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 31 毫秒
1.
Oxidation of amino acid residues in proteins can be caused by a variety of oxidizing agents normally produced by cells. The oxidation of methionine in proteins to methionine sulfoxide is implicated in aging as well as in pathological conditions, and it is a reversible reaction mediated by a ubiquitous enzyme, peptide methionine sulfoxide reductase. The reversibility of methionine oxidation suggests that it could act as a cellular regulatory mechanism although no such in vivo activity has been demonstrated. We show here that oxidation of a methionine residue in a voltage-dependent potassium channel modulates its inactivation. When this methionine residue is oxidized to methionine sulfoxide, the inactivation is disrupted, and it is reversed by coexpression with peptide methionine sulfoxide reductase. The results suggest that oxidation and reduction of methionine could play a dynamic role in the cellular signal transduction process in a variety of systems.  相似文献   

2.
3.
Methionine residues as endogenous antioxidants in proteins   总被引:3,自引:0,他引:3  
Cysteine and methionine are the two sulfur-containing residues normally found in proteins. Cysteine residues function in the catalytic cycle of many enzymes, and they can form disulfide bonds that contribute to protein structure. In contrast, the specific functions of methionine residues are not known. We propose that methionine residues constitute an important antioxidant defense mechanism. A variety of oxidants react readily with methionine to form methionine sulfoxide, and surface exposed methionine residues create an extremely high concentration of reactant, available as an efficient oxidant scavenger. Reduction back to methionine by methionine sulfoxide reductases would allow the antioxidant system to function catalytically. The effect of hydrogen peroxide exposure upon glutamine synthetase from Escherichia coli was studied as an in vitro model system. Eight of the 16 methionine residues could be oxidized with little effect on catalytic activity of the enzyme. The oxidizable methionine residues were found to be relatively surface exposed, whereas the intact residues were generally buried within the core of the protein. Furthermore, the susceptible residues were physically arranged in an array that guarded the entrance to the active site.  相似文献   

4.
Intact methionine residues in proteins were rapidly and precisely determined by measuring methyl thiocyanate released during the reaction with CNBr and separated by g.l.c. Conditions for the reaction and for chromatography on columns of Porapak P-S are described. The recovery of methyl thiocyanate from several methionine derivatives and analogues were examined. Carbamoylmethionine was adopted as a stable primary standard and ethyl thiocyanate as internal standard. The measured methionine content of several isolated proteins was close to the theoretical value indicated by previous work and the results for these and a range of food proteins agreed well with results obtained by ion-exchange chromatography after performic acid oxidation. Since CNBr does not react with methionine sulphoxide and a preliminary hydrolysis is not required, the method discriminates between methionine and any methionine sulphoxide that may be present. It could be useful in studies on the nutritional availability of methionine in processed foods.  相似文献   

5.
Whereas vertebrates possess only two thioredoxin genes, higher plants present a much greater diversity of thioredoxins. For example, Arabidopsis thaliana has five cytoplasmic thioredoxins (type h) and at least as many chloroplastic thioredoxins. The abundance of plant thioredoxins leads to the question whether the various plant thioredoxins play a similar role or have specific functions. Because most of these proteins display very similar activities on artificial or biological substrates in vitro, we developed an in vivo approach to answer this question. The disruption of both of the two Saccharomyces cerevisiae thioredoxin genes leads to pleiotropic effects including methionine auxotrophy, H2O2 hypersensitivity, altered cell cycle characteristics, and a limited ability to use methionine sulfoxide as source of methionine. We expressed eight plant thioredoxins (six cytoplasmic and two chloroplastic) in yeast trx1, trx2 double mutant cells and analyzed the different phenotypes. Arabidopsis type h thioredoxin 2 efficiently restored sulfate assimilation whereas Arabidopsis type h thioredoxin 3 conferred H2O2 tolerance. All thioredoxins tested could complement for reduction of methionine sulfoxide, whereas only type h thioredoxins were able to complement the cell cycle defect. These findings clearly indicate that specific interactions between plant thioredoxins and their targets occur in vivo.  相似文献   

6.
Jones et al. have reported that the epsilon-carbons of methionine residues in myoglobin can be enriched with stable isotope (13C) in two steps, i.e., methylation of methionine residues with 13CH3I in the protein and thiolysis using dithiothreitol [Jones, W.C., Rothgeb, T.M., and Gurd, F.R.N. (1976) J. Biol. Chem. 251,7452-7460]. Using their method, we failed to prepare active lysozyme in which the epsilon-carbons of methionine residues are enriched with 13C, because many side reactions took place under the thiolysis condition (pH 10.5, 37 degrees C). When we employed 2-aminothiophenol as a reagent for thiolysis, the reduction proceeded under a weakly acidic condition to afford fully active lysozyme, in which the epsilon-carbons of two methionine residues were enriched with 13C, in a 30% yield. Analysis of the 13C-edited NOESY spectra of 13C-enriched methionine lysozyme in the absence and presence of a substrate analogue indicated the occurrence of conformational change around Met 105 in lysozyme.  相似文献   

7.
An experiment was conducted to determine the utilization of ileal digestible methionine by growing pigs. Three methionine-deficient diets (0.09 g ileal digestible methionine/MJ digestible energy (DE)) were formulated using cottonseed meal, meat-and-bone meal and soya-bean meal respectively as the only source of methionine in the diet. An additional three diets were formulated with supplements of methionine to confirm that methionine was limiting in the first three diets. The growth performance and retention of methionine by pigs given the six diets over the 20-45 kg growth phase was then determined. Growth rates (g/d) of pigs given the three diets formulated to 0.09 g ileal digestible methionine/MJ DE were significantly different (P < 0.01): cottonseed meal 411, meat-and-bone meal 442, soya-bean meal 496 (SED 24.6). The response of pigs to the addition of methionine confirmed that methionine was limiting in these diets. Crude protein (N x 6.25) deposited by the pigs (g/d) was significantly higher (P < 0.05) for those given soya-bean meal (61) and meat-and-bone meal (57) relative to cottonseed meal (47; SED 3.3). The proportion of ileal digestible methionine retained by pigs given the three protein concentrates was: cottonseed meal 0.39, meat-and-bone meal 0.45, soya-bean meal 0.47 (SED 0.019). These results indicate that values for the ileal digestibility of methionine in protein concentrates do not reflect the proportion of methionine that can be utilized by the pig. It appears that, with heat-processed meals, a considerable proportion of the methionine is absorbed in a form(s) that is (are) inefficiently utilized.  相似文献   

8.
Experimental lots of 12 chicks each received, in addition to their standard diets, supplements of 1, 2 or 3% DL-threonine, in each threonine treatment 1, 2 or 3% methionine being added. After an 8-day feeding period the chicks were injected 35S-methionine and were killed after 6, 12 and 24 hours, respectively. The increasing methionine supplementation was found to markedly deteriorate chick growth. It was not possible to correct this effect by any of the threonine doses. The 35S-activity of the blood decreased very strongly from the 6th to the 12th hours following injection. The 35S-activity of the liver also declined clearly at low threonine supply, but less markedly at high threonine rates. In the breast muscle, the 35S-activity also proved three times higher 24 hours after the injection at equal methionine supplementation to the diet (1 or 2%) and at triplicated threonine level (3% against 1%). This finding is related with the fact that threonine supplements to not attenuate the negative imbalance symptoms after methionine overdosing, but that they are likely to aggrevate due to transport competition. The possibility of threonine degradation into glycine does not seem to be of importance for methionine detoxication in chicks.  相似文献   

9.
Methionine dependence is a metabolic defect that occurs in many human tumor cell lines but not normal in unestablished cell strains. Methionine-dependent tumor cell lines are unable to proliferate and arrest in the late S/G2 phase of the cell cycle when methionine is replaced by its immediate precursor homocysteine in the culture medium (MET-HCY+ medium). However, it is not known whether methionine dependence occurs in fresh patient tumors as it does in cell lines. In order to determine whether methionine dependence occurs in fresh patient tumors as well as whether methionine dependence occurs in fresh patient tumors as well as in cell lines we took advantage of the technique of sponge-gel-supported histoculture to grow tumors directly from surgery. We then measured nuclear DNA content by image analysis to determine the cell cycle position in MET-HCY+ compared to MET+HCY- medium in 21 human patient tumors. Human tumor cell lines found to be methionine dependent by cell count were used as positive controls and were found to have marked reduction of cells in G1 compared to total cells in the cell cycle in MET-HCY+ medium with respect to the G1: total cell ratio in MET+HCY- medium. Therefore late cell cycle arrest was used as a marker of methionine dependence for histocultured patient tumors. We found that 5 human tumors of 21, including tumors of the colon, breast, ovary, prostate, and a melanoma, were methionine dependent based on cell cycle analysis. These data on fresh human tumors indicate that methionine dependence may frequently occur in the cancer patient population. Implications for potential therapy based on methionine dependence are discussed.  相似文献   

10.
The activity of glutamine synthetase (EC 6.3.1.2) from Escherichia coli is regulated by the cyclic adenylylation and deadenylylation of Tyr-397 in each of the enzyme's 12 identical subunits. The nitration of Tyr-397 or of the nearby Tyr-326 by peroxynitrite can convert the unadenylylated enzyme to a form exhibiting regulatory characteristics similar to the form obtained by adenylylation. The adenylylated conformation can also be elicited by the oxidation of surface-exposed methionine residues to methionine sulfoxide. However, the nitration of tyrosine residues and the oxidation of methionine residues are oppositely directed by the presence and absence of CO2. At physiological concentrations of CO2, pH 7.4, nitration occurs but oxidation of methionine residues is inhibited. Conversely, in the absence of CO2 methionine oxidation is stimulated and nitration of tyrosine is prevented. It was further established that adenylylation of Tyr-397 precludes its nitration by peroxynitrite. Furthermore, nitration of Tyr-326 together with either nitration or adenylylation of Tyr-397 leads to inactivation of the enzyme. These results demonstrate that CO2 can alter the course of peroxynitrite-dependent reactions and serve notice that (i) the reactions have physiological significance only if they are shown to occur at physiological concentrations of CO2 and physiological pH; and (ii) the peroxynitrite-dependent nitration of tyrosine residues or the oxidation of methionine residues of metabolically regulated proteins can seriously compromise their biological function.  相似文献   

11.
Nucleotides in tRNAs that are conserved among isoacceptors are typically considered as candidates for tRNA synthetase recognition, with less importance attached to non-conserved nucleotides. Although the anticodon is an important contributor to the identity of methionine tRNAs, the class I methionine tRNA synthetase aminoacylates microhelices with high specificity. The microhelix substrates are comprised of as few as the 1st 4 base pairs of the acceptor stems of the elongator and initiator methionine tRNAs. For these two tRNAs, only the central 2:71 and 3:70 base pairs are common to the 1st 4 acceptor stem base pairs. We show here that, although the flanking 4:69 base pair is not conserved, a particular substitution at this position substantially reduces the gel electrophoresis-detected aminoacylation of an acceptor stem substrate that has the conserved 2:71 and 3:70 base pairs. Although the two methionine tRNAs have either U:A or G:C at position 4:69, substitution with C:G reduces charging of 9- or 4-base pair substrates that recreate part or all of the acceptor stem of a methionine tRNA. This effect is sufficient for methionine tRNA synthetase to discriminate between the closely related methionine and isoleucine tRNA acceptor stems. The ability to distinguish G:C and U:A from C:G is contrary to a simple scheme for recognition of atoms in the RNA minor groove.  相似文献   

12.
Methionine metabolism and transmethylation are central to the metabolism and differentiation of all known cells. In enkaryotic organisms, methionine metabolism and transmethylation are of paramount importance in modification and regulation of proteins, lipids, and nucleic acids. The differential methylation of genes regulates their expression in the myriad of cells in eukaryotic organisms. Disruption and abnormalities in methionine metabolism and transmethylation seems to be associated with the major diseases of mankind, including cancer, heart disease, aging, obesity, and Parkinson's disease. In this review, we describe how aberrant and abnormal methionine metabolism and transmethylation are related to these major diseases. Most importantly, we review and hypothesize how the developing therapeutic recombination methioninase (rMETase) can be utilized to cure or prevent all of these diseases.  相似文献   

13.
1. The fractional rate of loss of 14C and body-weight was measured in adult male rats after giving 14C-labelled methionine or leucine and maintaining rats for 30 d on a low-protein or a specific methionine+cystine-free diet: carcasses were then analysed for protein and fat 14C radioactivity. 2. The fractional loss of 14CO2 from [14C]methionine or [14C]leucine between day 20 and day 30 was always greater than the fractional loss of body-weight. 3. Carcass protein 14C radioactivity after giving [14C]leucine was higher than after giving [14C]methionine, but fat 14C radioactivity after either 14C-labelled amino acid was only a small proportion of the total body 14C radioactivity. 4. After correction of the fractional loss of 14CO2 for urinary 14C loss, but not body-weight loss, absolute amino acid loss was calculated using published values for methionine and leucine content of rats. 5. The best estimates of endogenous amino acid loss obtained using I-14C-labelled amino acids, expressed as mg/kg body-weight 0.75 per day were leucine 79, methionine 38.  相似文献   

14.
Experiments in vitro using rat liver slices indicated that the incorporation of the methionine methyl carbon into sarcosine and serine was dependent upon available glycine and most probably involves glycine methyltransferase. Although the sarcosine methyl carbon was rapidly oxidized to CO2, its formation accounted for only a small proportion of the oxidation of the methionine methyl carbon to CO2 under these conditions. In vivo experiments using a sarcosine trapping pool with 0.3% to 3.0% L-[methyl-14C]methionine in the diet indicated that from 5% to 14% of the absorbed methionine methyl carbon was metabolized via sarcosine, and that this accounted for only 10% to 20% of the observed oxidation of the methyl carbon to CO2. The adaptive response of the rat to high levels of dietary methionine, as indicated by greater oxidation of the methyl carbon to CO2, is in part due to increased sarcosine synthesis. The failure of supplemental glycine to stimulate oxidation of the methionine methyl carbon to CO2 in rats receiving 3% methionine plus 10% sarcosine may be due to sufficient glycine being produced from sarcosine metabolism.  相似文献   

15.
Experiments conducted with rats in vivo comparing the metabolism of methionine and S-methyl-L-cysteine and in vitro comparing methionine, S-methyl-L-cysteine and S-adenosyl-L-methionine indicate that a substantial portion of the oxidative metabolism of the methionine methyl group occurs by pathways that are independent of S-adenosylmethionine formation. Inclusion of 1.2% or 2.4% of S-methyl-L-cysteine in a diet containing 3% of L-methionine depressed the conversion of the methionine methyl and carboxyl carbons to CO2 by 39% and 28%, and 52% and 33%, respectively, for the two levels of S-methyl-L-cysteine. Inclusion of 1.65% of methionine in a diet containing 2.4% of S-methyl-L-cysteine did not affect the conversion of the methyl group of S-methylcysteine to CO2, but 3% of methionine depressed the conversion of the S-methylcysteine methyl group to CO2 to 87% of control values. Greater inhibitions were seen when these substrates were compared in a liver homogenate. In a rat liver homogenate system optimized for the conversion of the methyl group of methionine to CO2, the rate of conversion of the methyl group of S-adenosyl-L-methionine to CO2 was less than 1% of that of methionine even when the concentration of S-adenosylmethionine was saturating. Addition of saturating levels of unlabeled S-adenosymethionine to the homogenate system did not effect the rate of conversion of the methionine methyl carbon to CO2. Although S-adenosylmethionine-dependent metabolism of methionine, leading to incorporation of the methyl carbon into sarcosine and serine, could be demonstrated in liver homogenates, essentially all of the CO2 produced from the methionine methyl group was derived by a pathway or pathways independent of S-adenosylmethionine formation. Formaldehyde and formate have been tentatively identified as intermediates in catabolism of the methionine methyl group by this (these) pathway(s).  相似文献   

16.
Methionine synthase catalyzes the remethylation of homocysteine to methionine via a reaction in which methylcobalamin serves as an intermediate methyl carrier. Over time, the cob(I)alamin cofactor of methionine synthase becomes oxidized to cob(II)alamin rendering the enzyme inactive. Regeneration of functional enzyme requires reductive methylation via a reaction in which S-adenosylmethionine is utilized as a methyl donor. Patients of the cblE complementation group of disorders of folate/cobalamin metabolism who are defective in reductive activation of methionine synthase exhibit megaloblastic anemia, developmental delay, hyperhomocysteinemia, and hypomethioninemia. Using consensus sequences to predicted binding sites for FMN, FAD, and NADPH, we have cloned a cDNA corresponding to the "methionine synthase reductase" reducing system required for maintenance of the methionine synthase in a functional state. The gene MTRR has been localized to chromosome 5p15.2-15.3. A predominant mRNA of 3.6 kb is detected by Northern blot analysis. The deduced protein is a novel member of the FNR family of electron transferases, containing 698 amino acids with a predicted molecular mass of 77,700. It shares 38% identity with human cytochrome P450 reductase and 43% with the C. elegans putative methionine synthase reductase. The authenticity of the cDNA sequence was confirmed by identification of mutations in cblE patients, including a 4-bp frameshift in two affected siblings and a 3-bp deletion in a third patient. The cloning of the cDNA will permit the diagnostic characterization of cblE patients and investigation of the potential role of polymorphisms of this enzyme as a risk factor in hyperhomocysteinemia-linked vascular disease.  相似文献   

17.
The keto acid 2-oxo-4[methylthio]butanoic acid (OMTB) is an intermediate in the conversion of synthetic feed grade methionine sources to L-methionine in vivo in poultry and other animals. Because methionine sources are utilized by the chick with considerably less than 100% efficiency as sources of L-methionine, it is important to determine what metabolic process may limit the utilization of these sources. Because OMTB is converted to L-methionine by transamination, a study was conducted to determine which amino acids might serve as nitrogen donors in the conversion of OMTB to L-methionine in the chicken. Dialyzed tissue homogenates, mitochondria, and cytosol from liver, kidney, intestine, and skeletal muscle were incubated with OMTB and individual L-amino acids (isoleucine, leucine, valine, glutamic acid, aspartic acid, alanine, glutamine, asparagine, and phenylalanine) and the methionine that accumulated was determined by ion exchange chromatography. Tissues differed in the conversion of OMTB to methionine: kidney was most active, liver and intestinal mucosa were intermediate, and skeletal muscle had lowest activity. All amino acids supported methionine synthesis. Branched-chain amino acids and glutamic acid were the most effective substrates in tissue cytosols except in intestinal mucosa, in which asparagine was also effective. The preferred substrates in mitochondria were glutamate in liver mitochondria, isoleucine and alanine in kidney mitochondria, and branched-chain amino acids and glutamic acid in skeletal muscle mitochondria. All amino acids except alanine supported methionine synthesis from OMTB in mitochondria of intestinal mucosa. We conclude that a wide variety of amino acids can serve as substrates for transamination of OMTB in the chicken, and that the availability of nitrogen donors is unlikely to be a limiting factor in the conversion of OMTB to methionine.  相似文献   

18.
The renal transport and fractional reabsorption of inorganic sulfate is altered under conditions of sulfate deficiency or excess. The objective of this study was to examine the cellular mechanisms of adaptation of renal sodium/sulfate cotransport after varying dietary intakes of a sulfur containing amino acid, methionine. Female Lewis rats were divided into four groups and fed diets containing various concentrations of methionine (0, 0.3, 0.82 and 2.46%) for 8 days. Urinary excretion rates and renal clearance of sulfate were significantly decreased in the animals fed a 0% methionine diet or a 0.3% methionine diet, and significantly increased in the animals fed a 2.46% methionine diet when evaluated on days 4 and 7. Serum sulfate concentrations were unchanged by diet treatment in all animals. The fractional reabsorption of sulfate was significantly increased in the animals fed the 0% methionine diet and the 0.3% methionine diets, and decreased in the animals fed the 2.46% methionine diet. Increased mRNA and protein levels for the sodium/sulfate transporter (NaSi-1) were found in the kidney cortex following treatment with the 0 and 0.3% methionine diet groups. Sulfate homeostasis by renal reabsorption is maintained by an up-regulation of steady state levels of NaSi-1 mRNA and protein when the diet is low in methionine.  相似文献   

19.
1. The isolated perfused rat liver and suspensions of isolated rat hepatocytes fail to form glucose from histidine, in contrast with the liver in vivo. Both rat liver preparations readily metabolize histidine. The main end product is N-formiminoglutamate. In this respect the liver preparations behave like the liver of cobalamin- or folate-deficient mammals. 2. Additions of L-methionine in physiological concentrations (or of ethionine [2-amino-4-(ethylthio)butyric acid]) promotes the degradation of formiminoglutamate, as is already known to be the case in cobalamin of folate deficiency. Added methionine also promotes glucose formation from histidine. 3. Addition of methionine accelerates the oxidation of formate to bicarbonate by hepatocytes. 4. A feature common to cobalamin-deficient liver and the isolated liver preparations is taken to be a low tissue methionine concentration, to be expected in cobalamin deficiency through a decreased synthesis of methionine and caused in liver preparations by a washing out of amino acids during the handling of the tissue. 5. The available evidence is in accordance with the assumption that methionine does not directly increase the catalytic capacity of formyltetrahydrofolate dehydrogenase; rather, that an increased methionine concentration raises the concentration of S-adenosylmethionine, thus leading to the inhibition of methylenetetrahydrofolate reductase activity [Kutzbach & Stokstad (1967) Biochim. Biophys. Acta 139, 217-220; Kutzbach & Stokstad (1971) Methods Enzymol. 18B, 793-798], that this inhibition causes an increase in the concentration of methylenetetrahydrofolate and the C1 tetrahydrofolate derivatives in equilibrium with methylenetetrahydrofolate, including 10-formyltetrahydrofolate; that the increased concentration of the latter accelerates the formyltetrahydrofolate dehydrogenase reaction, because the normal concentration of the substrate is far below the Km value of the enzyme for the substrate. 6. The findings are relevant to the understanding of the regulation of both folate and methionine metabolism. When the methionine concentration is low, C1 units are preserved by the decreased activity of formyltetrahydrofolate dehydrogenase and are utilized for the synthesis of methionine, purines and pyrimidines. On the other hand when the concentration of methionine, and hence adenosylmethionine, is high and there is a surplus of C1 units as a result of excess of dietary supply, formyltetrahydrofolate dehydrogenase disposes of the excess. When ample dietary supply causes an excess of methionine, which has to be disposed of by degradation, the increased activity of formyltetrahydrofolate dehydrogenase decreases the supply of methyltetrahydrofolate. Thus homocysteine, instead of being remethylated, enters the pathway of degradation via cystathionine. 7. The findings throw light on the biochemical abnormalities associated with cobalamin deficiency (megaloblastic anaemia), especially on the 'methylfolate-trap hypothesis'. This is discussed. 8...  相似文献   

20.
This paper reviews current knowledge regarding the metabolism of the sulphur-containing amino acids methionine and cysteine in parasitic protozoa and helminths. Particular emphasis is placed on the unusual aspects of parasite biochemistry which may present targets for rational design of antiparasite drugs. In general, the basic pathways of sulphur amino acid metabolism in most parasites resemble those of their mammalian hosts, since the enzymes involved in (a) the methionine cycle and S-adenosylmethionine metabolism, (b) the trans-sulphuration sequence, (c) the transminative catabolism of methionine, (d) the oxidative catabolism of cysteine and (e) glutathione synthesis have been demonstrated variously in several helminth and protozoan species. Despite these common pathways, there also exist numerous differences between parasite and mammalian metabolism. Some of these differences are relatively subtle. For example, the biochemical properties (and primary amino acid structures) of certain parasite methionine cycle enzymes and S-adenosylmethionine decarboxylases differ from those of the corresponding mammalian enzymes, and nematodes and trichomonads possess a novel, non-mammalian form of the trans-sulphuration enzyme cystathionine beta-synthase. The most profound differences between parasite and mammalian biochemistry relate to a number of unusual enzymes and thiol metabolites found in parasitic protozoa. In certain protozoa the pathway for methionine recycling from 5'-methylthioadenosine differs markedly from the mammalian route, and involves 2 exclusively microbial enzymes. Trypanosomatid protozoa contain the non-mammalian antioxidant thiol compounds ovothiol A and trypanothione, together with unique trypanothione-linked enzymes. Specific anaerobic protozoa possess another exclusively microbial enzyme, methionine gamma-lyase, which catabolises methionine (and homocysteine); the physiological significance of these non-mammalian activities is not fully understood. These unusual features offer opportunities for chemotherapeutic exploitation, and in some cases represent metabolic similarities with bacteria. Additionally, some anaerobic protozoa contain unidentified thiols and this implies the presence of further unusual enzymes/pathways in these organisms. So far, no truly unique targets for chemotherapy have been found in helminth sulphur amino acid metabolism, and to some degree this reflects the relative lack of detailed study in the area.  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