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.
Amelogenin is believed to be involved in controlling the formation of the highly anisotropic and ordered hydroxyapatite crystallites that form enamel. The adsorption behavior of amelogenin proteins onto substrates is very important because protein–surface interactions are critical to its function. We have previously used LRAP, a splice variant of amelogenin, as a model protein for the full-length amelogenin in solid-state NMR and neutron reflectivity studies at interfaces. In this work, we examined the adsorption behavior of LRAP in greater detail using model self-assembled monolayers containing COOH, CH3, and NH2 end groups as substrates. Dynamic light scattering (DLS) experiments indicated that LRAP in phosphate buffered saline and solutions containing low concentrations of calcium and phosphate consisted of aggregates of nanospheres. Null ellipsometry and atomic force microscopy (AFM) were used to study protein adsorption amounts and quaternary structures on the surfaces. Relatively high amounts of adsorption occurred onto the CH3 and NH2 surfaces from both buffer solutions. Adsorption was also promoted onto COOH surfaces only when calcium was present in the solutions suggesting an interaction that involves calcium bridging with the negatively charged C-terminus. The ellipsometry and AFM studies revealed that LRAP adsorbed onto the surfaces as small subnanosphere-sized structures such as monomers or dimers. We propose that the monomers/dimers were present in solution even though they were not detected by DLS or that they adsorbed onto the surfaces by disassembling or “shedding” from the nanospheres that are present in solution. This work reveals the importance of small subnanosphere-sized structures of LRAP at interfaces.  相似文献   

2.
Atomic force microscopy has been used to follow in real time the adsorption from solution of two of the gliadin group of wheat seed storage proteins onto hydrophilic (mica) and hydrophobic (graphite) surfaces. The liquid cell of the microscope was used initially to acquire images of the substrate under a small quantity of pure solvent (1% acetic acid). Continuous imaging as an injection of gliadin solution entered the liquid cell enabled the adsorption process to be followed in situ from zero time. For ω‐gliadin, a monolayer was formed on the mica substrate during a period of ~2000 s, with the protein molecules oriented in parallel to the mica surface. In contrast, the ω‐gliadin had a relatively low affinity for the graphite substrate, as demonstrated by slow and weak adsorption to the surface. With γ‐gliadin, random deposition onto the mica surface was observed forming monodispersed structures, whereas on the graphite surface, monolayer islands of protein were formed with the protein molecules in a perpendicular orientation. Sequential adsorption experiments indicated strong interactions between the two proteins that, under certain circumstances, caused alterations to the surface morphologies of preadsorbed species. The results are relevant to our understanding of the interactions of proteins within the hydrated protein bodies of wheat grain and how these determine the processing properties of wheat gluten and dough. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 74–84, 2010. 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   

3.
With the decline in productivity of drug‐development efforts, novel approaches to rational drug design are being introduced and developed. Naturally occurring and synthetic peptides are emerging as novel promising compounds that can specifically and efficiently modulate signaling pathways in vitro and in vivo. We describe sequence‐based approaches that use peptides to mimic proteins in order to inhibit the interaction of the mimicked protein with its partners. We then discuss a structure‐based approach, in which protein‐peptide complex structures are used to rationally design and optimize peptidic inhibitors. We survey flexible peptide docking techniques and discuss current challenges and future directions in the rational design of peptidic inhibitors. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 505–513, 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   

4.
The existence of an “RNA world” as an early step in the history of life increases the interest for the characterization of these biomolecules. The hairpin ribozyme studied here is a self‐cleaving/ligating motif found in the minus strand of the satellite RNA associated with Tobacco ringspot virus. Surface‐enhanced Raman spectroscopy (SERS) is a powerful tool to study trace amounts of RNA. In controlled conditions, a SERS signal is proportional to the amount of free residues adsorbed on the metal surface. On RNA cleavage, residues are unpaired and free to interact with metal. SERS procedures are used to monitor and quantify the catalysis of ribozyme cleavage at biological concentrations in real time; thus, they propose an interesting alternative to electrophoretic methods. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 384–390, 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   

5.
A hemoprotein‐based supramolecular polymer that has a covalently linked heme moiety on the protein surface has been constructed based on interprotein heme–heme pocket interactions of the chemically modified apocytochrome b562 ( 1 ‐H63C). The thermodynamic properties of the polymer have been investigated by means of size exclusion chromatography, UV–vis spectroscopy, and circular dichroism spectroscopy. The results indicate that, as with other synthetic systems reported so far, the 1 ‐H63C hemoprotein assembly is thermodynamically controlled in aqueous solution: the degree of polymerization is dependent on the 1 ‐H63C concentration and is modulated by the addition of the end‐capping units, native heme, and/or apocytochrome b562 mutant (apoH63C). These properties suggest a potential use for the hemoprotein self‐assembly in preparation of stimuli‐responsive functional nanobiomaterials. © 2008 Wiley Periodicals, Inc. Biopolymers 91: 194–200, 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   

6.
We have applied optical waveguide lightmode spectroscopy combined with streaming potential measurements and Fourier-transformed infrared spectroscopy to investigate adsorption of amelogenin nanospheres onto polyelectrolytes. The long-term objective was to better understand the chemical nature of these assemblies and to gain further insight into the molecular mechanisms involved during self-assembly. It was found that monolayers of monomers and negatively charged nanospheres of a recombinant amelogenin (rM179) irreversibly adsorbed onto a positively charged polyelectrolyte multilayer films. On the basis of measurements performed at different temperatures, it was demonstrated that intermolecular interactions for the formation of nanospheres were not affected by their adsorption onto polyelectrolytes. Consecutive adsorption of nanospheres resulting in the formation of multilayer structures was possible by using cationic poly(l-lysine) as mediators. N-Acetyl-d-glucosamine (GlcNac) did not disturb the nanosphere-assembled protein's structure, and it only affected the adsorption of monomeric amelogenin. Infrared spectroscopy of adsorbed amelogenin revealed conformational differences between the monomeric and assembled forms of rM179. While there was a considerable amount of alpha-helices in the monomers, beta-turn and beta-sheet structures dominated the assembled proteins. Our work constitutes the first report on a structurally controlled in vitro buildup of an rM179 nanosphere monolayer-based matrix. Our data support the notion that amelogenin self-assembly is mostly driven by hydrophobic interactions and that amelogenin/PEM interactions are dominated by electrostatic forces. We suggest that similar forces can govern amelogenin interactions with non-amelogenins or the mineral phase during enamel biomineralization.  相似文献   

7.
The misfolding and aggregation of disease proteins is characteristic of numerous neurodegenerative diseases. Particular neuronal populations are more vulnerable to proteotoxicity while others are more apt to tolerate the misfolding and aggregation of disease proteins. Thus, the cellular environment must play a significant role in determining whether disease proteins are converted into toxic or benign forms. The endomembrane network of eukaryotes divides the cell into different subcellular compartments that possess distinct sets of molecular chaperones and protein interaction networks. Chaperones act as agonists and antagonists of disease protein aggregation to prevent the accumulation of toxic intermediates in the aggregation pathway. Interacting partners can also modulate the conformation and localization of disease proteins and thereby influence proteotoxicity. Thus, interplay between these protein homeostasis network components can modulate the self‐association of disease proteins and determine whether they elicit a toxic or benign outcome. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 229–236, 2010. 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   

8.
In vivo cell division protein FtsZ from E. coli forms rings and spirals which have only been observed by low resolution light microscopy. We show that these suprastructures are likely formed by molecular crowding which is a predominant factor in prokaryotic cells and enhances the weak lateral bonds between proto‐filaments. Although FtsZ assembles into single proto‐filaments in dilute aqueous buffer, with crowding agents above a critical concentration, it forms polymorphic supramolecular structures including rings and toroids (with multiple protofilaments) about 200 nm in diameter, similar in appearance to DNA toroids, and helices with pitches of several hundred nm as well as long, linear bundles. Helices resemble those observed in vivo, whereas the rings and toroids may represent a novel energy minimized state of FtsZ, at a later stage of Z‐ring constriction. We shed light on the molecular arrangement of FtsZ filaments within these suprastructures using high resolution electron microscopy. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 340–350, 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   

9.
Is linker DNA bent in the 30‐nm chromatin fiber at physiological conditions? We show here that electrostatic interactions between linker DNA and histone tails including salt condensation and release may bend linker DNA, thus affecting the higher order organization of chromatin. © 2005 Wiley Periodicals, Inc. Biopolymers 81: 20–28, 2006 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   

10.
Recently, ubiquitin was suggested as a promising anti‐inflammatory protein therapeutic. We found that a peptide fragment corresponding to the ubiquitin50–59 sequence (LEDGRTLSDY) possessed the immunosuppressive activity comparable with that of ubiquitin. CD and NMR spectroscopies were used to determine the conformational preferences of LEDGRTLSDY in solution. The peptide mixture, obtained by pepsin digestion of ubiquitin, was even more potent than the intact protein. Although the peptide exhibited a well‐defined conformation in methanol, its structure was distinct from the corresponding 50–59 fragment in the native ubiquitin molecule. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 423–431, 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   

11.
Dragline spider silk has been intensively studied for its superior qualities as a biomaterial. In previous studies, we made use of the baculovirus mediated expression system for the production of a recombinant Araneus diadematus spider silk dragline ADF4 protein and its self‐assembly into intricate fibers in host insect cells. In this study, our aim was to explore the function of the major repetitive domain of the dragline spider silk. Thus, we generated an array of synthetic proteins, each containing a different number of identical repeats up to the largest recombinantly expressed spider silk to date. Study of the self‐assembly properties of these proteins showed that depending on the increasing number of repeats they give rise to different assembly phenotypes, from a fully soluble protein to bona fide fibers with superior qualities. The different assembly forms, the corresponding chemical resistance properties obtained as well as ultrastructural studies, revealed novel insights concerning the structure and intermolecular interactions of the repetitive and nonrepetitive domains. Based on these observations and current knowledge in the field, we hereby present a comprehensive hypothetical model for the mechanism of dragline silk self‐assembly and fiber formation. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 458–468, 2010. 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   

12.
The secondary structures of proteins (alpha-helical, beta-sheet, beta-turn, and random coil) in the solid state and when bound to polymer beads, containing immobilized phenyl and butyl ligands such as those as commonly employed in hydrophobic interaction chromatography, have been investigated using FTIR-ATR spectroscopy and partial least squares (PLS) methods. Proteins with known structural features were used as models, including 12 proteins in the solid state and 7 proteins adsorbed onto the hydrophobic surfaces. A strong PLS correlation was achieved between predictions derived from the experimental data for 4 proteins adsorbed onto the phenyl-modified beads and reference data obtained from the X-ray crystallographic structures with r(2) values of 0.9974, 0.9864, 0.9924, and 0.9743 for alpha-helical, beta-sheet, beta-turn, and random coiled structures, respectively. On the other hand, proteins adsorbed onto the butyl sorbent underwent greater secondary structural changes compared to the phenyl sorbent as evidenced from the poorer PLS r(2) values (r(2) are 0.9658, 0.9106, 0.9571, and 0.9340). The results thus indicate that the secondary structures for these proteins were more affected by the butyl sorbent, whereas the secondary structure remains relatively unchanged for the proteins adsorbed onto the phenyl sorbent. This study has important ramifications for understanding the nature of protein secondary structural changes following adsorption onto hydrophobic sorbent surfaces. This knowledge could also enable the development of useful protocols for enhancing the chromatographic purification of proteins in their native bioactive states. (c) 2008 Wiley Periodicals, Inc. Biopolymers 89: 895-905, 2008.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   

13.
Despite detailed knowledge of the overall structural changes and stoichiometries of surfactant binding, little is known about which protein regions constitute the preferred sites of attack for initial unfolding. Here we have exposed three proteins to limited proteolysis at anionic (SDS) and cationic (DTAC) surfactant concentrations corresponding to specific conformational transitions, using the surfactant‐robust broad‐specificity proteases Savinase and Alcalase. Cleavage sites are identified by SDS‐PAGE and N‐terminal sequencing. We observe well‐defined cleavage fragments, which suggest that flexibility is limited to certain regions of the protein. Cleavage sites for α‐lactalbumin and myoglobin correspond to regions identified in other studies as partially unfolded at low pH or in the presence of organic solvents. For Tnfn3, which does not form partially folded structures under other conditions, cleavage sites can be rationalized from the structure of the protein's folding transition state and the position of loops in the native state. Nevertheless, they are more sensitive to choice of surfactant and protease, probably reflecting a heterogeneous and fluctuating ensemble of partially unfolded structures. Thus, for proteins accumulating stable intermediates on the folding pathway, surfactants encourage the formation of these states, while the situation is more complex for proteins that do not form these intermediates. © 2008 Wiley Periodicals, Inc. Biopolymers 91: 221–231, 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   

14.
Because over expression of Hsp70 molecular chaperones suppresses the toxicity of aberrantly folded proteins that occur in Alzheimer's disease (AD), Parkinson's disease (PD), amyotrophic lateral sclerosis, and various polyQ‐diseases (Huntington's disease and ataxias), Hsp70 is garnering attention as a possible therapeutic agent for these various diseases. Here, I review progress in this fascinating field of molecular chaperones and neurodegeneration and describe our current understanding of the mechanisms by which Hsp70 protects cells from the PD‐related protein called alpha‐synuclein (α‐syn). © 2009 Wiley Periodicals, Inc. Biopolymers 93: 218–228, 2010. 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   

15.
The vibrational structures of Nociceptin (FQ), its short bioactive fragments, and specifically‐modified [Tyr1]FQ (1‐6), [His1]FQ (1‐6), and [His1,4]FQ (1‐6) fragments were characterized. We showed that in the solid state, all of the aforementioned peptides except FQ adopt mainly turn and disordered secondary structures with a small contribution from an antiparallel β‐sheet conformation. FQ (1‐11), FQ (7‐17) [His1]FQ (1‐6), and [His1,4]FQ (1‐6) have an α‐helical backbone arrangement that could also slightly influence their secondary structure. The adsorption behavior of these peptides on a colloidal silver surface in an aqueous solution (pH = ~8.3) was investigated by means of surface‐enhanced Raman scattering (SERS). All of the peptides, excluding FQ (7‐17), chemisorbed on the colloidal silver surfaces through a Phe4 residue, which for FQ, FQ (1‐11), FQ (1‐6), [Tyr1]FQ (1‐6), and [His1]FQ (1‐6) lies almost flat on this surface, while for FQ (1‐13) and FQ (1‐13)NH2 adopts a slightly tilted orientation with respect to the surface. The Tyr1 residue in [Tyr1]FQ (1‐6) does not interact with the colloidal silver surface, suggesting that the Tyr1 and Phe4 side chains are located on the opposite sides of the peptide backbone, which can be also true for His1 and Phe4 in [His1]FQ (1‐6). The lone pair of electrons on the oxygen atom of the ionized carbonyl group of FQ (1‐13) and FQ (7‐17) appears to be coordinated to the colloidal silver nanoparticles, whereas in the case of the remaining peptides, it only assists in the adsorption process, similar to the ? NH2 group. We also showed that upon adsorption, the secondary structure of these peptides is altered. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1039–1054, 2010. 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   

16.
Poly‐L ‐proline has been used as a model system for various purposes over a period of more than 60 years. Its relevance among the protein/peptide community stems from its use as a reference system for determining the conformational distributions of unfolded peptides and proteins, its use as a molecular ruler, and from the pivotal role of proline residues in conformational transitions and protein–protein interactions. While several studies indicate that polyproline can aggregate and precipitate in aqueous solution, a systematic study of the aggregation process is still outstanding. We found, by means of UV‐circular dichroism and IR measurements, that polyproline is predominately monomeric at room temperature at millimolar concentrations. Upon heating, the polypeptide stays in its monomeric state until the temperature reaches a threshold of ca. 60°C. At higher temperatures, the peptide aggregates as a film on the inside surface of the employed cuvette. The process proceeds on a time scale of 103 s and can best be described by a bi‐exponential relaxation function. The respective CD and IR spectra are qualitatively different from the canonical spectra of polyproline in aqueous solution, and are indicative of a highly packed state. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 451–457, 2010. 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   

17.
Jun Gao  Zhijun Li 《Biopolymers》2010,93(4):340-347
It is widely accepted that a protein's sequence determines its structure. The surprising finding that proteins of distant sequence can adopt similar 3D structures has raised interesting questions regarding underlying conserved properties that are essential for protein folding and stability. Uncovering the conserved properties may shed light on the folding mechanism of proteins and help with the development of computational tools for protein structure prediction. We compiled and analyzed a structure pair dataset of 66 high‐resolution and low sequence identity (16–38%) soluble proteins. Structure deviation for each pair was confirmed by calculating its Cα SiMax value and comparing its potential energy per residue. Analysis of favorable inter‐residue interactions for each structure pair indicated that the average number of inter‐residue interactions within each structure represents a conserved feature of homologous structures of distant sequence. Detailed comparison of individual types of interactions showed that the average number of either hydrophobic or hydrogen bonding interactions remains unchanged for each structure pair. These findings should be of help to improving the quality of homology models based on templates of low sequence identity, thus broadening the application of homology modeling techniques for protein studies. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 340–347, 2010. 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   

18.
Protein engineering to functionalize the self‐assembling enamel matrix protein amelogenin with a cellulose binding domain (CBD) is used. The purpose is to examine the binding of the engineered protein, rh174CBD, to cellulose materials, and the possibility to immobilize self‐assembled amelogenin nanospheres on cellulose. rh174CBD assembled to nanospheres ≈35 nm in hydrodynamic diameter, very similar in size to wild type amelogenin (rh174). Uniform particles are formed at pH 10 for both rh174 and rh174CBD, but only rh174CBD nanospheres showes significant binding to cellulose (Avicel). Cellulose binding of rh174CBD is promoted when the protein is self‐assembled to nanospheres, compared to being in a monomeric form, suggesting a synergistic effect of the multiple CBDs on the nanospheres. The amount of bound rh174CBD nanospheres reached ≈15 mg/g Avicel, which corresponds to 4.2 to 6.3 × 10?7 mole/m2. By mixing rh174 and rh174CBD, and then inducing self‐assembly, composite nanospheres with a high degree of cellulose binding can be formed, despite a lower proportion of rh174CBD. This demonstrates that amelogenin variants like rh174 can be incorporated into the nanospheres, and still retain most of the binding to cellulose. Engineered amelogenin nanoparticles can thus be utilized to construct a range of new cellulose based hybrid materials, e.g. for wound treatment.  相似文献   

19.
The insulin‐linked polymorphic region (ILPR) is a VNTR region located upstream of the insulin (INS) gene consisting of the repeat 5′‐ACAGGGGTGTGGGG (repeat a) and several less abundant sequence repeats (b–n). Here, we have investigated the structural polymorphism of G‐quadruplexes formed from the most common repeat sequences (a–c) and their effect on insulin protein binding. We first established that the ILPR repeats “b” and “c” can form quadruplex structures. Insulin has previously been shown to bind a G‐quadruplex formed by a dimer of the repeat “a”. Our findings show that insulin binds preferentially to the repeat “a” G‐quadruplex (Kd = 0.17 ± 0.03 μM) over G‐quadruplexes formed from other ILPR repeats that were tested (Kds from 0.71 ± 0.15 to 1.07 ± 0.09 μM). Additionally, the Watson‐Crick complementary relationship between the loop regions of repeat “a” (ACA and TGT) seemingly play an important role in favoring a specific G‐quadruplex conformation, which based on our data is critical for insulin binding. Affinity for insulin is reduced in sequences lacking the putative WC complementarity, however upon engineered restoration of complementarity, insulin binding is recovered. A DMS footprinting assay on the repeat “a” G‐quadruplex in the presence of insulin, combined with binding affinities for ILPR mutants led to identification of a loop nucleotide critical for binding. Uniquely, insulin shows clear preference for binding to the G‐quadruplexes with the more antiparallel feature. Collectively, our results illustrate the specific nature of insulin binding to the ILPR G‐quadruplexes and begin to provide molecular details on such interactions. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 21–31, 2010. 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   

20.
Luteolin (LUT) is a polyphenolic compound, found in a variety of fruits, vegetables, and seeds, which has a variety of pharmacological properties. In the present contribution, binding of LUT to human serum albumin (HSA), the most abundant carrier protein in the blood, was investigated with the aim of describing the binding mode and parameters of the interaction. The application of circular dichroism, UV‐Vis absorption, fluorescence, Raman and surface‐enhanced Raman scattering spectroscopy combined with molecular modeling afforded a clear picture of the association mode of LUT to HSA. Specific interactions with protein amino acids were evidenced. LUT was found to be associated in subdomain IIA where an interaction with Trp‐214 is established. Hydrophobic and electrostatic interactions are the major acting forces in the binding of LUT to HSA. The HSA conformations were slightly altered by the drug complexation with reduction of α‐helix and increase of β‐turns structures, suggesting a partial protein unfolding. Also the configuration of at least two disulfide bridges were altered. Furthermore, the study of molecular modeling afforded the binding geometry. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 917–927, 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