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.
Myxobacteria are able to produce the important metabolite isovaleryl coenzyme A by a route other than leucine degradation. The first step into this pathway is mediated by LiuC, a member of the 3‐methylglutaconyl CoA hydratases (MGCH). Here we present crystal structures refined to 2.05 and 1.1 Å of LiuC in the apo form and bound to coenzyme A, respectively. By using simulated annealing we modeled the enzyme substrate complex and identified residues potentially involved in substrate binding, specificity, and catalysis. The dehydration of 3‐hydroxy‐3‐methylglutaconyl CoA to 3‐methylglutaconyl CoA catalyzed by LiuC involves Glu112 and Glu132 and likely employs the typical crotonase acid‐base mechanism. In this, Tyr231 and Arg69 are key players in positioning the substrate to enable catalysis. Surprisingly, LiuC shows higher sequence and structural similarity to human MGCH than to bacterial forms, although they convert the same substrate. This study provides structural insights into the alternative isovaleryl coenzyme A biosynthesis pathway and might open a path for biofuel research, as isovaleryl‐CoA is a source for isobutene, a precursor for renewable fuels and chemicals.  相似文献   

2.
CLK2 inhibition has been proposed as a potential mechanism to improve autism and neuronal functions in Phelan–McDermid syndrome (PMDS). Herein, the discovery of a very potent indazole CLK inhibitor series and the CLK2 X‐ray structure of the most potent analogue are reported. This new indazole series was identified through a biochemical CLK2 Caliper assay screen with 30k compounds selected by an in silico approach. Novel high‐resolution X‐ray structures of all CLKs, including the first CLK4 X‐ray structure, bound to known CLK2 inhibitor tool compounds (e.g., TG003, CX‐4945), are also shown and yield insight into inhibitor selectivity in the CLK family. The efficacy of the new CLK2 inhibitors from the indazole series was demonstrated in the mouse brain slice assay, and potential safety concerns were investigated. Genotoxicity findings in the human lymphocyte micronucleus test (MNT) assay are shown by using two structurally different CLK inhibitors to reveal a major concern for pan‐CLK inhibition in PMDS.  相似文献   

3.
Nitrosomonas europaea cytochrome c‐552 (Ne c‐552) variants with the same His/Met axial ligand set but with different EPR spectra have been characterized structurally, to aid understanding of how molecular structure determines heme electronic structure. Visible light absorption, Raman, and resonance Raman spectroscopy of the protein crystals was performed along with structure determination. The structures solved are those of Ne c‐552, which displays a “HALS” (or highly anisotropic low‐spin) EPR spectrum, and of the deletion mutant Ne N64Δ, which has a rhombic EPR spectrum. Two X‐ray crystal structures of wild‐type Ne c‐552 are reported; one is of the protein isolated from N. europaea cells (Ne c‐552n, 2.35 Å resolution), and the other is of recombinant protein expressed in Escherichia coli (Ne c‐552r, 1.63 Å resolution). Ne N64Δ crystallized in two different space groups, and two structures are reported [monoclinic (2.1 Å resolution) and hexagonal (2.3 Å resolution)]. Comparison of the structures of the wild‐type and mutant proteins reveals that heme ruffling is increased in the mutant; increased ruffling is predicted to yield a more rhombic EPR spectrum. The 2.35 Å Ne c‐552n structure shows 18 molecules in the asymmetric unit; analysis of the structure is consistent with population of more than one axial Met configuration, as seen previously by NMR. Finally, the mutation was shown to yield a more hydrophobic heme pocket and to expel water molecules from near the axial Met. These structures reveal that heme pocket residue 64 plays multiple roles in regulating the axial ligand orientation and the interaction of water with the heme. These results support the hypothesis that more ruffled hemes lead to more rhombic EPR signals in cytochromes c with His/Met axial ligation.  相似文献   

4.
Methylmalonyl-CoA epimerase (MMCE) is proposed to use general acid-base catalysis, but the proposed catalytic glutamic acids are highly asymmetrical in the active site unlike many other racemases. To gain insight into the puzzling relationships between catalytic mechanism, structure, and substrate preference, we solved Streptomyces coelicolor MMCE structures with substrate or 2-nitropropionyl-CoA, an intermediate/transition state analogue. Both ligand bound structures have a planar methylmalonate/2-nitropropionyl moiety indicating a deprotonated C2 with ≥4 Å distances to either catalytic acid. Both glutamates interact with the carboxylate/nitro group, either directly or through other residues. This suggests the proposed catalytic acids sequentially catalyze proton shifts between C2 and carboxylate of the substrate with an enolate intermediate. In addition, our structures provide a platform to design mutations for expanding substrate scope to support combinatorial biosynthesis.  相似文献   

5.
Despite extensive research into triosephosphate isomerases (TIMs), there exists a gap in understanding of the remarkable conjunction between catalytic loop‐6 (residues 166–176) movement and the conformational flip of Glu165 (catalytic base) upon substrate binding that primes the active site for efficient catalysis. The overwhelming occurrence of serine at position 96 (98 % of the 6277 unique TIM sequences), spatially proximal to E165 and the loop‐6 residues, raises questions about its role in catalysis. Notably, Plasmodium falciparum TIM has an extremely rare residue—phenylalanine—at this position whereas, curiously, the mutant F96S was catalytically defective. We have obtained insights into the influence of residue 96 on the loop‐6 conformational flip and E165 positioning by combining kinetic and structural studies on the PfTIM F96 mutants F96Y, F96A, F96S/S73A, and F96S/L167V with sequence conservation analysis and comparative analysis of the available apo and holo structures of the enzyme from diverse organisms.  相似文献   

6.
Analysis of the recently solved X‐ray crystal structures of Saccharomyces cerevisiae ribonucleotide reductase I (ScRnr1) in complex with effectors and substrates led to the discovery of a conserved water molecule located at the active site that interacted with the 2′‐hydroxy group of the nucleoside ribose. In this study 2′‐(2‐hydroxyethyl)‐2′‐deoxyadenosine 1 and the 5′‐diphosphate derivative 2 were designed and synthesized to see if the conserved water molecule could be displaced by a hydroxymethylene group, to generate novel RNR inhibitors as potential antitumor agents. Herein we report the synthesis of analogues 1 and 2 , and the co‐crystal structure of adenosine diphosphate analogue 2 bound to ScRnr1, which shows the conserved water molecule is displaced as hypothesized.  相似文献   

7.
The copolymers of (Z)‐4‐oxo‐4‐phenoxyl‐2‐butenoic acid with styrene (PSt/OPBA) and their macromolecular luminous lanthanide complexes (Ln‐PSt/OPBA) have been synthesized and characterized by means of GPC, elemental analysis, FTIR, X‐ray powder diffraction, spectral analysis, and thermal analysis. The IR studies showed that the carboxylic groups on the side chain of the polymer were coordinated to lanthanide ions by bidentate manner. However, the ethereal oxygen, instead of carbonyl, also bonded to the central lanthanide ions, which was an intriguing phenomenon for ester‐coordinated complexes. X‐ray diffraction experiments revealed that these PSt/OPBA copolymers were amorphous, but Ln‐PSt/OPBA were crystalline, in which the complex Eu‐Lc belonged to a high symmetric structure of orthorhombic quadratic system, with a = 10.59 ± 0.02 Å, c = 8.02 ± 0.01 Å; c/a = 0.763. In addition, the value δ (the number of free carboxylic groups) in Ln‐PSt/OPBA complexes increased with the decreasing mole ratio of styrene in the copolymers, while it decreased with increasing pH values of the solution. Eu3+ and Tb3+ complexes exhibited characteristic fluorescence with comparatively high brightness and good monochromaticity, and the fluorescence intensity was enhanced with increasing the content of lanthanide up to around 18 wt % without typical fluorescence concentration quenching behavior in the solid state. So using polymers as a matrix, Ln‐PSt/OPBA are likely to provide new materials that possess specific properties and desired features. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

8.
The synthesis and application of bifunctional mesoporous Al‐P‐V—O catalysts with both acidic and redox sites for selective oxidation of methanol to dimethoxymethane (DMM) is described. The catalysts were characterized by N2 adsorption/desorption, X‐ray diffraction, temperature‐programmed desorption, X‐ray photoelectron spectroscopy, and infrared spectroscopy. It is shown that porosity; redox property and surface acidity of the catalysts were greatly influenced by the Al/V/P ratio. The synergistic effect of phosphorus and vanadium was investigated. Al‐P‐V—O catalysts exhibited good catalytic activity because of the controlled reducibility and the acidic sites. © 2013 American Institute of Chemical Engineers AIChE J, 59: 2587–2593, 2013  相似文献   

9.
A mesoporous silica (SBA‐15)‐supported pyrimidine‐substituted N‐heterocyclic carbene iridium complex was prepared and used as a catalyst for both environmentally friendly N‐alkylation of amines and β‐alkylation of secondary alcohols with primary alcohols. The structure of the supported iridium catalyst was characterized by Fourier transform infrared (FT‐IR), 13C and 29Si solid‐state nuclear magnetic resonance (NMR), small‐angle X‐ray scattering (SAXS), transmission electron microscopy (TEM), iridium K‐edge X‐ray absorption near‐edge structure (XANES) and extended X‐ray absorption fine structure (EXAFS) spectroscopic analyses which demonstrated that the coordination environment of the iridium centre and the 3‐dimensional‐hexagonal pore structure of SBA‐15 were retained after the immobilization. The catalyst was found to be highly efficient for both kinds of reaction on a wide range of substrates under mild conditions. Moreover, the supported iridium catalyst was obviously superior to the unsupported one in the N‐alkylation of aniline and β‐alkylation of 1‐phenylethanol with benzyl alcohol as substrate, which indicated that not only the iridium complex moiety but also the support material contributed to the catalytic activity of the supported iridium catalyst in these reactions. The supported iridium catalyst can be easily recycled by simple washing without chemical treatment, and exhibited excellent recycling performance without notable decrease in catalytic efficiency even after twelve test cycles for N‐alkylation of aniline with benzyl alcohol, nine cycles for N‐alkylation of different amines with different alcohols, and eight cycles for β‐alkylation of 1‐phenylethanol with benzyl alcohol, respectively.  相似文献   

10.
A facile chemo‐enzymatic process has been developed for producing stereoisomers of 4‐substituted 2‐hydroxy‐4‐butyrolactones with good to excellent enantioselectivity. This process involves an easy separation of the diastereoisomers by column chromatography and efficient enzymatic resolution by whole cells of Escherichia coli JM109 expressing Fusarium proliferatum lactonase gene. This biocatalyst shows strong tolerance towards different substrate structures and at least three out four possible isomers could be obtained in excellent enantiomeric purity. Different substrate concentrations (10 mM–200 mM) were examined, giving a substrate to catalyst ratio of up to 26:1. This general and efficient enzymatic process provides access to stereoisomers of 4‐substituted 2‐hydroxy‐4‐butyrolactones readily and cost‐effectively. The stereochemical assignments were conducted systematically based on NMR, X‐ray diffraction and circular dichroism, leading to further understanding of the enzyme’s stereoselectivity.  相似文献   

11.
We have probed Pichia stipitis CBS 6054 Old Yellow Enzyme 2.6 (OYE 2.6) by several strategies including X‐ray crystallography, ligand binding and catalytic assays using the wild‐type as well as libraries of site‐saturation mutants. The alkene reductase crystallized in space group P 63 2 2 with unit cell dimensions of 127.1×123.4 Å and its structure was solved to 1.5 Å resolution by molecular replacement. The protein environment surrounding the flavin mononucleotide (FMN) cofactor was very similar to those of other OYE superfamily members; however, differences in the putative substrate binding site were also observed. Substrate analog complexes were analyzed by both UV‐Vis titration and X‐ray crystallography to provide information on possible substrate binding interactions. In addition, four active site residues were targeted for site saturation mutagenesis (Thr 35, Ile 113, His 188, His 191) and each library was tested against three representative Baylis–Hillman adducts. Thr 35 could be replaced by Ser with no change in activity; other amino acids (Ala, Cys, Leu, Met, Gln and Val) resulted in diminished catalytic efficiency. The Ile 113 replacement library yielded a range of catalytic activities, but had very little impact on stereoselectivity. Finally, the two His residues (188 and 191) were essentially intolerant of substitutions with the exception of the His 191 Asn mutant, which did show significant catalytic ability. Structural comparisons between OYE 2.6 and Saccharomyces pastorianus OYE1 suggest that the key interactions between the substrate hydroxymethyl groups and the side‐chain of Thr 35 and/or Tyr 78 play an important role in making OYE 2.6 an (S)‐selective alkene reductase.  相似文献   

12.
The binding mode of several substrate analogues, (2R)‐2‐benzyl‐3‐dehydroquinic acids 4 , which are potent reversible competitive inhibitors of type II dehydroquinase (DHQ2), the third enzyme of the shikimic acid pathway, has been investigated by structural and computational studies. The crystal structures of Mycobacterium tuberculosis and Helicobacter pylori DHQ2 in complex with one of the most potent inhibitor, p‐methoxybenzyl derivative 4 a , have been solved at 2.40 Å and 2.75 Å, respectively. This has allowed the resolution of the M. tuberculosis DHQ2 loop containing residues 20–25 for the first time. These structures show the key interactions of the aromatic ring in the active site of both enzymes and additionally reveal an important change in the conformation and flexibility of the loop that closes over substrate binding. The loop conformation and the binding mode of compounds 4 b – d has been also studied by molecular dynamics simulations, which suggest that the benzyl group of inhibitors 4 prevent appropriate orientation of the catalytic tyrosine of the loop for proton abstraction and disrupts its basicity.  相似文献   

13.
Aliphatic polycarbonate‐based polyurethane (PC‐PU) elastomers as well as their nanocomposites with organic‐modified clay (bentonite for organic system) were synthesized. Macrodiols (MD) (randomly copolymerized aliphatic PC‐glycols of molecular weight of about 2000: T5652, T4672, and T4692), hexamethylene diisocyanate, and butane‐1,4‐diol were used as starting materials. Solid‐state NMR and Fourier transform infrared spectroscopy, small‐angle X‐ray scattering, wide‐angle X‐ray diffraction, atomic force microscopy, and transmission electron microscopy were used for studying the bottom‐up self‐assembly of building units from the segmental level up that of organized structures of micrometer sizes. Contents of hard segments formed by the reaction of chain extender with diisocyanate plays a dominant role for the degree of ordering and related phenomena, while the MD chain has only limited effect on PC‐PU properties. The spectroscopy and scattering experiments suggest that bentonite particles incorporate well in the structure and promote the ordering of hard segment domains in PC‐PU matrix as compared with the nanofiller‐free analogue. © 2012 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

14.
We modified hydrophobic poly(3‐hydroxybutyrate‐co‐3‐hydroxyvalerate) (PHBHV) films with hydrophilic chains to control their surface properties. We designed and investigated surface‐initiated atom transfer radical polymerization (SI‐ATRP) to modify the PHBHV films by grafting poly(2‐hydroxyethyl methacrylate) (PHEMA) from the surface. This method consisted of two steps. In the first step, amino functions were formed on the surface by aminolysis; this was followed by the immobilization of an atom transfer radical polymerization initiator, 2‐bromoisobutyryl bromide. In the second step, the PHEMA chains were grafted to the substrate by a polymerization process initiated by the surface‐bound initiator. The SI‐ATRP technique was expected to favor a polymerization process with a controlled manner. The experimental results demonstrate that the grafting density was controlled by the reaction conditions in the first step. The grafted films were analyzed by Fourier transform infrared spectroscopy, contact angle testing, scanning electron microscopy, and energy‐dispersive X‐ray spectroscopy. The results show that grafted chains under the SI‐ATRP method were preferentially located on the surface for surface grafting and in the bulk for conventional free‐radical polymerization initiated by benzoyl peroxide. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

15.
The potential of diesel particulate filters wash‐coated with highly dispersed nano‐metric ceria particles for continuous regeneration has been investigated. To this end, catalytic filters were prepared, soot‐loaded (avoiding the formation of the cake layer), and regenerated—under isothermal conditions—at temperature ranging from 200–600°C. Results have shown that catalytic oxidation of soot starts from 300°C and, at all temperatures, the selectivity to CO2 is higher than 99%. 475°C is the minimum temperature at which the filter is regenerated via catalytic path. At this temperature, the catalytic filter maintains substantially the same performance over repeated cycles of soot loading and regeneration, indicating that the thermal stability of ceria is preserved. This has been further confirmed by comparison between the outcomes obtained from characterization (X‐ray powder diffraction, N2 adsorption at 77 K, Hg intrusion porosimetry, and scanning electron microscope/energy dispersive X‐ray analysis) of fresh filter and filter subjected to repeated regeneration tests. © 2017 American Institute of Chemical Engineers AIChE J, 63: 3442–3449, 2017  相似文献   

16.
The influence of cooling rates on the thermal behavior and solid‐state morphologies of polyhydroxyalkanoates have been investigated. The thermal behavior was studied by differential scanning calorimetry (DSC). The crystal structures (~ Å), lamellar (tens of nanometers), fibrillar (several hundred nanometers), and spherulitic (~ μm) morphologies of poly (3‐hydroxybutyrate) (PHB) and the copolymers of poly (3‐hydroxybutyric acid‐co‐3‐hydroxyvaleric acid) (PHBV) and poly (3‐hydroxybutyric acid‐co‐3‐hydroxyhexanoic acid) (PHBHx) crystallized under different cooling rates were studied using simultaneous small angle X‐ray scattering (SAXS) and wide angle X‐ray scattering, simultaneous ultra small angle X‐ray scattering (USAXS) and SAXS, and polarized optical microscopy, respectively. The experimental results showed that the lamellar and spherulitic morphologies depended strongly on cooling rates. However, there was little influence of cooling rates on the crystal structures. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

17.
The nickel catalysts derived from Cr‐doped LaNiO3 perovskite‐like precursors were characterized by X‐ray diffraction, high‐resolution transmission electron microscopy, temperature‐programmed oxidation, temperature‐programmed reduction, and X‐ray photoelectron spectroscopy. Their catalytic performance in CO2 reforming of methane under microwave irradiation was investigated. It was found that the structure and morphology of the oxide composites in this research were influenced by the ratio of Ni and Cr, and the mismatch of La3+, Ni3+, and Cr3+ may cause phase segregation. The catalytic performance of the Ni catalysts is dependent on the oxygen mobility of the perovskite oxide matrix, the content of the reduced Ni0, and the content of the remaining perovskite structure. The mobile oxygen in the perovskite matrix in the catalyst may enhance the conversion of CO2 during the reaction.  相似文献   

18.
Polystyrene‐grafted attapulgite (ATP‐g‐PS) was prepared using surface‐initiated redox initiation via graft from approach. ATP was modified with (3‐ aminopropyl) triethoxysilane (APTES) to anchor amine on the surface (ATP‐NH2), and then ATP‐NH2 was further treated with 4‐(diethylamino) salicylaldehyde (DEAS) to give aromatic tertiary amine groups functional ATP (ATP‐ATA). Subsequently, the surface‐initiated redox polymerization of styrene in the presence of ATP‐ATA and benzoyl peroxide (BPO) was performed to afford ATP‐g‐PS . The chemical grafting of PS on the surface of ATP was confirmed by Fourier transform infrared spectroscopy, X‐ray photoelectron spectroscopy, and thermogravimetric analysis . The crystal structure of PS‐grafted ATP was characterized by X‐ray diffraction (XRD) analysis. The morphology of ATP‐g‐PS was observed by scanning electron microscopy (SEM) . POLYM. ENG. SCI., 55:889–895, 2015. © 2014 Society of Plastics Engineers  相似文献   

19.
Wide‐angle (WAXS) and small‐angle X‐ray scattering (SAXS) studies of dry granular zein, zein fibers, zein–oleic acid resin, and zein–oleic acid films are reported. WAXS patterns showed two diffuse rings for these samples indicative of noncrystalline structures. Measured d‐spacings of ∼ 4.6 Å and ∼ 10.5 Å were found for zein–oleic acid resins and films, consistent with the presence of α‐helical segments. The granular zein and zein fibers had ∼ 4.6‐Å and ∼ 9.5‐Å spacings. Neither the films nor the fibers showed evidence of orientation of the molecular axes. SAXS studies of zein–oleic acid films indicated that the structure of the films was affected by preparation method. Biaxially drawn resin films showed periodicities of ∼ 170 Å along the film surface direction and ∼ 135 Å in the thickness direction, while the cast films had weaker intensity periodicities of ca. 80 Å for all beam directions; a weak, diffuse 45‐Å spacing was also observed for both samples. The 170‐Å periodicity was present in the resin before deformation and following uniaxial deformation. No SAXS periodicity was observed for the granular zein or zein fibers. Several structural models are presented for the resin films that are consistent with reports in the literature that zein, in solution, consist of prism‐like particles consisting of four or more molecules. ? 1999 John Wiley & Sons, Inc. J Appl Polym Sci 71: 1267‐1281, 1999  相似文献   

20.
The distribution of aroma chemical cis‐3‐hexenyl salicylate, on both longitudinal and cross‐sectional fiber directions, was identified through backscattered electron microscopy and X‐ray microanalysis including X‐ray spectrum and X‐ray map. Three fibers—cotton, lyocell, and polyester [poly(ethylene terephthalate) (PET)]—were used as substrates to evaluate the influence of fiber physical/chemical nature on the distribution of cis‐3‐hexenyl salicylate. It was found that the distribution of cis‐3‐hexenyl salicylate on the external and internal fiber surfaces correlated strongly with the chemical structure, roughness, and both pore and capillary structure of the textiles. cis‐3‐Hexenyl salicylate distributed through the whole cotton fiber cross section with higher concentrations in lumen and crenulations, whereas it distributed relatively uniformly in the surface and cross section of lyocell fiber. This is believed to relate to the macro‐ and micropores, macroscopic roughness, and the presence of a larger number of polar groups for these cellulose fibers. In contrast, cis‐3‐hexenyl salicylate accumulated at a few spots on the fiber surfaces of PET and in interfiber spaces of closely packed fibers, attributed to lower polarity, round cross‐sectional shape, smooth surfaces, and fewer voids of the PET fiber structure. © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 91: 3557–3564, 2004  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