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.
Phosphorylated ERK2 has an increased capacity to form homodimers relative to unphosphorylated ERK2. We have characterized the nature of the ERK2 dimer and have mutated residues in the crystal dimer interface to examine the impact of dimerization on ERK2 activity. Analysis of the mutants by gel filtration indicates that at least five residues must be mutated simultaneously to produce an ERK2 mutant that is predominantly monomeric. Mutants, whether monomers or dimers, have specific protein kinase activities under fixed assay conditions that are roughly equivalent to wild-type ERK2. The ratio of dimers to monomers is increased as the salt concentration increases, consistent with a strong hydrophobic contribution to the energy of dimer formation. ERK2 dimerization also requires divalent cations. Sedimentation analysis indicates that the related c-Jun N-terminal kinase SAPKalphaI/JNK2 also forms dimers, but dimerization displays no dependence on phosphorylation; the unphosphorylated and phosphorylated forms of the kinase behave similarly, with low micromolar dimer dissociation constants.  相似文献   

2.
FK506‐binding protein 22 (FKBP22) from the psychrotophic bacterium Shewanella sp. SIB1 (SIB1 FKBP22) is a homodimeric protein with peptidyl prolyl cis‐trans isomerase (PPIase) activity. Each monomer consists of the N‐terminal domain responsible for dimerization and C‐terminal catalytic domain. To reveal interactions at the dimer interface of SIB1 FKBP22, the crystal structure of the N‐domain of SIB1 FKBP22 (SN‐FKBP22, residues 1‐68) was determined at 1.9 Å resolution. SN‐FKBP22 forms a dimer, in which each monomer consists of three helices (α1, α2, and α3N). In the dimer, two monomers have head‐to‐head interactions, in which residues 8–64 of one monomer form tight interface with the corresponding residues of the other. The interface is featured by the presence of a Val‐Leu knot, in which Val37 and Leu41 of one monomer interact with Val41 and Leu37 of the other, respectively. To examine whether SIB1 FKBP22 is dissociated into the monomers by disruption of this knot, the mutant protein V37R/L41R‐FKBP22, in which Val37 and Leu41 of SIB1 FKBP22 are simultaneously replaced by Arg, was constructed and biochemically characterized. This mutant protein was indistinguishable from the SIB1 FKBP22 derivative lacking the N‐domain in oligomeric state, far‐UV CD spectrum, thermal denaturation curve, PPIase activity, and binding ability to a folding intermediate of protein, suggesting that the N‐domain of V37R/L41R‐FKBP22 is disordered. We propose that a Val‐Leu knot at the dimer interface of SIB1 FKBP22 is important for dimerization and dimerization is required for folding of the N‐domain.  相似文献   

3.
N‐acetylglucosamine 6‐phosphate deacetylase (NagA) catalyzes the conversion of N‐acetylglucosamine‐6‐phosphate to glucosamine‐6‐phosphate in amino sugar catabolism. This conversion is an essential step in the catabolism of sialic acid in several pathogenic bacteria, including Pasteurella multocida, and thus NagA is identified as a potential drug target. Here, we report the unique structural features of NagA from P. multocida (PmNagA) resolved to 1.95 Å. PmNagA displays an altered quaternary architecture with unique interface interactions compared to its close homolog, the Escherichia coli NagA (EcNagA). We confirmed that the altered quaternary structure is not a crystallographic artifact using single particle electron cryo‐microscopy. Analysis of the determined crystal structure reveals a set of hot‐spot residues involved in novel interactions at the dimer‐dimer interface. PmNagA binds to one Zn2+ ion in the active site and demonstrates kinetic parameters comparable to other bacterial homologs. Kinetic studies reveal that at high substrate concentrations (~10‐fold the KM), the tetrameric PmNagA displays hysteresis similar to its distant neighbor, the dimeric Staphylococcus aureus NagA (SaNagA). Our findings provide key information on structural and functional properties of NagA in P. multocida that could be utilized to design novel antibacterials.  相似文献   

4.
Diederik J. Opperman 《Proteins》2017,85(12):2252-2257
Ene‐reductases (ERs), or Old Yellow Enzymes, catalyze the asymmetric reduction of various activated alkenes. This class of biocatalysts is considered an attractive alternative to current chemical technologies for hydrogenation due to their high selectivity and specificity. Here the X‐ray crystal structure of RmER, a “thermophilic”‐like ER from Ralstonia (Cupriavidus) metallidurans, is reported. Unlike other members of this class of ERs, RmER is monomeric in solution which we previously related to its atypical elongated C‐terminus. A typical dimer interface was however observed in our crystal structure, with the conserved Arg‐“finger” forming part of the adjacent monomer's active site and the elongated C‐terminus extending into the active site through contacting the “capping” domain. This dimerization also resulted in the loss of one FMN cofactor from each dimer pair. This potential transient dimerization and dissociation of FMN could conceivably explain the rapid rates previously observed when an FMN light‐driven cofactor regeneration system was used during catalysis with RmER.  相似文献   

5.
Crystallographic structures and experimental assays of human CXC chemokine receptor type 4 (CXCR4) provide strong evidence for the capacity to homodimerize, potentially as a means of allosteric regulation. Even so, how this homodimer forms and its biological significance has yet to be fully characterized. By applying principles from network analysis, sequence‐based approaches such as statistical coupling analysis to determine coevolutionary residues, can be used in conjunction with molecular dynamics simulations to identify residues relevant to dimerization. Here, the predominant coevolution sector lies along the observed dimer interface, suggesting functional relevance. Furthermore, coevolution scoring provides a basis for determining significant nodes, termed hubs, in the network formed by residues found along the interface of the homodimer. These node residues coincide with hotspots indicating potential druggability. Drug design efforts targeting such key residues could potentially result in modulation of binding and therapeutic benefits for disease states, such as lung cancers, lymphomas and latent HIV‐1 infection. Furthermore, this method may be applied to any protein–protein interaction.  相似文献   

6.
Computational prediction of side‐chain conformation is an important component of protein structure prediction. Accurate side‐chain prediction is crucial for practical applications of protein structure models that need atomic‐detailed resolution such as protein and ligand design. We evaluated the accuracy of eight side‐chain prediction methods in reproducing the side‐chain conformations of experimentally solved structures deposited to the Protein Data Bank. Prediction accuracy was evaluated for a total of four different structural environments (buried, surface, interface, and membrane‐spanning) in three different protein types (monomeric, multimeric, and membrane). Overall, the highest accuracy was observed for buried residues in monomeric and multimeric proteins. Notably, side‐chains at protein interfaces and membrane‐spanning regions were better predicted than surface residues even though the methods did not all use multimeric and membrane proteins for training. Thus, we conclude that the current methods are as practically useful for modeling protein docking interfaces and membrane‐spanning regions as for modeling monomers. Proteins 2014; 82:1971–1984. © 2014 Wiley Periodicals, Inc.  相似文献   

7.
Human heat shock protein of 90 kDa (hHsp90) is a homodimer that has an essential role in facilitating malignant transformation at the molecular level. Inhibiting hHsp90 function is a validated approach for treating different types of tumors. Inhibiting the dimerization of hHsp90 via its C-terminal domain (CTD) should provide a novel way to therapeutically interfere with hHsp90 function. Here, we predicted hot spot residues that cluster in the CTD dimerization interface by a structural decomposition of the effective energy of binding computed by the MM-GBSA approach and confirmed these predictions using in silico alanine scanning with DrugScorePPI. Mutation of these residues to alanine caused a significant decrease in the melting temperature according to differential scanning fluorimetry experiments, indicating a reduced stability of the mutant hHsp90 complexes. Size exclusion chromatography and multi-angle light scattering studies demonstrate that the reduced stability of the mutant hHsp90 correlates with a lower complex stoichiometry due to the disruption of the dimerization interface. These results suggest that the identified hot spot residues can be used as a pharmacophoric template for identifying and designing small-molecule inhibitors of hHsp90 dimerization.  相似文献   

8.
Size exclusion chromatographic analyses showed that Ca2+‐free VILIP‐1 contained both monomeric and dimeric forms, while no appreciable dimerization was noted with Ca2+‐free VILIP‐3. Swapping of EF‐hands 3 and 4 of VILIP‐1 with those of VILIP‐3 caused the inability of the resulting chimeric protein to form dimeric protein. Nonreducing SDS‐PAGE analyses revealed that most of the dimeric VILIP‐1 was noncovalently bound together. Reduced glutathione (GSH)/oxidized glutathione (GSSG) treatment notably enhanced the formation of disulfide‐linked VILIP‐1 dimer, while Ca2+ and Mg2+ enhanced disulfide dimerization of VILIP‐1 marginally in the presence of thiol compounds. Cys‐187 at the C‐terminus of VILIP‐1 contributed greatly to form S‐S‐crosslinked dimer as revealed by mutagenesis studies. The ability of GSH/GSSG‐treated VILIP‐1 to activate guanylyl cyclase B was reduced by substituting Cys‐187 with Ala. Together with disulfide dimer of VILIP‐1 detected in rat brain extracts, our data may imply the functional contribution of disulfide dimer to the interaction of VILIP‐1 with its physiological target(s). Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
A plethora of both experimental and computational methods have been proposed in the past 20 years for the identification of hot spots at a protein–protein interface. The experimental determination of a protein–protein complex followed by alanine scanning mutagenesis, though able to determine hot spots with much precision, is expensive and has no guarantee of success while the accuracy of the current computational methods for hot‐spot identification remains low. Here, we present a novel structure‐based computational approach that accurately determines hot spots through docking into a set of proteins homologous to only one of the two interacting partners of a compound capable of disrupting the protein–protein interaction (PPI). This approach has been applied to identify the hot spots of human activin receptor type II (ActRII) critical for its binding toward Cripto‐I. The subsequent experimental confirmation of the computationally identified hot spots portends a potentially accurate method for hot‐spot determination in silico given a compound capable of disrupting the PPI in question. The hot spots of human ActRII first reported here may well become the focal points for the design of small molecule drugs that target the PPI. The determination of their interface may have significant biological implications in that it suggests that Cripto‐I plays an important role in both activin and nodal signal pathways.  相似文献   

10.
Aldehyde dehydrogenases (ALDHs) couple the oxidation of aldehydes to the reduction of NAD(P)+. These enzymes have gained importance as they have been related to the detoxification of aldehydes generated in several diseases involving oxidative stress. It has been determined that tetrameric ALDHs work only with two of their four active sites (half‐of‐the‐sites reactivity), but the mechanistic reason for this feature remains unknown. In this study, tetrameric human aldehyde dehydrogenase class 1A1 (ALDH1A1) was dimerized to study the correlation of the oligomeric structure with the presence of half‐of‐the‐sites reactivity. Stable dimers from ALDH1A1 were generated by combining the mutation of two residues of the dimer–dimer interface in the tetramer (previously shown to render a low‐active and unstable enzyme) and the fusion of green fluorescent protein (GFP) in the C‐terminus of the mutant. Some kinetic properties of the GFP‐fusion mutant resembled those of human aldehyde dehydrogenase class 3A1, a native dimer, in that the fusion dimer did not show burst in the generation of nicotinamide adenine dinucleotide (NADH) and was less sensitive to the action of specific modulators. The presence of primary isotope effect indicated that the rate‐limiting step changed from NADH release to hydride transfer. The mutant showed higher activity with malondialdehyde and acrolein and was more resistant to inactivation by acrolein compared with the wild type. The mutant kinetic profile showed two hyperbolic components when the substrates were varied, suggesting the presence of two active sites with different affinities and catalytic capacities. In conclusion, the ALDH1A1–GFP dimeric mutant exhibits full site reactivity, suggesting that only the tetrameric structure induces the half‐of‐the‐sites reactivity. Proteins 2013; 81:1330–1339. © 2013 Wiley Periodicals, Inc.  相似文献   

11.
Zhenhua Li  Jinyan Li 《Proteins》2010,78(16):3304-3316
A protein interface can be as “wet” as a protein surface in terms of the number of immobilized water molecules. This important water information has not been explicitly taken by computational methods to model and identify protein binding hot spots, overlooking the water role in forming interface hydrogen bonds and in filing cavities. Hot spot residues are usually clustered at the core of the protein binding interfaces. However, traditional machine learning methods often identify the hot spot residues individually, breaking the cooperativity of the energetic contribution. Our idea in this work is to explore the role of immobilized water and meanwhile to capture two essential properties of hot spots: the compactness in contact and the far distance from bulk solvent. Our model is named geometrically centered region (GCR). The detection of GCRs is based on novel tripartite graphs, and atom burial levels which are a concept more intuitive than SASA. Applying to a data set containing 355 mutations, we achieved an F measure of 0.6414 when ΔΔG ≥ 1.0 kcal/mol was used to define hot spots. This performance is better than Robetta, a benchmark method in the field. We found that all but only one of the GCRs contain water to a certain degree, and most of the outstanding hot spot residues have water‐mediated contacts. If the water is excluded, the burial level values are poorly related to the ΔΔG, and the model loses its performance remarkably. We also presented a definition for the O‐ring of a GCR as the set of immediate neighbors of the residues in the GCR. Comparative analysis between the O‐rings and GCRs reveals that the newly defined O‐ring is indeed energetically less important than the GCR hot spot, confirming a long‐standing hypothesis. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

12.
The equilibrium properties of a HIV‐1‐protease precursor are studied by means of an efficient molecular dynamics scheme, which allows for the simulation of the folding of the protein monomers and their dimerization into an active form and compare them with those of the mature protein. The results of the model provide, with atomic detail, an overall account of several experimental findings, including the NMR conformation of the mature dimer, the calorimetric properties of the system, the effects of the precursor tail on the dimerization constant, the secondary chemical shifts of the monomer, and the paramagnetic relaxation enhancement data associated with the conformations of the precursor. It is found that although the mature protein can dimerize in a unique, single way, the precursor populates several dimeric conformations in which monomers are always native‐like, but their binding can be non‐native. Proteins 2014; 82:633–639. © 2013 Wiley Periodicals, Inc.  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.
Zhong N  Zhang S  Zou P  Chen J  Kang X  Li Z  Liang C  Jin C  Xia B 《Journal of virology》2008,82(9):4227-4234
The main protease (M(pro)) of severe acute respiratory syndrome coronavirus (SARS-CoV) plays an essential role in the extensive proteolytic processing of the viral polyproteins (pp1a and pp1ab), and it is an important target for anti-SARS drug development. It was found that SARS-CoV M(pro) exists in solution as an equilibrium of both monomeric and dimeric forms, and the dimeric form is the enzymatically active form. However, the mechanism of SARS-CoV M(pro) dimerization, especially the roles of its N-terminal seven residues (N-finger) and its unique C-terminal domain in the dimerization, remain unclear. Here we report that the SARS-CoV M(pro) C-terminal domain alone (residues 187 to 306; M(pro)-C) is produced in Escherichia coli in both monomeric and dimeric forms, and no exchange could be observed between them at room temperature. The M(pro)-C dimer has a novel dimerization interface. Meanwhile, the N-finger deletion mutant of SARS-CoV M(pro) also exists as both a stable monomer and a stable dimer, and the dimer is formed through the same C-terminal-domain interaction as that in the M(pro)-C dimer. However, no C-terminal domain-mediated dimerization form can be detected for wild-type SARS-CoV M(pro). Our study results help to clarify previously published controversial claims about the role of the N-finger in SARS-CoV M(pro) dimerization. Apparently, without the N-finger, SARS-CoV M(pro) can no longer retain the active dimer structure; instead, it can form a new type of dimer which is inactive. Therefore, the N-finger of SARS-CoV M(pro) is not only critical for its dimerization but also essential for the enzyme to form the enzymatically active dimer.  相似文献   

15.
Diaminopimelate (DAP) epimerase is involved in the biosynthesis of meso-DAP and lysine, which are important precursors for the synthesis of peptidoglycan, housekeeping proteins, and virulence factors in bacteria. Accordingly, DAP epimerase is a promising antimicrobial target. Previous studies report that DAP epimerase exists as a monomeric enzyme. However, we show using analytical ultracentrifugation, X-ray crystallography, and enzyme kinetic analyses that DAP epimerase from Escherichia coli exists as a functional dimer in solution and the crystal state. Furthermore, the 2.0-Å X-ray crystal structure of the E. coli DAP epimerase dimer shows for the first time that the enzyme exists in an open, active conformation. The importance of dimerization was subsequently probed by using site-directed mutagenesis to generate a monomeric mutant (Y268A). Our studies show that Y268A is catalytically inactive, thus demonstrating that dimerization of DAP epimerase is essential for catalysis. Molecular dynamics simulations indicate that the DAP epimerase monomer is inherently more flexible than the dimer, suggesting that dimerization optimizes protein dynamics to support function. Our findings offer insight into the development of novel antimicrobial agents targeting the dimeric antibiotic target DAP epimerase.  相似文献   

16.
The identification of protein–protein interactions is vital for understanding protein function, elucidating interaction mechanisms, and for practical applications in drug discovery. With the exponentially growing protein sequence data, fully automated computational methods that predict interactions between proteins are becoming essential components of system‐level function inference. A thorough analysis of protein complex structures demonstrated that binding site locations as well as the interfacial geometry are highly conserved across evolutionarily related proteins. Because the conformational space of protein–protein interactions is highly covered by experimental structures, sensitive protein threading techniques can be used to identify suitable templates for the accurate prediction of interfacial residues. Toward this goal, we developed eFindSitePPI, an algorithm that uses the three‐dimensional structure of a target protein, evolutionarily remotely related templates and machine learning techniques to predict binding residues. Using crystal structures, the average sensitivity (specificity) of eFindSitePPI in interfacial residue prediction is 0.46 (0.92). For weakly homologous protein models, these values only slightly decrease to 0.40–0.43 (0.91–0.92) demonstrating that eFindSitePPI performs well not only using experimental data but also tolerates structural imperfections in computer‐generated structures. In addition, eFindSitePPI detects specific molecular interactions at the interface; for instance, it correctly predicts approximately one half of hydrogen bonds and aromatic interactions, as well as one third of salt bridges and hydrophobic contacts. Comparative benchmarks against several dimer datasets show that eFindSitePPI outperforms other methods for protein‐binding residue prediction. It also features a carefully tuned confidence estimation system, which is particularly useful in large‐scale applications using raw genomic data. eFindSitePPI is freely available to the academic community at http://www.brylinski.org/efindsiteppi . Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

17.
The tRNA-modifying enzyme tRNA-guanine transglycosylase (Tgt) is a putative target for new selective antibiotics against Shigella bacteria. The formation of a Tgt homodimer was suggested on the basis of several crystal structures of Tgt in complex with RNA. In the present study, noncovalent mass spectrometry was used (i) to confirm the dimeric oligomerization state of Tgt in solution and (ii) to evidence the binding stoichiometry of the complex formed between Tgt and its full-length substrate tRNA. To further investigate the importance of Tgt protein-protein interaction, point mutations were introduced into the dimer interface in order to study their influence on the formation of the catalytically active complex. Enzyme kinetics revealed a reduced catalytic activity of these mutated variants, which could be related to a destabilization of the dimer formation as evidenced by both noncovalent mass spectrometry and X-ray crystallography. Finally, the effect of inhibitor binding was investigated by noncovalent mass spectrometry, thus providing the binding stoichiometries of Tgt:inhibitor complexes and showing competitive interactions in the presence of tRNA. Inhibitors that display an influence on the formation of the dimer interface in the crystal structure are promising candidates to alter the protein-protein interaction, which could provide a new way to inhibit Tgt.  相似文献   

18.
The viral immediate-early protein 1 (IE1) is crucial for efficient replication of cytomegalovirus (CMV). A recent crystal structure of the IE1 protein from rhesus CMV revealed that the protein exhibits a novel fold and crystallizes in two slightly different dimeric arrangements. Molecular dynamics simulations and energetic analyses performed in this study show that both dimers are stable and allowed us to identify a common set of five residues that appear particularly important for dimer formation. These residues are distributed over the entire dimer interface and do not form a typical hot spot for protein interactions. In addition, the dimer interface of IE1 proved to include a high portion of hydrophilic interactions pointing toward the transient nature of dimer formation. Characterization of monomeric and dimeric IE1 revealed three sequentially discontinuous dynamic domains that exhibit correlated motion within the domain and are simultaneously anti-correlated to the adjacent domains. The hinge motions observed between the dynamic domains increase the shape complementarity to the coiled–coil region of tripartite motif proteins, suggesting that the detected dynamics of IE1 might be physiologically important by enabling a better interaction with its cellular target molecules.  相似文献   

19.
Rhamnogalacturonan‐II (RG‐II) is structurally the most complex glycan in higher plants, containing 13 different sugars and 21 distinct glycosidic linkages. Two monomeric RG‐II molecules can form an RG‐II‐borate diester dimer through the two apiosyl (Api) residues of side chain A to regulate cross‐linking of pectin in the cell wall. But the relationship of Api biosynthesis and RG‐II dimer is still unclear. In this study we investigated the two homologous UDP‐D‐apiose/UDP‐D‐xylose synthases (AXSs) in Arabidopsis thaliana that synthesize UDP‐D‐apiose (UDP‐Api). Both AXSs are ubiquitously expressed, while AXS2 has higher overall expression than AXS1 in the tissues analyzed. The homozygous axs double mutant is lethal, while heterozygous axs1/+ axs2 and axs1 axs2/+ mutants display intermediate phenotypes. The axs1/+ axs2 mutant plants are unable to set seed and die. By contrast, the axs1 axs2/+ mutant plants exhibit loss of shoot and root apical dominance. UDP‐Api content in axs1 axs2/+ mutants is decreased by 83%. The cell wall of axs1 axs2/+ mutant plants is thicker and contains less RG‐II‐borate complex than wild‐type Col‐0 plants. Taken together, these results provide direct evidence of the importance of AXSs for UDP‐Api and RG‐II‐borate complex formation in plant growth and development.  相似文献   

20.
Human cystatin C (HCC) is one of the amyloidogenic proteins to be shown to oligomerize via a three‐dimensional domain swapping mechanism. This process precedes the formation of a stable dimer and proceeds particularly easily in the case of the L68Q mutant. According to the proposed mechanism, dimerization of the HCC precedes conformational changes within the β2 and β3 strands. In this article, we present conformational studies, using circular dichroism and MD methods, of the β2‐L1‐β3 (His43‐Thr72) fragment of the HCC involved in HCC dimer formation. We also carried out studies of the β2‐L1‐β3 peptide, in which the Val57 residue was replaced by residues promoting β‐turn structure formation (Asp, Asn, or Pro). The present study established that point mutation could modify the structure of the L1 loop in the β‐hairpin peptide. Our results showed that the L1 loop in the peptide excised from human cystatin C is broader than that in cystatin C. In the HCC protein, broadening of the L1 loop together with the unfavorable L68Q mutation in the hydrophobic pocket could be a force sufficient to cause the partial unfolding and then the opening of HCC or its L68Q mutant structure for further dimerization. We presume further that the Asp57 and Asn57 mutations in the L1 loop of HCC could stabilize the closed form of HCC, whereas the Pro57 mutation could lead to the opening of the HCC structure and then to dimer/oligomer formation. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 373–383, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  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