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.
Cell Biology and Toxicology - The human enzyme As(III) S-adenosylmethionine methyltransferase (AS3MT) catalyzes arsenic biotransformations and is considered to contribute to arsenic-related...  相似文献   

2.
The arsRDABC operon of Escherichia coli plasmid R773 encodes the ArsAB pump that catalyzes extrusion of the metalloids As(III) and Sb(III), conferring metalloid resistance. The catalytic subunit, ArsA, is an ATPase with two homologous halves, A1 and A2, connected by a short linker. Each half contains a nucleotide binding domain. The overall rate of ATP hydrolysis is slow in the absence of metalloid and is accelerated by metalloid binding. The results of photolabeling of ArsA with the ATP analogue 8-azidoadenosine 5'-[alpha-(32)P]-triphosphate at 4 degrees C indicate that metalloid stimulation correlates with a >10-fold increase in affinity for nucleotide. To investigate the relative contributions of the two nucleotide binding domains to catalysis, a thrombin site was introduced in the linker. This allowed discrimination between incorporation of labeled nucleotides into the two halves of ArsA. The results indicate that both the A1 and A2 nucleotide binding domains bind and hydrolyze trinucleotide, even in the absence of metalloid. Sb(III) increases the affinity of the A1 nucleotide binding domain to a greater extent than the A2 nucleotide binding domain. The ATP analogue labeled with (32)P at the gamma position was used to measure hydrolysis of trinucleotide at 37 degrees C. Under these catalytic conditions, both nucleotide binding domains hydrolyze ATP, but hydrolysis in A1 is stimulated to a greater degree by Sb(III) than A2. These results suggest that the two homologous halves of the ArsA may be functionally nonequivalent.  相似文献   

3.
The ArsD metallochaperone delivers trivalent metalloids, As(III) or Sb(III), to the ArsA ATPase, the catalytic subunit of the ArsAB As(III) efflux pump. Transfer of As(III) increases the affinity of ArsA for As(III), allowing resistance to environmental arsenic concentrations. As(III) transfer is channelled from chaperone to ATPase, implying that ArsD and ArsA form an interface at their metal binding sites. A genetic approach was used to test this hypothesis. Thirteen ArsD mutants exhibiting either weaker or stronger interaction with ArsA were selected by either repressed transactivator yeast two-hybrid or reverse yeast two-hybrid assays. Additionally, Lys-37 and Lys-62 were identified as being involved in ArsD function by site-directed mutagenesis and chemical modification. Substitution at either position with arginine was tolerated, suggesting participation of a positive charge. By yeast two-hybrid analysis K37A and K62A mutants lost interaction with ArsA. All 15 mutations were mapped on the surface of the ArsD structure, and their locations are consistent with a structural model generated by in silico docking. Four are close to metalloid binding site residues Cys-12, Cys-13 and Cys-18, and seven are on the surface of helix 1. These results suggest that the interface involves one surface of helix 1 and the metalloid binding site.  相似文献   

4.
The arsRDABC operon of Escherichia coli plasmid R773 encodes the ArsAB extrusion pump for the trivalent metalloids As(III) and Sb(III). ArsA, the catalytic subunit has two homologous halves, A1 and A2. Each half has a consensus signal transduction domain that physically connects the nucleotide-binding domain to the metalloid-binding domain. The relation between metalloid binding by ArsA and transport through ArsB is unclear. In this study, direct metalloid binding to ArsA was examined. The results show that ArsA binds a single Sb(III) with high affinity only in the presence of Mg(2+)-nucleotide. Mutation of the codons for Cys-113 and Cys-422 eliminated Sb(III) binding to purified ArsA. C113A/C422A ArsA has basal ATPase activity similar to that of the wild type but lacks metalloid-stimulated activity. Accumulation of metalloid was assayed in intact cells, where reduced uptake results from active extrusion by the ArsAB pump. Cells expressing the arsA(C113A/C422A)B genes had an intermediate level of metalloid resistance and accumulation between those expressing only arsB alone and those expressing wild type arsAB genes. The results indicate that, whereas metalloid stimulation of ArsA activity enhances the ability of the pump to reduce the intracellular concentration of metalloid, high affinity binding of metalloid by ArsA is not obligatory for transport or resistance. Yet, in mixed populations of cells bearing either arsAB or arsA(C113A/C422A)B growing in subtoxic concentrations of arsenite, cells bearing wild type arsAB replaced cells with mutant arsA(C113A/C422A)B in less than 1 week, showing that the metalloid binding site confers an evolutionary advantage.  相似文献   

5.
Arsenic, a toxic metalloid widely existing in the environment, causes a variety of health problems. The ars operon encoded by Escherichia coli plasmid R773 has arsD and arsA genes, where ArsA is an ATPase that is the catalytic subunit of the ArsAB As(III) extrusion pump, and ArsD is an arsenic chaperone for ArsA. ArsD transfers As(III) to ArsA and increases the affinity of ArsA for As(III), allowing resistance to environmental concentrations of arsenic. Cys12, Cys13 and Cys18 in ArsD form a three sulfur-coordinated As(III) binding site that is essential for metallochaperone activity. ATP hydrolysis by ArsA is required for transfer of As(III) from ArsD to ArsA, suggesting that transfer occurs with a conformation of ArsA that transiently forms during the catalytic cycle. The 1.4 Å x-ray crystal structure of ArsD shows a core of four ??-strands flanked by four ??-helices in a thioredoxin fold. Docking of ArsD with ArsA was modeled in silico. Independently ArsD mutants exhibiting either weaker or stronger interaction with ArsA were selected. The locations of the mutations mapped on the surface of ArsD are consistent with the docking model. The results suggest that the interface with ArsA involves one surface of ??1 helix and metalloid binding site of ArsD.  相似文献   

6.
ArsD is a metallochaperone that delivers As(III) to the ArsA ATPase, the catalytic subunit of the ArsAB pump encoded by the arsRDABC operon of Escherichia coli plasmid R773. Conserved ArsD cysteine residues (Cys12, Cys13 and Cys18) construct the As(III) binding site of the protein, however a global structural understanding of this arsenic binding remains unclear. We have obtained NMR assignments for ArsD as a starting point for probing structural changes on the protein that occur in response to metalloid binding and upon formation of a complex with ArsA. The predicted solution structure of ArsD is in agreement with recently published crystallographic structural results.  相似文献   

7.
The ArsAB extrusion pump encoded by the ars operon of Escherichia coli plasmid R773 confers resistance to the toxic trivalent metalloids arsenite [As(III)] and antimonite [Sb(III)]. The ArsA ATPase, the catalytic subunit of the pump, has two homologous halves, A1 and A2. At the interface of these two halves are two nucleotide-binding domains and a metalloid-binding domain. Cys-113 and Cys-422 have been shown to form a high-affinity metalloid binding site. The crystal structure of ArsA shows two other bound metalloid atoms, one liganded to Cys-172 and His-453, and the other liganded to His-148 and Ser-420. The contribution of those putative metalloid sites was examined. There was little effect of mutagenesis of residues His-148 and Ser-420 on metalloid binding. However, a C172A ArsA mutant and C172A/H453A double mutant exhibited significantly decreased affinity for Sb(III). These results suggest first that there is only a single high-affinity metalloid binding site in ArsA, and second that Cys-172 controls the affinity of this site for metalloid and hence the efficiency of metalloactivation of the ArsAB efflux pump.  相似文献   

8.
Asp kinase catalyzes the first step of the Asp-derived essential amino acid pathway in plants and microorganisms. Depending on the source organism, this enzyme contains up to four regulatory ACT domains and exhibits several isoforms under the control of a great variety of allosteric effectors. We report here the dimeric structure of a Lys and S-adenosylmethionine-sensitive Asp kinase isoform from Arabidopsis thaliana in complex with its two inhibitors. This work reveals the structure of an Asp kinase and an enzyme containing two ACT domains cocrystallized with its effectors. Only one ACT domain (ACT1) is implicated in effector binding. A loop involved in the binding of Lys and S-adenosylmethionine provides an explanation for the synergistic inhibition by these effectors. The presence of S-adenosylmethionine in the regulatory domain indicates that ACT domains are also able to bind nucleotides. The organization of ACT domains in the present structure is different from that observed in Thr deaminase and in the regulatory subunit of acetohydroxyacid synthase III.  相似文献   

9.
RumA catalyzes transfer of a methyl group from S-adenosylmethionine (SAM) specifically to uridine 1939 of 23S ribosomal RNA in Escherichia coli to yield 5-methyluridine. We determined the crystal structure of RumA at 1.95 A resolution. The protein is organized into three structural domains: The N-terminal domain contains sequence homology to the conserved TRAM motif and displays a five-stranded beta barrel architecture characteristic of an oligosaccharide/oligonucleotide binding fold. The central domain contains a [Fe(4)S(4)] cluster coordinated by four conserved cysteine residues. The C-terminal domain displays the typical SAM-dependent methyltransferase fold. The catalytic nucleophile Cys389 lies in a motif different from that in DNA 5-methylcytosine methyltransferases. The electrostatic potential surface reveals a predominately positively charged area that covers the concave surface of the first two domains and suggests an RNA binding mode. The iron-sulfur cluster may be involved in the correct folding of the protein or may have a role in RNA binding.  相似文献   

10.
Rat liver glycine methyltransferase, a homotetramer, exhibits sigmoidal rate behavior with respect to S-adenosylmethionine (Ogawa, H., and Fujioka, M. (1982) J. Biol. Chem. 257, 3447-3452). The binding experiment shows that the sigmoidicity observed in initial velocity kinetics is explained by the cooperative binding of S-adenosylmethionine to the catalytic sites residing on each subunit. Limited proteolysis of glycine methyltransferase with trypsin in the presence of S-adenosylmethionine yields an enzyme lacking the NH2-terminal 8 residues. The proteolytically modified enzyme retains a tetrameric structure. The truncated enzyme shows no cooperativity with respect to S-adenosylmethionine binding and kinetics. It has values of Vmax and Km for glycine identical to those of the native enzyme, but a 3-fold lower [S]0.5 value for S-adenosylmethionine. The proteolytic modification is without effect on the circular dichroism and fluorescence spectra. Furthermore, the protein fluorescence of the modified enzyme is quenched upon addition of S-adenosylmethionine to the same extent as observed with the native enzyme. These results suggest that a short NH2-terminal segment, which lies outside the active site, is important for communication between subunits.  相似文献   

11.
Ribosome biogenesis in eukaryotes requires the participation of a large number of ribosome assembly factors. The highly conserved eukaryotic nucleolar protein Nep1 has an essential but unknown function in 18S rRNA processing and ribosome biogenesis. In Saccharomyces cerevisiae the malfunction of a temperature-sensitive Nep1 protein (nep1-1(ts)) was suppressed by the addition of S-adenosylmethionine (SAM). This suggests the participation of Nep1 in a methyltransferase reaction during ribosome biogenesis. In addition, yeast Nep1 binds to a 6-nt RNA-binding motif also found in 18S rRNA and facilitates the incorporation of ribosomal protein Rps19 during the formation of pre-ribosomes. Here, we present the X-ray structure of the Nep1 homolog from the archaebacterium Methanocaldococcus jannaschii in its free form (2.2 A resolution) and bound to the S-adenosylmethionine analog S-adenosylhomocysteine (SAH, 2.15 A resolution) and the antibiotic and general methyltransferase inhibitor sinefungin (2.25 A resolution). The structure reveals a fold which is very similar to the conserved core fold of the SPOUT-class methyltransferases but contains a novel extension of this common core fold. SAH and sinefungin bind to Nep1 at a preformed binding site that is topologically equivalent to the cofactor-binding site in other SPOUT-class methyltransferases. Therefore, our structures together with previous genetic data suggest that Nep1 is a genuine rRNA methyltransferase.  相似文献   

12.
抗砷性微生物及其抗砷分子机制研究进展   总被引:4,自引:0,他引:4  
蔡林  王革娇 《微生物学通报》2009,36(8):1253-1259
砷(Arsenic, As)是一种剧毒类金属(Metalloid), 在自然环境中主要以三价亚砷酸盐[Arsenite, AsO2-, As(III)]和五价砷酸盐[Arsenate, AsO43-, As(V)]的无机形式广泛存在。许多微生物在含砷环境的长期适应过程中, 进化了多种不同的砷解毒抗性机制。目前研究发现主要存在4种类型的砷抗性机理, 包括: As(III)氧化, 细胞质As(V)还原, 呼吸性As(V)还原, As(III)甲基化, 这些机制赋予微生物砷抗性并在砷的转化和地球化学循环中起着极  相似文献   

13.
Marapakala K  Qin J  Rosen BP 《Biochemistry》2012,51(5):944-951
The enzyme As(III) S-adenosylmethionine methyltransferase (EC 2.1.1.137) (ArsM or AS3MT) is found in members of every kingdom, from bacteria to humans. In these enzymes, there are three conserved cysteine residues at positions 72, 174, and 224 in the CmArsM orthologue from the thermophilic eukaryotic alga Cyanidioschyzon sp. 5508. Substitution of any of the three led to loss of As(III) methylation. In contrast, a C72A mutant still methylated trivalent methylarsenite [MAs(III)]. Protein fluorescence of a single-tryptophan mutant reported binding of As(III) or MAs(III). As(GS)(3) and MAs(GS)(2) bound significantly faster than As(III), suggesting that the glutathionylated arsenicals are preferred substrates for the enzyme. Protein fluorescence also reported binding of Sb(III), and the purified enzyme methylated and volatilized Sb(III). The results suggest that all three cysteine residues are necessary for the first step in the reaction, As(III) methylation, but that only Cys174 and Cys224 are required for the second step, methylation of MAs(III) to dimethylarsenite [DMAs(III)]. The rate-limiting step was identified as the conversion of DMAs(III) to trimethylarsine, and DMAs(III) accumulates as the principal product.  相似文献   

14.
W C Lam  D H Tsao  A H Maki  K A Maegley  N O Reich 《Biochemistry》1992,31(43):10438-10442
The interactions of an arsenic (III) reagent, (CH3)2AsSCH2CONH2, with two Escherichia coli RI methyltransferase mutants, W183F and C223S, have been studied by phosphorescence, optically detected magnetic resonance, and fluorescence spectroscopy. The phosphorescence spectrum of the W183F mutant containing only one tryptophan at position 225 reveals a single 0,0-band that is red-shifted by 9.8 nm upon binding of As(III). Fluorescence titration of W183F with (CH3)2AsSCH2CONH2 produces a large tryptophan fluorescence quenching. Analysis of the quenching data points to a single high-affinity As(III) binding site that is associated with the fluorescence quenching. Triplet-state kinetic measurements performed on the perturbed tryptophan show large reductions in the lifetimes of the triplet sublevels, especially that of the T chi sublevel. As(III) binding to the enzyme at a site very close to the Trp225 residue induces an external heavy-atom effect, showing that the perturber atom is in van der Waals contact with the indole chromophore. In the case of the C223S mutant, a single tryptophan 0,0-band also is observed in the phosphorescence spectrum, but no change occurs upon addition of the As(III) reagent. Fluorescence titration of C223S with As(III) shows essentially no quenching of tryptophan fluorescence, in contrast with W183F. These results, along with previous triplet-state and biochemical studies on the wild-type enzyme [Tsao, D. H.H., & Maki, A. H. (1991) Biochemistry 30, 4565-4572], show that As(III) binds with high affinity to the Cys223 residue and that the Trp225 side chain is located close enough to that of Cys223 to produce a heavy-atom perturbation when As(III) is bound.  相似文献   

15.
The ars operon of plasmid R773 encodes an As(III)/Sb(III) extrusion pump. The catalytic subunit, the ArsA ATPase, has two homologous halves, A1 and A2, each with a consensus nucleotide-binding sequence. ATP hydrolysis is slow in the absence of metalloid and is accelerated by metalloid binding. ArsA M446W has a single tryptophan adjacent to the A2 nucleotide-binding site. Tryptophan fluorescence increased upon addition of ATP, ADP, or a nonhydrolyzable ATP analogue. Mg(2+) and Sb(III) produced rapid quenching of fluorescence with ADP, no quenching with a nonhydrolyzable analogue, and slow quenching with ATP. The results suggest that slow quenching with ATP reflects hydrolysis of ATP to ADP in the A2 nucleotide-binding site. In an A2 nucleotide-binding site mutant, nucleotides had no effect. In contrast, in an A1 nucleotide-binding mutant, nucleotides still increased fluorescence, but there was no quenching with Mg(2+) and Sb(III). This suggests that the A2 site hydrolyzes ATP only when Sb(III) or As(III) is present and when the A1 nucleotide-binding domain is functional. These results support previous hypotheses in which only the A1 nucleotide-binding domain hydrolyzes ATP in the absence of activator (unisite catalysis), and both the A1 and A2 sites hydrolyze ATP when activated (multisite catalysis).  相似文献   

16.
Inorganic arsenic is an established human carcinogen, but its metabolism is incompletely defined. The ATP binding cassette protein, multidrug resistance protein (MRP1/ABCC1), transports conjugated organic anions (e.g. leukotriene C(4)) and also co-transports certain unmodified xenobiotics (e.g. vincristine) with glutathione (GSH). MRP1 also confers resistance to arsenic in association with GSH; however, the mechanism and the species of arsenic transported are unknown. Using membrane vesicles prepared from the MRP1-overexpressing lung cancer cell line, H69AR, we found that MRP1 transports arsenite (As(III)) only in the presence of GSH but does not transport arsenate (As(V)) (with or without GSH). The non-reducing GSH analogs L-gamma-glutamyl-L-alpha-aminobutyryl glycine and S-methyl GSH did not support As(III) transport, indicating that the free thiol group of GSH is required. GSH-dependent transport of As(III) was 2-fold higher at pH 6.5-7 than at a more basic pH, consistent with the formation and transport of the acid-stable arsenic triglutathione (As(GS)(3)). Immunoblot analysis of H69AR vesicles revealed the unexpected membrane association of GSH S-transferase P1-1 (GSTP1-1). Membrane vesicles from an MRP1-transfected HeLa cell line lacking membrane-associated GSTP1-1 did not transport As(III) even in the presence of GSH but did transport synthetic As(GS)(3). The addition of exogenous GSTP1-1 to HeLa-MRP1 vesicles resulted in GSH-dependent As(III) transport. The apparent K(m) of As(GS)(3) for MRP1 was 0.32 microM, suggesting a remarkably high relative affinity. As(GS)(3) transport by MRP1 was osmotically sensitive and was inhibited by several conjugated organic anions (MRP1 substrates) as well as the metalloid antimonite (K(i) 2.8 microM). As(GS)(3) transport experiments using MRP1 mutants with substrate specificities differing from wild-type MRP1 suggested a commonality in the substrate binding pockets of As(GS)(3) and leukotriene C(4). Finally, human MRP2 also transported As(GS)(3). In conclusion, MRP1 transports inorganic arsenic as a tri-GSH conjugate, and GSTP1-1 may have a synergistic role in this process.  相似文献   

17.
Anamorsin is a recently identified molecule that inhibits apoptosis during hematopoiesis. It contains an N‐terminal methyltransferase‐like domain and a C‐terminal Fe‐S cluster motif. Not much is known about the function of the protein. To better understand the function of anamorsin, we have solved the crystal structure of the N‐terminal domain at 1.8 Å resolution. Although the overall structure resembles a typical S‐adenosylmethionine (SAM) dependent methyltransferase fold, it lacks one α‐helix and one β‐strand. As a result, the N‐terminal domain as well as the full‐length anamorsin did not show S‐adenosyl‐l ‐methionine (AdoMet) dependent methyltransferase activity. Structural comparisons with known AdoMet dependent methyltransferases reveals subtle differences in the SAM binding pocket that preclude the N‐terminal domain from binding to AdoMet. The N‐terminal methyltransferase‐like domain of anamorsin probably functions as a structural scaffold to inhibit methyl transfers by out‐competing other AdoMet dependant methyltransferases or acts as bait for protein–protein interactions.Proteins 2014; 82:1066–1071. © 2013 Wiley Periodicals, Inc.  相似文献   

18.
Microbial methylation of metalloids: arsenic, antimony, and bismuth.   总被引:5,自引:0,他引:5  
A significant 19th century public health problem was that the inhabitants of many houses containing wallpaper decorated with green arsenical pigments experienced illness and death. The problem was caused by certain fungi that grew in the presence of inorganic arsenic to form a toxic, garlic-odored gas. The garlic odor was actually put to use in a very delicate microbiological test for arsenic. In 1933, the gas was shown to be trimethylarsine. It was not until 1971 that arsenic methylation by bacteria was demonstrated. Further research in biomethylation has been facilitated by the development of delicate techniques for the determination of arsenic species. As described in this review, many microorganisms (bacteria, fungi, and yeasts) and animals are now known to biomethylate arsenic, forming both volatile (e.g., methylarsines) and nonvolatile (e.g., methylarsonic acid and dimethylarsinic acid) compounds. The enzymatic mechanisms for this biomethylation are discussed. The microbial conversion of sodium arsenate to trimethylarsine proceeds by alternate reduction and methylation steps, with S-adenosylmethionine as the usual methyl donor. Thiols have important roles in the reductions. In anaerobic bacteria, methylcobalamin may be the donor. The other metalloid elements of the periodic table group 15, antimony and bismuth, also undergo biomethylation to some extent. Trimethylstibine formation by microorganisms is now well established, but this process apparently does not occur in animals. Formation of trimethylbismuth by microorganisms has been reported in a few cases. Microbial methylation plays important roles in the biogeochemical cycling of these metalloid elements and possibly in their detoxification. The wheel has come full circle, and public health considerations are again important.  相似文献   

19.
Microbial Methylation of Metalloids: Arsenic, Antimony, and Bismuth   总被引:10,自引:0,他引:10       下载免费PDF全文
A significant 19th century public health problem was that the inhabitants of many houses containing wallpaper decorated with green arsenical pigments experienced illness and death. The problem was caused by certain fungi that grew in the presence of inorganic arsenic to form a toxic, garlic-odored gas. The garlic odor was actually put to use in a very delicate microbiological test for arsenic. In 1933, the gas was shown to be trimethylarsine. It was not until 1971 that arsenic methylation by bacteria was demonstrated. Further research in biomethylation has been facilitated by the development of delicate techniques for the determination of arsenic species. As described in this review, many microorganisms (bacteria, fungi, and yeasts) and animals are now known to biomethylate arsenic, forming both volatile (e.g., methylarsines) and nonvolatile (e.g., methylarsonic acid and dimethylarsinic acid) compounds. The enzymatic mechanisms for this biomethylation are discussed. The microbial conversion of sodium arsenate to trimethylarsine proceeds by alternate reduction and methylation steps, with S-adenosylmethionine as the usual methyl donor. Thiols have important roles in the reductions. In anaerobic bacteria, methylcobalamin may be the donor. The other metalloid elements of the periodic table group 15, antimony and bismuth, also undergo biomethylation to some extent. Trimethylstibine formation by microorganisms is now well established, but this process apparently does not occur in animals. Formation of trimethylbismuth by microorganisms has been reported in a few cases. Microbial methylation plays important roles in the biogeochemical cycling of these metalloid elements and possibly in their detoxification. The wheel has come full circle, and public health considerations are again important.  相似文献   

20.
Processive versus distributive methyl group transfer was assessed for pea Rubisco large subunit methyltransferase, a SET domain protein lysine methyltransferase catalyzing the formation of trimethyllysine-14 in the large subunit of Rubisco. Catalytically competent complexes between an immobilized form of des(methyl) Rubisco and Rubisco large subunit methyltransferase were used to demonstrate enzyme release that was co-incident with and dependent on formation of trimethyllysine. Catalytic rate constants determined for formation of trimethyllysine were considerably lower ( approximately 10-fold) than rate constants determined for total radiolabel incorporation from [3H-methyl]-S-adenosylmethionine. Double-reciprocal velocity plots under catalytic conditions favoring monomethyllysine indicated a random or ordered reaction mechanism, while conditions favoring trimethyllysine suggested a hybrid ping-pong mechanism. These results were compared with double-reciprocal velocity plots and product analyses obtained for HsSET7/9 (a monomethyltransferase) and SpCLR4 (a dimethyltransferase) and suggest a predictive ability of double-reciprocal velocity plots for single versus multiple methyl group transfers by SET domain protein lysine methyltransferases. A model is proposed for SET domain protein lysine methyltransferases in which initial binding of polypeptide substrate and S-adenosylmethionine is random, with polypeptide binding followed by deprotonation of the epsilon-amine of the target lysyl residue and subsequent methylation. Following methyl group transfer, S-adenosylhomocysteine and monomethylated polypeptide dissociate from monomethyltransferases, but di- and trimethyltransferases begin a successive and catalytically obligatory deprotonation of enzyme-bound methylated lysyl intermediates, which along with binding and release of S-adenosylmethionine and S-adenosylhomocysteine is manifested as a hybrid ping-pong-like reaction mechanism.  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