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 15 毫秒
1.
Short‐chain dehydrogenase/reductase (SDR) is distributed in many organisms, from bacteria to humans, and has significant roles in metabolism of carbohydrates, lipids, amino acids, and other biomolecules. An important intermediate in acidic polysaccharide metabolism is 2‐keto‐3‐deoxy‐d ‐gluconate (KDG). Recently, two short and long loops in Sphingomonas KDG‐producing SDR enzymes (NADPH‐dependent A1‐R and NADH‐dependent A1‐R′) involved in alginate metabolism were shown to be crucial for NADPH or NADH coenzyme specificity. Two SDR family enzymes—KduD from Pectobacterium carotovorum (PcaKduD) and DhuD from Streptococcus pyogenes (SpyDhuD)—prefer NADH as coenzyme, although only PcaKduD can utilize both NADPH and NADH. Both enzymes reduce 2,5‐diketo‐3‐deoxy‐d ‐gluconate to produce KDG. Tertiary and quaternary structures of SpyDhuD and PcaKduD and its complex with NADH were determined at high resolution (approximately 1.6 Å) by X‐ray crystallography. Both PcaKduD and SpyDhuD consist of a three‐layered structure, α/β/α, with a coenzyme‐binding site in the Rossmann fold; similar to enzymes A1‐R and A1‐R′, both arrange the two short and long loops close to the coenzyme‐binding site. The primary structures of the two loops in PcaKduD and SpyDhuD were similar to those in A1‐R′ but not A1‐R. Charge neutrality and moderate space at the binding site of the nucleoside ribose 2′ coenzyme region were determined to be structurally crucial for dual‐coenzyme specificity in PcaKduD by structural comparison of the NADH‐ and NADPH‐specific SDR enzymes. The corresponding site in SpyDhuD was negatively charged and spatially shallow. This is the first reported study on structural determinants in SDR family KduD related to dual‐coenzyme specificity. Proteins 2016; 84:934–947. © 2016 Wiley Periodicals, Inc.  相似文献   

2.
Aims: To characterize of a thermostable recombinant α‐l ‐arabinofuranosidase from Caldicellulosiruptor saccharolyticus for the hydrolysis of arabino‐oligosaccharides to l ‐arabinose. Methods and Results: A recombinant α‐l ‐arabinofuranosidase from C. saccharolyticus was purified by heat treatment and Hi‐Trap anion exchange chromatography with a specific activity of 28·2 U mg?1. The native enzyme was a 58‐kDa octamer with a molecular mass of 460 kDa, as measured by gel filtration. The catalytic residues and consensus sequences of the glycoside hydrolase 51 family of α‐l ‐arabinofuranosidases were completely conserved in α‐l ‐arabinofuranosidase from C. saccharolyticus. The maximum enzyme activity was observed at pH 5·5 and 80°C with a half‐life of 49 h at 75°C. Among aryl‐glycoside substrates, the enzyme displayed activity only for p‐nitrophenyl‐α‐l ‐arabinofuranoside [maximum kcat/Km of 220 m(mol l?1)?1 s?1] and p‐nitrophenyl‐α‐l ‐arabinopyranoside. This substrate specificity differs from those of other α‐l ‐arabinofuranosidases. In a 1 mmol l?1 solution of each sugar, arabino‐oligosaccharides with 2–5 monomer units were completely hydrolysed to l ‐arabinose within 13 h in the presence of 30 U ml?1 of enzyme at 75°C. Conclusions: The novel substrate specificity and hydrolytic properties for arabino‐oligosaccharides of α‐l ‐arabinofuranosidase from C. saccharolyticus demonstrate the potential in the commercial production of l ‐arabinose in concert with endoarabinanase and/or xylanase. Significance and Impact of the Study: The findings of this work contribute to the knowledge of hydrolytic properties for arabino‐oligosaccharides performed by thermostable α‐l ‐arabinofuranosidase.  相似文献   

3.
4.
Histatin‐5 (Hst‐5, DSHAKRHHGYKRKFHEKHHSHRGY) is a member of a histidine‐rich peptide family secreted by major salivary glands, exhibiting high fungicidal activity against Candida albicans. In the present work, we demonstrate the 3D structure of the head‐to‐tail cyclic variant of Hst‐5 in TFE solution determined using NMR spectroscopy and molecular dynamics simulations. The cyclic histatin‐5 reveals a helix‐loop‐helix motif with α‐helices at positions Ala4‐His7 and Lys11‐Ser20. Both helical segments are arranged relative to each other at an angle of ca. 142°. The head‐to‐tail cyclization increases amphipathicity of the peptide, this, however, does not affect its antimicrobial potency. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
Bacillus thuringiensis is a soil‐dwelling Gram positive bacterium that has been utilized as a biopesticide for well over 60 years. It is known to contain flagella that are important for motility. One of the proteins found in flagella is flagellin, which is post‐translationally modified by O‐glycosylation with derivatives of pseudaminic acid. The biosynthetic pathway for the production of CMP‐pseudaminic acid in B. thuringiensis, starting with UDP‐N‐acetyl‐d ‐glucosamine (UDP‐GlcNAc), requires seven enzymes. Here, we report the three‐dimensional structures of Pen and Pal, which catalyze the first and second steps, respectively. Pen contains a tightly bound NADP(H) cofactor whereas Pal is isolated with bound NAD(H). For the X‐ray analysis of Pen, the site‐directed D128N/K129A mutant variant was prepared in order to trap its substrate, UDP‐GlcNAc, into the active site. Pen adopts a hexameric quaternary structure with each subunit showing the bilobal architecture observed for members of the short‐chain dehydrogenase/reductase superfamily. The hexameric quaternary structure is atypical for most members of the superfamily. The structure of Pal was determined in the presence of UDP. Pal adopts the more typical dimeric quaternary structure. Taken together, Pen and Pal catalyze the conversion of UDP‐GlcNAc to UDP‐4‐keto‐6‐deoxy‐l ‐N‐acetylaltrosamine. Strikingly, in Gram negative bacteria such as Campylobacter jejuni and Helicobacter pylori, only a single enzyme (FlaA1) is required for the production of UDP‐4‐keto‐6‐deoxy‐l ‐N‐acetylaltrosamine. A comparison of Pen and Pal with FlaA1 reveals differences that may explain why FlaA1 is a bifunctional enzyme whereas Pen and Pal catalyze the individual steps leading to the formation of the UDP‐sugar product. This investigation represents the first structural analysis of the enzymes in B. thuringiensis that are required for CMP‐pseudaminic acid formation.  相似文献   

6.
Aims: Characterization of substrate specificity of a d ‐lyxose isomerase from Serratia proteamaculans and application of the enzyme in the production of d ‐lyxose and d ‐mannose. Methods and Results: The concentrations of monosaccharides were determined using a Bio‐LC system. The activity of the recombinant protein from Ser. proteamaculans was the highest for d ‐lyxose among aldoses, indicating that it is a d‐ lyxose isomerase. The native recombinant enzyme existed as a 54‐kDa dimer, and the maximal activity for d‐ lyxose isomerization was observed at pH 7·5 and 40°C in the presence of 1 mmol l?1 Mn2+. The Km values for d ‐lyxose, d ‐mannose, d ‐xylulose, and d ‐fructose were 13·3, 32·2, 3·83, and 19·4 mmol l?1, respectively. In 2 ml of reaction volume at pH 7·5 and 35°C, d ‐lyxose was produced at 35% (w/v) from 50% (w/v) d ‐xylulose by the d‐ lyxose isomerase in 3 h, while d ‐mannose were produced at 10% (w/v) from 50% (w/v) d ‐fructose in 5 h. Conclusions: We identified the putative sugar isomerase from Ser. proteamaculans as a d ‐lyxose isomerase. The enzyme exhibited isomerization activity for aldose substrates with the C2 and C3 hydroxyl groups in the left‐hand configuration. High production rates of d‐ lyxose and d ‐mannose by the enzyme were obtained. Significance and Impact of the Study: A new d‐ lyxose isomerase was found, and this enzyme had higher activity for d ‐lyxose and d ‐mannose than previously reported enzymes. Thus, the enzyme can be applied in industrial production of d ‐lyxose and d ‐mannose.  相似文献   

7.
Bypass of the penicillin‐binding proteins by an l ,d ‐transpeptidase (Ldtfm) confers cross‐resistance to β‐lactam and glycopeptide antibiotics in mutants of Enterococcus faecium selected in vitro. Ldtfm is produced by the parental strain D344S although it insignificantly contributes to peptidoglycan cross‐linking as pentapeptide stems cannot be used as acyl donors by this enzyme. Here we show that production of the tetrapeptide substrate of Ldtfm is controlled by a two‐component regulatory system (DdcRS) and a metallo‐d ,d ‐carboxypeptidase (DdcY). The locus was silent in D344S and its activation was due to amino acid substitutions in DdcS or DdcR that led to production of DdcY and hydrolysis of the C‐terminal d ‐Ala residue of the cytoplasmic peptidoglycan precursor UDP‐MurNAc‐pentapeptide. The T161A and T161M substitutions affected a position of DdcS known to be essential for the phosphatase activity of related sensor kinases. Complete elimination of UDP‐MurNAc‐pentapeptide, which was required specifically for resistance to glycopeptides, involved substitutions in DdcY that increased the catalytic efficiency of the enzyme (E127K) and affected its interaction with the cell envelope (I14N). The ddc locus displays striking similarities with portions of the van vancomycin resistance gene clusters, suggesting possible routes of emergence of cross‐resistance to glycopeptides and β‐lactams in natural conditions.  相似文献   

8.
Methylated inositol, d ‐pinitol (3‐O‐methyl‐d ‐chiro‐inositol), is a common constituent in legumes. It is synthesized from myo‐inositol in two reactions: the first reaction, catalyzed by myo‐inositol‐O‐methyltransferase (IMT), consists of a transfer of a methyl group from S‐adenosylmethionine to myo‐inositol with the formation of d ‐ononitol, while the second reaction, catalyzed by d ‐ononitol epimerase (OEP), involves epimerization of d ‐ononitol to d ‐pinitol. To identify the genes involved in d ‐pinitol biosynthesis in a model legume Medicago truncatula, we conducted a BLAST search on its genome using soybean IMT cDNA as a query and found putative IMT (MtIMT) gene. Subsequent co‐expression analysis performed on publicly available microarray data revealed two potential OEP genes: MtOEPA, encoding an aldo‐keto reductase and MtOEPB, encoding a short‐chain dehydrogenase. cDNAs of all three genes were cloned and expressed as recombinant proteins in E. coli. In vitro assays confirmed that putative MtIMT enzyme catalyzes methylation of myo‐inositol to d ‐ononitol and showed that MtOEPA enzyme has NAD+‐dependent d ‐ononitol dehydrogenase activity, while MtOEPB enzyme has NADP+‐dependent d ‐pinitol dehydrogenase activity. Both enzymes are required for epimerization of d ‐ononitol to d ‐pinitol, which occurs in the presence of NAD+ and NADPH. Introduction of MtIMT, MtOEPA, and MtOEPB genes into tobacco plants resulted in production of d ‐ononitol and d ‐pinitol in transformants. As this two‐step pathway of d ‐ononitol epimerization is coupled with a transfer of reducing equivalents from NADPH to NAD+, we speculate that one of the functions of this pathway might be regeneration of NADP+ during drought stress.  相似文献   

9.
A solution of optically pure kynurenine (KYN), i.e., D ‐KYN or L ‐KYN, was administered intravenously to male Sprague‐Dawley rats (10 mg kg?1 ml?1). The time‐course of changes in the concentrations of urinary monoamines and their metabolites such as 5‐hydroxytryptamine (5‐HT), 5‐hydroxyindole acetic acid (5‐HIAA), dopamine, and 3‐methoxytyramine were investigated by reversed‐phase high‐performance liquid chromatography with electrochemical detection after precolumn derivatization with (2R)‐2,5‐dioxopyrrolidin‐1‐yl‐2,5,7,8‐tetramethyl‐6‐(tetrahydro‐2H‐pyran‐2‐yloxy)chroman‐2‐carboxylate (NPCA). We observed a stereoselective difference in the effects of the KYN enantiomers. Only D ‐KYN, not L ‐KYN, caused a significant increase in urinary 5‐HT levels within 30 min after its administration. With regard to the metabolites, urinary 3‐MT level was increased by D ‐KYN administration. On the other hand, no significant change in the DA level was observed after administration of either D ‐KYN or L ‐KYN. These results suggest that D ‐KYN could affect the activity of neuroactive amines, especially 5‐HT, in vivo. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

10.
Four diastereomeric‐Leu‐Leu‐Aib‐Leu‐Leu‐Aib‐peptides, Boc‐D ‐Leu‐L ‐Leu‐Aib‐L ‐Leu‐L ‐Leu‐Aib‐OMe (1), Boc‐L ‐Leu‐D ‐Leu‐Aib‐L ‐Leu‐L ‐Leu‐Aib‐OMe (2), Boc‐L ‐Leu‐L ‐Leu‐Aib‐D ‐Leu‐L ‐Leu‐Aib‐OMe (3), and Boc‐L ‐Leu‐L ‐Leu‐Aib‐L ‐Leu‐D ‐Leu‐Aib‐OMe (4), were synthesized. The crystals of the four hexapeptides were characterized by X‐ray crystallographic analysis. Two diastereomeric hexapeptides 1 and 2 having D ‐Leu(1) or D ‐Leu(2) were folded into right‐handed (P) 3 10 ‐helical structures, while peptide 3 having D ‐Leu(4) was folded into a turn structure nucleated by type III′ and I$' \bf{\beta}$ ‐turns, and peptide 4 having D ‐Leu(5) was folded into a left‐handed (M) 3 10 ‐helical structure. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
Recent studies have demonstrated that the O‐antigens of some pathogenic bacteria such as Brucella abortus, Francisella tularensis, and Campylobacter jejuni contain quite unusual N‐formylated sugars (3‐formamido‐3,6‐dideoxy‐d ‐glucose or 4‐formamido‐4,6‐dideoxy‐d ‐glucose). Typically, four enzymes are required for the formation of such sugars: a thymidylyltransferase, a 4,6‐dehydratase, a pyridoxal 5'‐phosphate or PLP‐dependent aminotransferase, and an N‐formyltransferase. To date, there have been no published reports of N‐formylated sugars associated with Mycobacterium tuberculosis. A recent investigation from our laboratories, however, has demonstrated that one gene product from M. tuberculosis, Rv3404c, functions as a sugar N‐formyltransferase. Given that M. tuberculosis produces l ‐rhamnose, both a thymidylyltransferase (Rv0334) and a 4,6‐dehydratase (Rv3464) required for its formation have been identified. Thus, there is one remaining enzyme needed for the production of an N‐formylated sugar in M. tuberculosis, namely a PLP‐dependent aminotransferase. Here we demonstrate that the M. tuberculosis rv3402c gene encodes such an enzyme. Our data prove that M. tuberculosis contains all of the enzymatic activities required for the formation of dTDP‐4‐formamido‐4,6‐dideoxy‐d ‐glucose. Indeed, the rv3402c gene product likely contributes to virulence or persistence during infection, though its temporal expression and location remain to be determined.  相似文献   

12.
3,4‐Dihydroxy‐2‐butanone‐4‐phosphate synthase (DHBPS) encoded by ribB gene is one of the first enzymes in riboflavin biosynthesis pathway and catalyzes the conversion of ribulose‐5‐phosphate (Ru5P) to 3,4‐dihydroxy‐2‐butanone‐4‐phosphate and formate. DHBPS is an attractive target for developing anti‐bacterial drugs as this enzyme is essential for pathogens, but absent in humans. The recombinant DHBPS enzyme of Salmonella requires magnesium ion for its activity and catalyzes the formation of 3,4‐dihydroxy‐2‐butanone‐4‐phosphate from Ru5P at a rate of 199 nmol min?1 mg?1 with Km value of 116 μM at 37°C. Further, we have determined the crystal structures of Salmonella DHBPS in complex with sulfate, Ru5P and sulfate‐zinc ion at a resolution of 2.80, 2.52, and 1.86 Å, respectively. Analysis of these crystal structures reveals that the acidic loop (residues 34–39) responsible for the acid‐base catalysis is disordered in the absence of substrate or metal ion at the active site. Upon binding either substrate or sulfate and metal ions, the acidic loop becomes stabilized, adopts a closed conformation and interacts with the substrate. Our structure for the first time reveals that binding of substrate Ru5P alone is sufficient for the stabilization of the acidic active site loop into a closed conformation. In addition, the Glu38 residue from the acidic active site loop undergoes a conformational change upon Ru5P binding, which helps in positioning the second metal ion that stabilizes the Ru5P and the reaction intermediates. This is the first structural report of DHBPS in complex with either substrate or metal ion from any eubacteria. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

13.
The primary structure of the B chain of the N‐acetyl‐D ‐galactosamine‐recognizing mistletoe lectin‐3 (ML‐3B) has been deduced from proteolytic digest peptides of the purified glycoprotein, their HPLC‐separation and Edman degradation and confirmation of the peptide sequences by MALDI‐MS. ML‐3B consists of 262 amino acid residues including 10 cysteine moieties. The structure and linkage of the carbohydrate side chains, connected to two N‐glycosylation sites at positions Asn95 and Asn135 of the lectin, were determined by a combination of glycosidase treatment and MALDI‐MS of corresponding glycopeptide fragments. The sequence alignment reveals a high homology with other B chains of type‐II RIPs, although there are remarkable differences in the D ‐galactose‐specific mistletoe lectin‐1B chain. The recently published primary structure of the mistletoe lectin‐3A chain 1 and the now available primary sequence of the 3B chain allowed the construction of a preliminary homology model of ML‐3. The model demonstrates, unequivocally, that ML‐3 is a member of the type‐II RIP family with rigid conservation of the enzymatic active site of the A chain and an identical overall protein fold. Specific amino acid residue exchanges and the different glycosylation pattern in comparison with ML‐1 are discussed and related to the properties of the two glycoproteins. The knowledge of the complete primary structure of mistletoe lectin‐3 is a major contribution towards more insight into the mechanism of the biological activity of commercial mistletoe preparations. Copyright © 2004 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
The relationship between the conformation and biological activity of the peptide allosteric modulator of the interleukin‐1 receptor 101.10 (D ‐Arg‐D ‐Tyr‐D ‐Thr‐D ‐Val‐D ‐Glu‐D ‐Leu‐D ‐Ala‐NH2) has been studied using (R)‐ and (S)‐Bgl residues. Twelve Bgl peptides were synthesized using (R)‐ and (S)‐cyclic sulfamidate reagents derived from L ‐ and D ‐aspartic acid in an optimized Fmoc‐compatible protocol for efficient lactam installment onto the supported peptide resin. Examination of these (R)‐ and (S)‐Bgl 101.10 analogs for their potential to inhibit IL‐1β‐induced thymocyte cell proliferation using a novel fluorescence assay revealed that certain analogs exhibited retained and improved potency relative to the parent peptide 101.10. In light of previous reports that Bgl residues may stabilize type II′β‐turn‐like conformations in peptides, CD spectroscopy was performed on selected compounds to identify secondary structure necessary for peptide biological activity. Results indicate that the presence of a fold about the central residues of the parent peptide may be important for activity. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
N‐carbamoyl‐amino‐acid amidohydrolase (also known as N‐carbamoylase) is the stereospecific enzyme responsible for the chirality of the D ‐ or L ‐amino acid obtained in the “Hydantoinase Process.” This process is based on the dynamic kinetic resolution of D ,L ‐5‐monosubstituted hydantoins. In this work, we have demonstrated the capability of a recombinant L ‐N‐carbamoylase from the thermophilic bacterium Geobacillus stearothermophilus CECT43 (BsLcar) to hydrolyze N‐acetyl and N‐formyl‐L ‐amino acids as well as the known N‐carbamoyl‐L ‐amino acids, thus proving its substrate promiscuity. BsLcar showed faster hydrolysis for N‐formyl‐L ‐amino acids than for N‐carbamoyl and N‐acetyl‐L ‐derivatives, with a catalytic efficiency (kcat/Km) of 8.58 × 105, 1.83 × 104, and 1.78 × 103 (s?1 M?1), respectively, for the three precursors of L ‐methionine. Optimum reaction conditions for BsLcar, using the three N‐substituted‐L ‐methionine substrates, were 65°C and pH 7.5. In all three cases, the metal ions Co2+, Mn2+, and Ni2+ greatly enhanced BsLcar activity, whereas metal‐chelating agents inhibited it, showing that BsLcar is a metalloenzyme. The Co2+‐dependent activity profile of the enzyme showed no detectable inhibition at high metal ion concentrations. © 2010 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

16.
N‐formylated sugars have been observed on the O‐antigens of such pathogenic Gram‐negative bacteria as Campylobacter jejuni and Francisella tularensis. Until recently, however, little was known regarding the overall molecular architectures of the N‐formyltransferases that are required for the biosynthesis of these unusual sugars. Here we demonstrate that the protein encoded by the wbtj gene from F. tularensis is an N‐formyltransferase that functions on dTDP‐4‐amino‐4,6‐dideoxy‐d ‐glucose as its substrate. The enzyme, hereafter referred to as WbtJ, demonstrates a strict requirement for N10‐formyltetrahydrofolate as its carbon source. In addition to the kinetic analysis, the three‐dimensional structure of the enzyme was solved in the presence of dTDP‐sugar ligands to a nominal resolution of 2.1 Å. Each subunit of the dimeric enzyme is dominated by a “core” domain defined by Met 1 to Ser 185. This core motif harbors the active site residues. Following the core domain, the last 56 residues fold into two α‐helices and a β‐hairpin motif. The hairpin motif is responsible primarily for the subunit:subunit interface, which is characterized by a rather hydrophobic pocket. From the study presented here, it is now known that WbtJ functions on C‐4′ amino sugars. Another enzyme recently investigated in the laboratory, WlaRD, formylates only C‐3′ amino sugars. Strikingly, the quaternary structures of WbtJ and WlaRD are remarkably different. In addition, there are several significant variations in the side chains that line their active site pockets, which may be important for substrate specificity. Details concerning the kinetic and structural properties of WbtJ are presented.  相似文献   

17.
The mannosylated derivative of adamant‐1‐yl tripeptide (D ‐(Ad‐1‐yl)Gly‐L ‐Ala‐D ‐isoGln) was prepared to study the effects of mannosylation on adjuvant (immunostimulating) activity. Mannosylated adamant‐1‐yl tripeptide (Man‐OCH2CH(Me)CO‐D ‐(Ad‐1‐yl)Gly‐L ‐Ala‐D ‐isoGln) is a non‐pyrogenic, H2O‐soluble, and non‐toxic compound. Adjuvant activity of mannosylated adamantyl tripeptide was tested in the mouse model with ovalbumin as an antigen and in comparison to the parent tripeptide and peptidoglycan monomer (PGM, β‐D ‐GlcNAc‐(1→4)‐D ‐MurNAc‐L ‐Ala‐D ‐isoGln‐mesoDAP(εNH2)‐D ‐Ala‐D ‐Ala), a well‐known effective adjuvant. The mannosylation of adamantyl tripeptide caused the amplification of its immunostimulating activity in such a way that it was comparable to that of PGM.  相似文献   

18.
Phosphoserine aminotransferase (SerC) from Escherichia coli (E. coli) MG1655 is engineered to catalyze the deamination of homoserine to 4‐hydroxy‐2‐ketobutyrate, a key reaction in producing 1,3‐propanediol (1,3‐PDO) from glucose in a novel glycerol‐independent metabolic pathway. To this end, a computation‐based rational approach is used to change the substrate specificity of SerC from l ‐phosphoserine to l ‐homoserine. In this approach, molecular dynamics simulations and virtual screening are combined to predict mutation sites. The enzyme activity of the best mutant, SerCR42W/R77W, is successfully improved by 4.2‐fold in comparison to the wild type when l ‐homoserine is used as the substrate, while its activity toward the natural substrate l ‐phosphoserine is completely deactivated. To validate the effects of the mutant on 1,3‐PDO production, the “homoserine to 1,3‐PDO” pathway is constructed in E. coli by coexpression of SerCR42W/R77W with pyruvate decarboxylase and alcohol dehydrogenase. The resulting mutant strain achieves the production of 3.03 g L?1 1,3‐PDO in fed‐batch fermentation, which is 13‐fold higher than the wild‐type strain and represents an important step forward to realize the promise of the glycerol‐independent synthetic pathway for 1,3‐PDO production from glucose.  相似文献   

19.
Klebsiella pneumoniae HR526, a new isolated 1,3‐propanediol (1,3‐PD) producer, exhibited great productivity. However, the accumulation of lactate in the late‐exponential phase remained an obstacle of 1,3‐PD industrial scale production. Hereby, mutants lacking D ‐lactate pathway were constructed by knocking out the ldhA gene encoding fermentative D ‐lactate dehydrogenase (LDH) of HR526. The mutant K. pneumoniae LDH526 with the lowest LDH activity was studied in aerobic fed‐batch fermentation. In experiments using pure glycerol as feedstock, the 1,3‐PD concentrations, conversion, and productivity increased from 95.39 g L?1, 0.48 and 1.98 g L?1 h?1 to 102. 06 g L?1, 0.52 mol mol?1 and 2.13 g L?1 h?1, respectively. The diol (1,3‐PD and 2,3‐butanediol) conversion increased from 0.55 mol mol?1 to a maximum of 0.65 mol mol?1. Lactate would not accumulate until 1,3‐PD exceeded 84 g L?1, and the final lactate concentration decreased dramatically from more than 40 g L?1 to <3 g L?1. Enzymic measurements showed LDH activity decreased by 89–98% during fed‐batch fermentation, and other related enzyme activities were not affected. NADH/NAD+ enhanced more than 50% in the late‐exponential phase as the D ‐lactate pathway was cut off, which might be the main reason for the change of final metabolites concentrations. The ability to utilize crude glycerol from biodiesel process and great genetic stability demonstrated that K. pnemoniae LDH526 was valuable for 1,3‐PD industrial production. Biotechnol. Bioeng. 2009; 104: 965–972. © 2009 Wiley Periodicals, Inc.  相似文献   

20.
Endo‐β‐1,4‐d ‐mannanase from the Antarctic springtail, Cryptopygus antarcticus (CaMan), is a cold‐adapted β‐mannanase that has the lowest optimum temperature (30°C) of all known β‐mannanases. Here, we report the apo‐ and mannopentaose (M5) complex structures of CaMan. Structural comparison of CaMan with other β‐mannanases from the multicellular animals reveals that CaMan has an extended loop that alters topography of the active site. Structural and mutational analyses suggest that this extended loop is linked to the cold‐adapted enzymatic activity. From the CaMan‐M5 complex structure, we defined the mannose‐recognition subsites and observed unreported M5 binding site on the surface of CaMan. Proteins 2014; 82:3217–3223. © 2014 Wiley Periodicals, Inc.  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