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.
Here, the term “module” is redefined for trans‐acyltransferase (trans‐AT) assembly lines to agree with how its domains cooperate and evolutionarily co‐migrate. The key domain in both the polyketide synthase (PKS) and nonribosomal peptide synthetase (NRPS) modules of assembly lines is the acyl carrier protein (ACP). ACPs not only relay growing acyl chains through the assembly line but also collaborate with enzymes in modules, both in cis and in trans, to add a specific chemical moiety. A ketosynthase (KS) downstream of ACP often plays the role of gatekeeper, ensuring that only a single intermediate generated by the enzymes of a module is passed downstream. Bioinformatic analysis of 526 ACPs from 33 characterized trans‐AT assembly lines reveals ACPs from the same module type generally clade together, reflective of the co‐evolution of these domains with their cognate enzymes. While KSs downstream of ACPs from the same module type generally also clade together, KSs upstream of ACPs do not—in disagreement with the traditional definition of a module. Beyond nomenclature, the presented analysis impacts our understanding of module function, the evolution of assembly lines, pathway prediction, and assembly line engineering.  相似文献   

2.
This paper reports the enantioseparation of β‐hydroxy‐1,2,3‐triazole derivatives, which present a broad range of biological properties, by supercritical fluid chromatography (SFC) and high‐performance liquid chromatography techniques (HPLC). Polysaccharide‐based chiral columns (cellulose and amylose) were used to evaluate the separation in SFC and HPLC. Time of analyses, consumption of solvent, and parameter optimization were reduced using SFC technique. The columns based on cellulose chiral stationary phase using 2‐propanol and ethanol as modifiers showed the best results for the enantioresolution of the (±)‐β‐hydroxy‐1,2,3‐triazoles by SFC analyses. These techniques were applied to evaluate the selectivity of biocatalytic reduction of β‐keto‐1,2,3‐triazoles by marine‐derived fungus Penicillium citrinum CBMAI 1186 to obtain the (±)‐β‐hydroxy‐1,2,3‐triazoles.  相似文献   

3.
Aims: Optimal production conditions of conjugated γ‐linolenic acid (CGLA) from γ‐linolenic acid using washed cells of Lactobacillus plantarum AKU 1009a as catalysts were investigated. Methods and Results: Washed cells of Lact. plantarum AKU 1009a exhibiting a high level of CGLA productivity were obtained by cultivation in a nutrient medium supplemented with 0·03% (w/v) α‐linolenic acid as an inducer. Under the optimal reaction conditions with 13 mg ml?1γ‐linolenic acid as a substrate in 5 ‐ml reaction volume, the washed cells [32% (wet cells, w/v) corresponding to 46 mg ml?1 dry cells] as the catalysts produced 8·8 mg CGLA per millilitre reaction mixture (68% molar yield) in 27 h. The produced CGLA was a mixture of two isomers, i.e., cis‐6,cis‐9,trans‐11‐octadecatrienoic acid (CGLA1, 40% of total CGLA) and cis‐6,trans‐9,trans‐11‐octadecatrienoic acid (CGLA2, 60% of total CGLA), and accounted for 66% of total fatty acid obtained. The CGLA produced was obtained as free fatty acids adsorbed mostly on the surface of the cells of Lact. plantarum AKU1009a. Conclusion: The practical process of CGLA production from γ‐linolenic acid using washed cells of Lact. plantarum AKU 1009a was successfully established. Significance and Impact of the Study: We presented the first example of microbial production of CGLA. CGLA produced by the process is valuable for evaluating their physiological and nutritional effects, and chemical characteristics.  相似文献   

4.
A fluorescence method was established for a α‐glucosidase activity assay and inhibitor screening based on β‐cyclodextrin‐coated quantum dots. p‐Nitrophenol, the hydrolysis product of the α‐glucosidase reaction, could quench the fluorescence of β‐cyclodextrin‐coated quantum dots via an electron transfer process, leading to fluorescence turn‐off, whereas the fluorescence of the system turned on in the presence of α‐glucosidase inhibitors. Taking advantage of the excellent properties of quantum dots, this method provided a very simple, rapid and sensitive screening method for α‐glucosidase inhibitors. Two α‐glucosidase inhibitors, 2,4,6‐tribromophenol and acarbose, were used to evaluate the feasibility of this screening model, and IC50 values of 24 μM and 0.55 mM were obtained respectively, which were lower than those previously reported. The method may have potential application in screening α‐glucosidase inhibitors. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

5.
The β‐N‐acetylhexosaminidase FDL specifically removes the β‐1,2‐GlcNAc residue conjugated to the α‐1,3‐mannose residue of the core structure of insect N‐glycans, playing significant physiological roles in post‐translational modification in the Golgi apparatus. Little is known about its enzymatic properties. We obtained the OfFDL gene from the insect Ostrinia furnacalis by RT‐PCR. The full length cDNA of FDL is 2241 bp carrying an opening reading frame of 1923 bp encoding 640 amino acids. The recombinant protein OfFDL in a soluble and active form was obtained with high purity through a two‐step purification strategy. The recombinant OfFDL exclusively hydrolyzes the terminal β‐1,2‐GlcNAc residue from the α‐1,3 branch instead of the α‐1,6 branch of the substrate GnGn‐PA. Several kinetic parameters including kcat/Km values toward four artificial substrates and Ki values of three representative hexosaminidase inhibitors were obtained.  相似文献   

6.
Biosynthesis of asymmetric carotenoids such as α‐carotene and lutein in plants and green algae involves the two enzymes lycopene β‐cyclase (LCYB) and lycopene ε‐cyclase (LCYE). The two cyclases are closely related and probably resulted from an ancient gene duplication. While in most plants investigated so far the two cyclases are encoded by separate genes, prasinophyte algae of the order Mamiellales contain a single gene encoding a fusion protein comprised of LCYB, LCYE and a C‐terminal light‐harvesting complex (LHC) domain. Here we show that the lycopene cyclase fusion protein from Ostreococcus lucimarinus catalyzed the simultaneous formation of α‐carotene and β‐carotene when heterologously expressed in Escherichia coli. The stoichiometry of the two products in E. coli could be altered by gradual truncation of the C‐terminus, suggesting that the LHC domain may be involved in modulating the relative activities of the two cyclase domains in the algae. Partial deletions of the linker region between the cyclase domains or replacement of one or both cyclase domains with the corresponding cyclases from the green alga Chlamydomonas reinhardtii resulted in pronounced shifts of the α‐carotene‐to‐β‐carotene ratio, indicating that both the relative activities of the cyclase domains and the overall structure of the fusion protein have a strong impact on the product stoichiometry. The possibility to tune the product ratio of the lycopene cyclase fusion protein from Mamiellales renders it useful for the biotechnological production of the asymmetric carotenoids α‐carotene or lutein in bacteria or fungi.  相似文献   

7.
Originally annotated as the initiator of fatty acid synthesis (FAS), β‐ketoacyl‐acyl carrier protein synthase III (KAS III) is a unique component of the bacterial FAS system. Novel variants of KAS III have been identified that promote the de novo use of additional extracellular fatty acids by FAS. These KAS III variants prefer longer acyl‐groups, notably octanoyl‐CoA. Acinetobacter baumannii, a clinically important nosocomial pathogen, contains such a multifunctional KAS III (AbKAS III). To characterize the structural basis of its substrate specificity, we determined the crystal structures of AbKAS III in the presence of different substrates. The acyl‐group binding cavity of AbKAS III and co‐crystal structure of AbKAS III and octanoyl‐CoA confirmed that the cavity can accommodate acyl groups with longer alkyl chains. Interestingly, Cys264 formed a disulfide bond with residual CoA used in the crystallization, which distorted helices at the putative interface with acyl‐carrier proteins. The crystal structure of KAS III in the alternate conformation can also be utilized for designing novel antibiotics.  相似文献   

8.
The process by which α-stereocenters of polyketide intermediates are set by modular polyketide synthases (PKSs) when condensation is not immediately followed by reduction is mysterious. However, the reductase-incompetent ketoreductase (KR) from the third module of 6-deoxyerythronolide B synthase has been proposed to operate as a racemase, aiding in the epimerization process that reverses the orientation of the α-methyl group of the polyketide intermediate generated by the ketosynthase to the configuration observed in the 6-deoxyerythronolide B final product. To learn more about the epimerization process, the structure of the C2-type KR from the third module of the pikromycin synthase, analogous to the KR from the third module of 6-deoxyerythronolide B synthase, was determined to 1.88 Å resolution. This first structural analysis of this KR-type reveals differences from reductase-competent KRs such as that the site NADPH binds to reductase-competent KRs is occluded by side chains and the putative catalytic tyrosine possesses more degrees of freedom. The active-site geometry may enable C2-type KRs to align the thioester and β-keto groups of a polyketide intermediate to reduce the pKa of the α-proton and accelerate its abstraction. Results from in vivo assays of engineered PKSs support that C2-type KRs cooperate with epimer-specific ketosynthases to set the configurations of substituent-bearing α-carbons.  相似文献   

9.
Ligand binding of neutral progesterone, basic propranolol, and acidic warfarin to human α1‐acid glycoprotein (AGP) was investigated by Raman spectroscopy. The binding itself is characterized by a uniform conformational shift in which a tryptophan residue is involved. Slight differences corresponding to different contacts of the individual ligands inside the β‐barrel are described. Results are compared with in silico ligand docking into the available crystal structure of deglycosylated AGP using quantum/molecular mechanics. Calculated binding energies are ?18.2, ?14.5, and ?11.5 kcal/mol for warfarin, propranolol, and progesterone, respectively. These calculations are consistent with Raman difference spectroscopy; nevertheless, minor discrepancies in the precise positions of the ligands point to structural differences between deglycosylated and native AGP. Thermal dynamics of AGP with/without bounded warfarin was followed by Raman spectroscopy in a temperature range of 10–95 °C and analyzed by principal component analysis. With increasing temperature, a slight decrease of α‐helical content is observed that coincides with an increase in β‐sheet content. Above 45 °C, also β‐strands tend to unfold, and the observed decrease in β‐sheet coincides with an increase of β‐turns accompanied by a conformational shift of the nearby disulfide bridge from high‐energy trans‐gauche‐trans to more relaxed gauche‐gauche‐trans. This major rearrangement in the vicinity of the bridge is not only characterized by unfolding of the β‐sheet but also by subsequent ligand release. Hereby, ligand binding alters the protein dynamics, and the more rigid protein–ligand complex shows an improved thermal stability, a finding that contributes to the reported chaperone‐like function of AGP. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

10.
Introduction – Bioautographic assays using TLC play an important role in the search for active compounds from plants. A TLC assay has previously been established for the detection of β‐glucosidase inhibitors but not for α‐glucosidase. Nonetheless, α‐glucosidase inhibition is an important target for therapeutic agents against of type 2 diabetes and anti‐viral infections. Objective – To develop a TLC bioautographic method to detect α‐ and β‐glucosidase inhibitors in plant extracts. Methodology – The enzymes α‐ and β‐d ‐glucosidase were dissolved in sodium acetate buffer. After migration of the samples, the TLC plate was sprayed with enzyme solution and incubated at room temperature for 60 min in the case of α‐d ‐glucosidase, and 37°C for 20 min in the case of β‐d ‐glucosidase. For detection of the active enzyme, solutions of 2‐naphthyl‐α‐D‐glucopyranoside or 2‐naphthyl‐β‐D‐glucopyranoside and Fast Blue Salt were mixed at a ratio of 1 : 1 (for α‐d ‐glucosidase) or 1 : 4 (for β‐d ‐glucosidase) and sprayed onto the plate to give a purple background colouration after 2–5 min. Results – Enzyme inhibitors were visualised as white spots on the TLC plates. Conduritol B epoxide inhibited α‐d ‐glucosidase and β‐d ‐glucosidase down to 0.1 µg. Methanol extracts of Tussilago farfara and Urtica dioica after migration on TLC gave enzymatic inhibition when applied in amounts of 100 µg for α‐glucosidase and 50 µg for β‐glucosidase. Conclusion – The screening test was able to detect inhibition of α‐ and β‐glucosidases by pure reference substances and by compounds present in complex matrices, such as plant extracts. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

11.
Chiral sulfoxides/N‐oxides (R)‐ 1 and (R,R)‐ 2 are effective chiral promoters in the enantioselective allylation of α‐keto ester N‐benzoylhydrazone derivatives 3a , 3b , 3c , 3d , 3e , 3f , 3g to generate the corresponding N‐benzoylhydrazine derivatives 4a , 4b , 4c , 4d , 4e , 4f , 4g , with enantiomeric excesses as high as 98%. Representative hydrazine derivatives 4a , 4b were subsequently treated with SmI2, and the resulting amino esters 5a , 5b with LiOH to obtain quaternary α‐substituted α‐allyl α‐amino acids 6a , 6b , whose absolute configuration was assigned as (S), with fundament on chemical correlation and electronic circular dichroism (ECD) data. Chirality 25:529–540, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

12.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
Clostridium propionicum is the only organism known to ferment β‐alanine, a constituent of coenzyme A (CoA) and the phosphopantetheinyl prosthetic group of holo‐acyl carrier protein. The first step in the fermentation is a CoA‐transfer to β‐alanine. Subsequently, the resulting β‐alanyl‐CoA is deaminated by the enzyme β‐alanyl‐CoA:ammonia lyase (Acl) to reversibly form ammonia and acrylyl‐CoA. We have determined the crystal structure of Acl in its apo‐form at a resolution of 0.97 Å as well as in complex with CoA at a resolution of 1.59 Å. The structures reveal that the enyzme belongs to a superfamily of proteins exhibiting a so called “hot dog fold” which is characterized by a five‐stranded antiparallel β‐sheet with a long α‐helix packed against it. The functional unit of all “hot dog fold” proteins is a homodimer containing two equivalent substrate binding sites which are established by the dimer interface. In the case of Acl, three functional dimers combine to a homohexamer strongly resembling the homohexamer formed by YciA‐like acyl‐CoA thioesterases. Here, we propose an enzymatic mechanism based on the crystal structure of the Acl·CoA complex and molecular docking. Proteins 2014; 82:2041–2053. © 2014 Wiley Periodicals, Inc.  相似文献   

14.
5‐Aminolevulinate synthase (ALAS) controls the rate‐limiting step of heme biosynthesis in mammals by catalyzing the condensation of succinyl‐coenzyme A and glycine to produce 5‐aminolevulinate, coenzyme‐A (CoA), and carbon dioxide. ALAS is a member of the α‐oxoamine synthase family of pyridoxal 5′‐phosphate (PLP)‐dependent enzymes and shares high degree of structural similarity and reaction mechanism with the other members of the family. The X‐ray crystal structure of ALAS from Rhodobacter capsulatus reveals that the alkanoate component of succinyl‐CoA is coordinated by a conserved arginine and a threonine. The functions of the corresponding acyl‐CoA‐binding residues in murine erthyroid ALAS (R85 and T430) in relation to acyl‐CoA binding and substrate discrimination were examined using site‐directed mutagenesis and a series of CoA‐derivatives. The catalytic efficiency of the R85L variant with octanoyl‐CoA was 66‐fold higher than that of the wild‐type protein, supporting the proposal of this residue as key in discriminating substrate binding. Substitution of the acyl‐CoA‐binding residues with hydrophobic amino acids caused a ligand‐induced negative dichroic band at 420 nm in the CD spectra, suggesting that these residues affect substrate‐mediated changes to the PLP microenvironment. Transient kinetic analyses of the R85K variant‐catalyzed reactions confirm that this substitution decreases microscopic rates associated with formation and decay of a key reaction intermediate and show that the nature of the acyl‐CoA tail seriously affect product binding. These results show that the bifurcate interaction of the carboxylate moiety of succinyl‐CoA with R85 and T430 is an important determinant in ALAS function and may play a role in substrate specificity.  相似文献   

15.
A series of polymer‐supported chiral β‐hydroxy amides and C2‐symmetric β‐hydroxy amides have been synthesized and successfully used for the enantioselective addition of phenylacetylene to aldehydes. High yields (up to 93%) and enantioselectivities (up to 92% ee) were achieved by using polymer‐supported chiral β‐hydroxy amide 4b . The resin 4b is reused four times, giving the product with enantioselectivity 80% ee. Fortunately, it is found that this heterogonous system is suitable not only for aromatic aldehydes but also aliphatic aldehyde. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

16.
Peroxisomes are thought to have played a key role in the evolution of metabolic networks of photosynthetic organisms by connecting oxidative and biosynthetic routes operating in different compartments. While the various oxidative pathways operating in the peroxisomes of higher plants are fairly well characterized, the reactions present in the primitive peroxisomes (microbodies) of algae are poorly understood. Screening of a Chlamydomonas insertional mutant library identified a strain strongly impaired in oil remobilization and defective in Cre05.g232002 (CrACX2), a gene encoding a member of the acyl‐CoA oxidase/dehydrogenase superfamily. The purified recombinant CrACX2 expressed in Escherichia coli catalyzed the oxidation of fatty acyl‐CoAs into trans‐2‐enoyl‐CoA and produced H2O2. This result demonstrated that CrACX2 is a genuine acyl‐CoA oxidase, which is responsible for the first step of the peroxisomal fatty acid (FA) β‐oxidation spiral. A fluorescent protein‐tagging study pointed to a peroxisomal location of CrACX2. The importance of peroxisomal FA β‐oxidation in algal physiology was shown by the impact of the mutation on FA turnover during day/night cycles. Moreover, under nitrogen depletion the mutant accumulated 20% more oil than the wild type, illustrating the potential of β‐oxidation mutants for algal biotechnology. This study provides experimental evidence that a plant‐type FA β‐oxidation involving H2O2‐producing acyl‐CoA oxidation activity has already evolved in the microbodies of the unicellular green alga Chlamydomonas reinhardtii.  相似文献   

17.
The conformation of oligomers of β‐amino acids of the general type Ac‐[β‐Xaa]n‐NHMe (β‐Xaa = β‐Ala, β‐Aib, and β‐Abu; n = 1–4) was systematically examined at different levels of ab initio molecular orbital theory (HF/6‐31G*, HF/3‐21G). The solvent influence was considered employing two quantum‐mechanical self‐consistent reaction field models. The results show a wide variety of possibilities for the formation of characteristic elements of secondary structure in β‐peptides. Most of them can be derived from the monomer units of blocked β‐peptides with n = 1. The stability and geometries of the β‐peptide structures are considerably influenced by the side‐chain positions, by the configurations at the Cα‐ and Cβ‐atoms of the β‐amino acid constituents, and especially by environmental effects. Structure peculiarities of β‐peptides, in particular those of various helix alternatives, are discussed in relation to typical elements of secondary structure in α‐peptides. © 1999 John Wiley & Sons, Inc. Biopoly 50: 167–184, 1999  相似文献   

18.
BamA of Escherichia coli is an essential component of the hetero‐oligomeric machinery that mediates β‐barrel outer membrane protein (OMP) assembly. The C‐ and N‐termini of BamA fold into trans‐membrane β‐barrel and five soluble POTRA domains respectively. Detailed characterization of BamA POTRA 1 missense and deletion mutants revealed two competing OMP assembly pathways, one of which is followed by the archetypal trimeric β‐barrel OMPs, OmpF and LamB, and is dependent on POTRA 1. Interestingly, our data suggest that BamA also requires its POTRA 1 domain for proper assembly. The second pathway is independent of POTRA 1 and is exemplified by TolC. Site‐specific cross‐linking analysis revealed that the POTRA 1 domain of BamA interacts with SurA, a periplasmic chaperone required for the assembly of OmpF and LamB, but not that of TolC and BamA. The data suggest that SurA and BamA POTRA 1 domain function in concert to assist folding and assembly of most β‐barrel OMPs except for TolC, which folds into a unique soluble α‐helical barrel and an OM‐anchored β‐barrel. The two assembly pathways finally merge at some step beyond POTRA 1 but presumably before membrane insertion, which is thought to be catalysed by the trans‐membrane β‐barrel domain of BamA.  相似文献   

19.
Guangcai Ma  Wenyou Zhu  Yongjun Liu 《Proteins》2016,84(11):1606-1615
Pectate lyase utilizes the anti‐β‐elimination chemistry to catalyze the cleavage of α‐1,4 glycosidic bond between D‐galacturonate regions during the degradation of plant polysaccharide pectin. We report here detailed mechanistic studies of the Bacillus subtilis pectate lyase (BsPel) using QM/MM calculations. It was found that the residue Arg279 serves as the catalytic base to abstract the α‐proton from C52 atom of substrate Ada2 subsite, forming an unstable carbanion intermediate. The glycosidic bond of this intermediate is scissile to generate the 4,5‐unsaturated digalacturonate product and a negatively charged β‐leaving group. Two active site residues (Lys247 and Arg279) and two Ca2+ ions (Ca2 and Ca3) form hydrogen‐bonding and coordination interactions with C52? COO? of Ada2, respectively, which facilitate the proton abstraction and stabilize the generated carbanion intermediates. Arg284 is not the potential proton donor to saturate the leaving group. Actually, the proton source of leaving group is the solvent water molecule rather than any active site acidic residues. In addition, the calculation results suggest that careful selections of QM‐ and Active‐regions are essential to accurately explore the enzymatic reactions. Proteins 2016; 84:1606–1615. © 2016 Wiley Periodicals, Inc.  相似文献   

20.
Macrolactam antibiotics such as incednine and cremimycin possess an aliphatic β‐amino acid as a starter unit of their polyketide chain. In the biosynthesis of incednine and cremimycin, unique stand‐alone adenylation enzymes IdnL1 and CmiS6 select and activate the proper aliphatic β‐amino acid as a starter unit. In this study, we describe the enzymatic characterization and the structural basis of substrate specificity of IdnL1 and CmiS6. Functional analysis revealed that IdnL1 and CmiS6 recognize 3‐aminobutanoic acid and 3‐aminononanoic acid, respectively. We solved the X‐ray crystal structures of IdnL1 and CmiS6 to understand the recognition mechanism of these aliphatic β‐amino acids. These structures revealed that IdnL1 and CmiS6 share a common recognition motif that interacts with the β‐amino group of the substrates. However, the hydrophobic side‐chains of the substrates are accommodated differently in the two enzymes. IdnL1 has a bulky Leu220 located close to the terminal methyl group of 3‐aminobutanoate of the trapped acyl‐adenylate intermediate to construct a shallow substrate‐binding pocket. In contrast, CmiS6 possesses Gly220 at the corresponding position to accommodate 3‐aminononanoic acid. This structural observation was supported by a mutational study. Thus, the size of amino acid residue at the 220 position is critical for the selection of an aliphatic β‐amino acid substrate in these adenylation enzymes. Proteins 2017; 85:1238–1247. © 2017 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