phosphate dependent aminotransferase PctC converted N‐acetyl‐d ‐glucosaminyl‐3‐aminoacetophonone into 3′‐amino‐3′‐deoxy‐N‐acetyl‐d ‐glucosaminyl‐3‐aminoacetophenone. Further, N‐acetyl‐d ‐glucosaminyl‐3‐aminophenyl‐β‐oxopropanoic acid ethyl ester was converted into the corresponding 3′‐amino derivative. However, PctP did not oxidize most of the tested d ‐glucose derivatives, including UDP‐GlcNAc. Thus, modification of the GlcNAc moiety in pactamycin biosynthesis appears to occur after the glycosylation of aniline derivatives.  相似文献   

13.
BACKGROUND: 1,3‐1,4‐β‐D‐glucanase (1,3‐1,4‐β‐D‐glucan 4‐glucanohydrolase; EC 3.2.1.73) has been used in a range of industrial processes. As a biocatalyst, it is better to use immobilized enzymes than free enzymes, therefore, the immobilization of 1,3‐1,4‐β‐D‐glucanase was investigated. RESULTS: A 1,3‐1,4‐β‐D‐glucanase gene from Fibrobacter succinogenes was overexpressed in Escherichia coli as a recombinant protein fused to the N terminus of oleosin, a unique structural protein of seed oil bodies. With the reconstitution of the artificial oil bodies (AOBs), refolding, purification, and immobilization of active 1,3‐1,4‐β‐D‐glucanase was accomplished simultaneously. Response surface modeling (RSM), with central composite design (CCD), and regression analysis were successfully applied to determine the optimal temperature and pH conditions of the AOB‐immobilized 1,3‐1,4‐β‐D‐glucanase. The optimal conditions for the highest immobilized 1,3‐1,4‐β‐D‐glucanase activity (7.1 IU mg?1 of total protein) were observed at 39 °C and pH 8.8. Furthermore, AOB‐immobilized 1,3‐1,4‐β‐D‐glucanase retained more than 70% of its initial activity after 120 min at 39 °C, and it was easily and simply recovered from the surface of the solution by brief centrifugation; it could be reused eight times while retaining more than 80% of its activity. CONCLUSIONS: These results indicate that the AOB‐based system is a comparatively simple and effective method for simultaneous refolding, purification, and immobilization of 1,3‐1,4‐β‐D‐glucanase. Copyright © 2009 Society of Chemical Industry  相似文献   

14.
A series of optically active polyamides containing di‐O‐methyl‐L ‐tartaryl moieties in the main chain were synthesized by polycondensation of di‐O‐methyl‐L ‐tartaryl chloride 5 with diamines and characterized by gel permeation chromatography, UV–vis, circular dichroism (CD), IR, and NMR spectroscopies. The polycondensation reaction could be carried out under mild conditions and the reaction time was short (2–3 h). The key monomer 5 prepared from L ‐tartaric acid via esterification, etherification, hydrolysis, and chlorination was easily purified by vacuum sublimation. These polyamides with number average molecular weights ranging from 14,000 to 35,000, displayed large optical activity in dimethyl sulfoxide solution, and their specific optical rotations oscillated between 87.2° and 210.7° depending on the structures of the diamines. The glass transition temperatures of these polyamides were in the range of 106–191°C, and the 10% mass loss occurred at temperature above 300°C. The polyamides derived from aromatic diamines exhibited higher Tg and thermal stability than those derived from aliphatic diamines. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

15.
D ‐glucose and D ‐galactose end‐functionalized polylactide oligomers were synthesized by controlled ring‐opening polymerization of lactide using aluminium triisopropoxide, triethylaluminium or stannous octoate as promoter. Accordingly, two selectively protected monosaccharides were studied as co‐initiators, either 1,2;5,6‐di‐O‐isopropylidene‐α‐D ‐glucofuranose (1) and 1,2;3,4‐di‐O‐isopropylidene‐α‐D ‐galactopyranose (2). In contrast to what is known in polymerization of ?‐caprolactone, both protected monosaccharides proved to be efficient co‐initiators and yielded end‐functionalized polylactide chains with controlled regioselectivity (C‐3 or C‐6 linkage), predictable molecular weights and narrow molecular weight distributions. © 2003 Society of Chemical Industry  相似文献   

16.
Multi‐responsive hydrogels have recently received considerable attention for bioapplications. Here, novel temperature‐ and redox‐responsive polypetide hydrogels have been developed. Thermo‐sensitive hydrogels based on poly(ethyleneglycol)‐block ‐poly(γ‐propargyl‐l ‐glutamate) (PEG‐PPLG ) were first synthesized by the ring opening polymerization of γ‐propargyl‐l ‐glutamate N ‐carboxyanhydride (PLG‐NCA ) with amino group terminated PEG monomethyl ether (mPEG‐NH2 ) as macroinitiator and were then functionalized via the ‘thiol‐yne’ click reaction between the propargyl pendents and the thiol‐containing 1‐propanethiol. The sol ? gel phase transition of the obtained copolymer aqueous solution in response to temperature change was studied. The mass loss of the hydrogel in vitro was accelerated in the presence of H2O2 , exhibiting a redox‐responsive property. Further, the methyl thiazolyl tetrazolium viability results revealed that this polypetide hydrogel has excellent biocompatibility, presenting potential applications in the biomedical field. © 2016 Society of Chemical Industry  相似文献   

17.
The synthesis and detailed characterization of racemic 3‐methyl‐1,4‐dioxan‐2‐one (3‐MeDX) are reported. The bulk ring‐opening polymerization of 3‐MeDX, to yield a poly(ester‐ether) meant for biomedical applications, in the presence of various initiators such as tin(II) octanoate, tin(II) octanoate/n‐butyl alcohol, aluminium tris‐isopropoxide and an aluminium Schiff base complex (HAPENAlOiPr) under varying experimental conditions is here detailed for the first time. Polymerization kinetics were investigated and compared with those of 1,4‐dioxan‐2‐one. The studies reveal that the rate of polymerization of 3‐MeDX is less than that of 1,4‐dioxan‐2‐one. Experimental conditions to achieve relatively high molar masses have been established. Thermodynamic parameters such as enthalpy and entropy of 3‐MeDX polymerization as well as ceiling temperature have been determined. Poly(D ,L ‐3‐MeDX) is found to possess a much lower ceiling temperature than poly(1,4‐dioxan‐2‐one). Poly(D ,L ‐3‐MeDX) was characterized using NMR spectroscopy, matrix‐assisted laser desorption ionization mass spectrometry, size exclusion chromatography and differential scanning calorimetry. This polymer is an amorphous material with a glass transition temperature of about ?20 °C. Copyright © 2010 Society of Chemical Industry  相似文献   

18.
An expression system based on Escherichia coli and the T5 promoter allowed the overproduction of a his‐tagged rhamnulose‐1‐phosphate aldolase (RhuA; EC 4.1.2.19), an enzyme with applications in the production of deoxyazasugars and deoxysugars compounds. Shake flask and bioreactor cultivation with E coli M15 (pQErham) were performed under different media and inducing conditions for RhuA expression. A Defined Medium (DM) with glucose as carbon source gave a high volumetric and enzyme productivity (3460 AU dm?3 and 288 AU dm?3 h?1 respectively) compared with Luria–Bertoni (LB) medium (2292 AU dm? 3 and 255 AU dm?3 h?1). The minimum quantity of (isopropyl‐β‐D ‐thiogalactoside) IPTG for optimal induction was estimated in 18–20 µmol IPTG gDCW?1. The highest volumetric production of RhuA (8333 AU dm?3) was obtained when IPTG was added in the late log‐phase. No significant differences were found in specific RhuA activity for induction temperatures of 30 and 37 °C. An effective two‐step purification process comprising affinity chromatography and gel permeation has been developed (overall recovery 66.5%). These studies provide the basis for the further development of an integrated process for recombinant RhuA production suitable for biotransformation applications. Copyright © 2003 Society of Chemical Industry  相似文献   

19.
Two series of polymers with low‐molecular‐weight L ‐lactide side‐arms were prepared by ring‐opening polymerization using D ‐glucosamine and N‐acetyl‐D ‐glucosamine as polymer core molecules, and methanesulfonic acid as solvent and catalyst. This simple synthetic route does not rely on the use of organometallic catalysts, and has also proven useful to the authors for chitosan grafting with lactones. NMR spectroscopy reveals a high degree of substitution (>3), and Fourier transform infrared/NMR spectra suggest the existence of three different lactate tautomers likely to be responsible for coloration. These D ‐glucosamine‐based polymers also display glass transition temperatures approximately 10 °C above that of the human body, which points to the potential of these lactone‐grafted aminosugars with tunable amphiphilic character in the design of submicrometer‐sized drug delivery vehicles. They are also viewed as interesting hydrophobic chelating agents for catalytic transition metal centers. Copyright © 2010 Society of Chemical Industry  相似文献   

20.
The putative hydrolase gene bhp from the balhimycin biosynthetic gene cluster has been cloned and overexpressed in Escherichia coli. The corresponding enzyme Bhp was purified to homogeneity by nickel‐chelating chromatography and characterized. Although Bhp has sequence similarities to hydrolases with “haloperoxidase”/perhydrolase activity, it did not show any enzymatic activity with standard “haloperoxidase”/perhydrolase substrates (e.g., monochlorodimedone and phenol red), nonspecific esterase substrates (such as p‐nitrophenyl acetate, p‐nitrophenyl phosphate and S‐thiophenyl acetate) or the model lactonase substrate dihydrocoumarin. However, Bhp could be shown to catalyse the hydrolysis of S‐β‐hydroxytyrosyl‐N‐acetyl cysteamine thioester (β‐OH‐Tyr‐SNAC) with 15 times the efficiency of S‐L ‐tyrosyl‐N‐acetyl cysteamine thioester (L ‐Tyr‐SNAC). This is in agreement with the suggestion that Bhp is involved in balhimycin biosynthesis, during which it was supposed to catalyse the hydrolysis of β‐OH‐Tyr‐S‐PCP (PCP=peptidyl carrier protein) to free β‐hydroxytyrosine (β‐OH‐Tyr) and strongly suggests that Bhp is a thioesterase with high substrate specificity for PCP‐bound β‐OH‐Tyr and not a “haloperoxidase”/perhydrolase or nonspecific esterase.  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