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.
Understanding the biophysical mechanism of polyglutamine (polyGln) aggregation is important to unravel the role of aggregates in the pathology of polyGln repeat disorders. To achieve this, synthetic polyGln peptides are widely used. Their disaggregation and solubilization is essential because it plays a crucial role in reproducing biophysical experimental data under in vitro conditions. Pretreatment with trifluoroacetic acid (TFA) and hexafluoroisopropanol (HFIP) at a 1:1 ratio is currently the method of choice to achieve solubility of polyGln peptides. Here we report that the disaggregation and solubilization of polyGln peptides can be achieved by TFA alone. We tested TFA due to the close similarity of it with HFIP in the nature of H-bond breakage and formation, higher cost, and the problems faced by us in the availability of HFIP. Our results demonstrate that the TFA disaggregated polyGln sequences give similar solubilization yield, aggregation kinetics, thioflavin T (ThT) binding, and structural features in comparison with the TFA/HFIP method. Furthermore, we show by limited validation studies that the proposed TFA method can replace the existing TFA/HFIP disaggregation method of polyGln sequences.  相似文献   

2.
Fragments of proteins containing an expanded polyglutamine (polyQ) tract are thought to initiate aggregation and toxicity in at least nine neurodegenerative diseases, including Huntington''s disease. Because proteasomes appear unable to digest the polyQ tract, which can initiate intracellular protein aggregation, preventing polyQ peptide aggregation by chaperones should greatly improve polyQ clearance and prevent aggregate formation. Here we expressed polyQ peptides in cells and show that their intracellular aggregation is prevented by DNAJB6 and DNAJB8, members of the DNAJ (Hsp40) chaperone family. In contrast, HSPA/Hsp70 and DNAJB1, also members of the DNAJ chaperone family, did not prevent peptide-initiated aggregation. Intriguingly, DNAJB6 and DNAJB8 also affected the soluble levels of polyQ peptides, indicating that DNAJB6 and DNAJB8 inhibit polyQ peptide aggregation directly. Together with recent data showing that purified DNAJB6 can suppress fibrillation of polyQ peptides far more efficiently than polyQ expanded protein fragments in vitro, we conclude that the mechanism of DNAJB6 and DNAJB8 is suppression of polyQ protein aggregation by directly binding the polyQ tract.  相似文献   

3.
Abnormally expanded polyglutamine domains in proteins are associated with several neurodegenerative diseases, including Huntington's disease. Expansion of the polyglutamine (polyQ) domain facilitates aggregation of the affected protein, and several studies directly link aggregation to neurotoxicity. Studies of synthetic polyQ peptides have contributed substantially to our understanding of the mechanism of aggregation. In this report, polyQ fibrils were immobilized onto a sensor, and their elongation by polyQ peptides of various length and conformation was examined using quartz crystal microbalance with dissipation monitoring (QCM-D). The rate of elongation increased as the peptide length increased from 8 to 24 glutamines (Q8, Q20, and Q24). Monomer conformation affected elongation rates: insertion of a β-turn template d-Pro-Gly in the center of the peptide increased elongation rates several-fold, while insertion of Pro-Pro dramatically slowed elongation. Dissipation measurements of the QCM-D provided qualitative information about mechanical properties of the elongating fibrils. These data showed clear differences in the characteristics of the elongating aggregates, depending on the specific identity of the associating polyQ peptide. Elongation rates were sensitive to the pH and ionic strength of the buffer. Comparison of QCM-D data with those obtained by optical waveguide lightmode spectroscopy revealed that very little water was associated with the elongation of fibrils by the peptide containing d-Pro-Gly, but a significant amount of water was associated when the fibrils were elongated by Q20. Together, the data indicate that elongation of polyQ fibrils can occur without full consolidation to the fibril structure, resulting in variations to the aggregate structure during elongation.  相似文献   

4.
Polyglutamine (polyQ, a peptide) with an abnormal repeat length is the causative agent of polyQ diseases, such as Huntington’s disease. Although glutamine is a polar residue, polyQ peptides form insoluble aggregates in water, and the mechanism for this aggregation is still unclear. To elucidate the detailed mechanism for the nucleation and aggregation of polyQ peptides, replica exchange molecular dynamics simulations were performed for monomers and dimers of polyQ peptides with several chain lengths. Furthermore, to determine how the aggregation mechanism of polyQ differs from those of other peptides, we compared the results for polyQ with those of polyasparagine and polyleucine. The energy barrier between the monomeric and dimeric states of polyQ was found to be relatively low, and it was observed that polyQ dimers strongly favor the formation of antiparallel β-sheet structures. We also found a characteristic behavior of the monomeric polyQ peptide: a turn at the eighth residue is always present, even when the chain length is varied. We previously showed that a structure including more than two sets of β-turns is stable, so a long monomeric polyQ chain can act as an aggregation nucleus by forming several pairs of antiparallel β-sheet structures within a single chain. Since the aggregation of polyQ peptides has some features in common with an amyloid fibril, our results shed light on the mechanism for the aggregation of polyQ peptides as well as the mechanism for the formation of general amyloid fibrils, which cause the onset of amyloid diseases.  相似文献   

5.
A methodology is presented for the facile synthesis of Arg‐containing peptides modified at the guanidine headgroup as substituted amidine cores. This process allows for the iterative construction of these Arg isosteres while the peptide is being built out on the solid support, providing a high potential for diversity in substitution pattern in the resulting peptide. A series of N‐Pmc‐substituted thioamides were condensed with deprotected δ‐N Orn‐bearing peptides while attached to the solid support using Mukaiyama's reagent as coupling reagent, yielding isosteric Arg‐containing analogs. Peptides were cleaved using trimethylsilyl trifluoromethanesulfonate/TFA and analyzed in their crude form in order to illustrate the amenability of this process toward production of peptide isolates in high crude purity. Arg‐containing peptides having a single Arg isostere were utilized to show the general utility of this approach as well as a multiple‐Arg‐containing construct, illustrating the amenability of this method toward stepwise construction of differently substituted amidine headgroups within the same peptide. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
Expansion of polyglutamine (polyQ) tracts in proteins results in protein aggregation and is associated with cell death in at least nine neurodegenerative diseases. Disease age of onset is correlated with the polyQ insert length above a critical value of 35-40 glutamines. The aggregation kinetics of isolated polyQ peptides in vitro also shows a similar critical-length dependence. While recent experimental work has provided considerable insights into polyQ aggregation, the molecular mechanism of aggregation is not well understood. Here, using computer simulations of isolated polyQ peptides, we show that a mechanism of aggregation is the conformational transition in a single polyQ peptide chain from random coil to a parallel beta-helix. This transition occurs selectively in peptides longer than 37 glutamines. In the beta-helices observed in simulations, all residues adopt beta-strand backbone dihedral angles, and the polypeptide chain coils around a central helical axis with 18.5 +/- 2 residues per turn. We also find that mutant polyQ peptides with proline-glycine inserts show formation of antiparallel beta-hairpins in their ground state, in agreement with experiments. The lower stability of mutant beta-helices explains their lower aggregation rates compared to wild type. Our results provide a molecular mechanism for polyQ-mediated aggregation.  相似文献   

7.
Deposition of insoluble fibrillar aggregates of β‐amyloid (Aβ) peptides in the brain is a hallmark of Alzheimer's disease. Apart from forming fibrils, these peptides also exist as soluble aggregates. Fibrillar and a variety of nonfibrillar aggregates of Aβ have also been obtained in vitro. Hexafluoroisopropanol (HFIP) has been widely used to dissolve Aβ and other amyloidogenic peptides. In this study, we show that the dissolution of Aβ40, 42, and 43 in HFIP followed by drying results in highly ordered aggregates. Although α‐helical conformation is observed, it is not stable for prolonged periods. Drying after prolonged incubation of Aβ40, 42, and 43 peptides in HFIP leads to structural transition from α‐helical to β‐conformation. The peptides form short fibrous aggregates that further assemble giving rise to highly ordered ring‐like structures. Aβ16–22, a highly amyloidogenic peptide stretch from Aβ, also formed very similar rings when dissolved in HFIP and dried. HFIP could not induce α‐helical conformation in Aβ16–22, and rings were obtained from freshly dissolved peptide. The rings formed by Aβ40, 42, 43, and Aβ16–22 are composed of the peptides in β‐conformation and cause enhancement in thioflavin T fluorescence, suggesting that the molecular architecture of these structures is amyloid‐like. Our results clearly indicate that dissolution of Aβ40, 42 and 43 and the amyloidogenic fragment Aβ16–22 in HFIP results in the formation of annular amyloid‐like structures. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
Chatterjee C  Gerig JT 《Biopolymers》2007,87(2-3):115-123
It has been suggested that aggregation of fluorinated alcohols in water solutions is involved with the abilities of these alcohols to provoke conformational changes in peptides and proteins. The extent of fluoroalcohol aggregation depends on the degree of fluorination: hexafluoroisopropanol (HFIP) is more extensively aggregated than is TFE. We previously described a study of the interactions of HFIP with the peptide Trp-cage and provided evidence for the formation of long-lived complexes between this fluoroalcohol and the peptide. In the present work, we have examined the interactions of the less-fluorinated TFE with Trp-cage, in order to probe the role of fluoroalcohol aggregation in the phenomena observed. Intermolecular (1)H{(19)F} nuclear Overhauser effects arising from interactions of TFE with the hydrogens of the peptide in a solution containing 42% TFE were determined at sample temperatures from 5 to 45 degrees C. It is shown that the folded state of the peptide under these conditions is essentially the same as that observed in water and in 30% HFIP-water. The observed peptide-solvent NOEs indicate formation of complexes of Trp-cage with TFE that persist for times of the order of 1 ns. The interactions leading to complexes with TFE are somewhat weaker than those involved in complex formation with HFIP. There are no indications that the aggregation of fluoroalcohol is a necessary concomitant of the interactions of TFE or HFIP with Trp-cage. Rather, the stronger and more long-lived interactions of HFIP with Trp-cage appear to be primarily the result of the greater hydrogen-bonding ability and hydrophobicity of this fluoroalcohol.  相似文献   

9.
In this study, a novel N‐acetyl‐glucosaminylated asparagine derivative was developed. This derivative carried TFA‐sensitive protecting groups and was derived from commercially available compounds only in three steps. It was applicable to the ordinary 9‐fluorenylmethoxycarbonyl (Fmoc)‐based solid‐phase peptide synthesis (SPPS) method, and the protecting groups on the carbohydrate moiety could be removed by a single step of TFA cocktail treatment generally used for the final deprotection step in Fmoc‐SPPS. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Huntington disease (HD), a neurodegenerative disorder, is caused by an expansion of more than 35-40 polyglutamine (polyQ) repeats located near the N-terminus of the huntingtin (htt) protein. The expansion of the polyQ domain results in the ordered assembly of htt fragments into fibrillar aggregates that are the main constituents of inclusion bodies, which are a hallmark of the disease. This paper describes protocols for studying the aggregation of mutant htt fragments and synthetic polyQ peptides with atomic force microscopy (AFM). Ex situ AFM is used to characterize aggregate formation in protein incubation as a function of time. Methods to quickly and unambiguously distinguish specific aggregate species from complex, heterogeneous aggregation reactions based on simple morphological features are presented. Finally, the application of time lapse atomic force microscopy in solution is presented for studying synthetic model polyQ peptides, which allows for tracking the formation and fate of individual aggregates on surfaces over time. This ability allows for dynamic studies of the aggregation process and direct observation of the interplay between different types of aggregates.  相似文献   

11.
We report here the synthesis of the first selenocysteine SPPS derivatives which bear TFA‐labile sidechain protecting groups. New compounds Fmoc‐Sec(Xan)‐OH and Fmoc‐Sec(Trt)‐OH are presented as useful and practical alternatives to the traditional Fmoc‐Sec‐OH derivatives currently available to the peptide chemist. From a bis Fmoc‐protected selenocystine precursor, multiple avenues of diselenide reduction were attempted to determine the most effective method for subsequent attachment of the protecting group electrophiles. Our previously reported one‐pot reduction methodology was ultimately chosen as the optimal approach toward the synthesis of these novel building blocks, and both were easily obtained in high yield and purity. Fmoc‐Sec(Xan)‐OH was discovered to be bench‐stable for extended timeframes while the corresponding Fmoc‐Sec(Trt)‐OH derivative appeared to detritylate slowly when not stored at ?20 °C. Both Sec derivatives were incorporated into single‐ and multiple‐Sec‐containing test peptides in order to ascertain the peptides' deprotection behavior and final form upon TFA cleavage. Single‐Sec‐containing test peptides were always isolated as their corresponding diselenide dimers, while dual‐Sec‐containing peptide sequences were afforded exclusively as their intramolecular diselenides. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Polyglutamine (polyQ) expansion in exon1 (XN1) of the huntingtin protein is linked to Huntington''s disease. When the number of glutamines exceeds a threshold of approximately 36–40 repeats, XN1 can readily form amyloid aggregates similar to those associated with disease. Many experiments suggest that misfolding of monomeric XN1 plays an important role in the length-dependent aggregation. Elucidating the misfolding of a XN1 monomer can help determine the molecular mechanism of XN1 aggregation and potentially help develop strategies to inhibit XN1 aggregation. The flanking sequences surrounding the polyQ region can play a critical role in determining the structural rearrangement and aggregation mechanism of XN1. Few experiments have studied XN1 in its entirety, with all flanking regions. To obtain structural insights into the misfolding of XN1 toward amyloid aggregation, we perform molecular dynamics simulations on monomeric XN1 with full flanking regions, a variant missing the polyproline regions, which are hypothesized to prevent aggregation, and an isolated polyQ peptide (Qn). For each of these three constructs, we study glutamine repeat lengths of 23, 36, 40 and 47. We find that polyQ peptides have a positive correlation between their probability to form a β-rich misfolded state and their expansion length. We also find that the flanking regions of XN1 affect its probability to^x_page_count=28 form a β-rich state compared to the isolated polyQ. Particularly, the polyproline regions form polyproline type II helices and decrease the probability of the polyQ region to form a β-rich state. Additionally, by lengthening polyQ, the first N-terminal 17 residues are more likely to adopt a β-sheet conformation rather than an α-helix conformation. Therefore, our molecular dynamics study provides a structural insight of XN1 misfolding and elucidates the possible role of the flanking sequences in XN1 aggregation.  相似文献   

13.
Huntington's and eight other neurodegenerative diseases occur because of CAG repeat expansion mutation culminating into an expanded polyglutamine tract in respective protein. In Huntington's disease (HD), a number of CAG repeats beyond normal repeat length (>36) lead to the formation of mutant protein, the proteolytic cleavage of which induces aggregation in polyglutamine length‐dependent manner. The neurodegeneration in this disease is linked to aggregation, and its inhibition is a potential approach for therapeutic development. Although peptides and other molecules have been developed for inhibiting aggregation, peptides in general are susceptible to degradation in vivo conditions. To understand their clinical significance, they also need to be delivered through blood–brain barrier. Here, for the first time, we have synthesized poly‐d ,l ‐lactide‐co‐glycolide nanoparticles containing a polyglutamine aggregation inhibitor peptide PGQ9[P2], by nanoprecipitation method. This process yielded less than 200 nm spherical nanoparticles with uniform distribution. Characterization studies by infrared spectroscopy‐based and HPLC‐based assays show the presence of PGQ9[P2] in nanoparticles. In vitro release kinetics demonstrates that nanoparticles release PGQ9[P2] by erosion and diffusion processes. When the PGQ9[P2]‐loaded nanoparticles are incubated with aggregation‐prone Q35P10 peptide, representing N‐terminal part of Huntingtin protein, it arrests the elongation phase of Q35P10 aggregation. These findings propose the first step toward delivery of a peptide inhibitor against polyglutamine aggregation in HD. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
The structure and self‐assembly of the peptide corresponding to the third transmembrane domain (TMD3) of Slc11a1 and its E139A mutant are studied in 1,1,1,3,3,3‐hexafluoro‐2‐propanol (HFIP) aqueous solution by NMR and CD experiments. Slc11a1 is an integral membrane protein with 12 putative TMDs and functions as a pH‐coupled divalent metal cation transporter. Glu139 of Slc11a1 is highly conserved within predicted TMD3 of the Slc11 protein family and function‐associated. Here, we provide the first direct experimental evidence for the structural features of two 24‐residue peptides corresponding to TMD3 of Slc11a1 and its E139A mutant in 60% HFIP‐d2 aqueous solution using CD and NMR spectroscopies. Our study shows that the membrane‐spanning peptide folds as a typical amphipathic α‐helix structure from Ile5 to Met20 with hydrophilic residues Glu12 (Glu139 in Slc11a1) and Asp19 lying on the same side of the helix. The substitution of Glu139 by an alanine residue has little effect on the structure of the peptide, but increases hydrophobicity and facilitates self‐assembly of the peptide. Although the wildtype peptide is monomeric in HFIP aqueous solution, the E139A mutant forms a dimer. The increase in hydrophobicity of the membrane‐spanning peptide and/or change in the interactions between transmembrane segments induced by E139A mutation may affect the metal ion transport of the protein. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
The palette of fluorescent proteins (FPs) available for live‐cell imaging contains proteins that strongly differ in their biophysical properties. FPs cannot be assumed to be equivalent and in certain cases could significantly perturb the behavior of fluorescent reporters. We employed Saccharomyces cerevisiae to comprehensively study the impact of FPs on the toxicity of polyglutamine (polyQ) expansion proteins associated with Huntington's disease. The toxicity of polyQ fusion constructs is highly dependent on the sequences flanking the polyQ repeats. Thus, they represent a powerful tool to study the impact of fluorescent fusion partners. We observed significant differences on polyQ aggregation and toxicity between commonly used FPs. We generated a novel series of vectors with latest yeast‐optimized FPs for investigation of Htt toxicity, including a newly optimized blue FP for expression in yeast. Our study highlights the importance of carefully choosing the optimal FPs when designing tagging strategies.   相似文献   

16.
Huntington disease (HD) is a neurodegenerative disorder caused by expansion of CAG trinucleotide repeats, leading to an elongated polyglutamine sequence (polyQ) in the huntingtin protein. Misfolding of mutant polyQ proteins with expanded tracts results in aggregation, causing cytotoxicity. Oxidative stress in HD has been documented in humans as important to disease progression. Using yeast cells as a model of HD, we report that when grown at high glucose concentration, cells expressing mutant polyQ do not show apparent oxidative stress. At higher cell densities, when glucose becomes limiting and cells are metabolically shifting from fermentation to respiration, protein oxidation and catalase activity increases in relation to the length of the polyQ tract. Oxidative stress, either endogenous as a result of mutant polyQ expression or exogenously generated, increases Sir2 levels. Δ sir2 cells expressing expanded polyQ lengths show signs of oxidative stress even at the early exponential phase. In a wild-type background, isonicotinamide, a Sir2 activator, decreases mutant polyQ aggregation and the stress generated by expanded polyQ. Taken together, these results describe mutant polyQ proteins as being more toxic in respiring cells, causing oxidative stress and an increase in Sir2 levels. Activation of Sir2 would play a protective role against this toxicity.  相似文献   

17.
A significant body of evidence shows that polyglutamine (polyQ) tracts are important for various biological functions. The characteristic polymorphism of polyQ length is thought to play an important role in the adaptation of organisms to their environment. However, proteins with expanded polyQ are prone to form amyloids, which cause diseases in humans and animals and toxicity in yeast. Saccharomyces cerevisiae contain at least 8 proteins which can form heritable amyloids, called prions, and most of them are proteins with glutamine- and asparagine-enriched domains. Yeast prion amyloids are susceptible to fragmentation by the protein disaggregase Hsp104, which allows them to propagate and be transmitted to daughter cells during cell divisions. We have previously shown that interspersion of polyQ domains with some non-glutamine residues stimulates fragmentation of polyQ amyloids in yeast and that yeast prion domains are often enriched in one of these residues. These findings indicate that yeast prion domains may have derived from polyQ tracts via accumulation and amplification of mutations. The same hypothesis may be applied to polyasparagine (polyN) tracts, since they display similar properties to polyQ, such as length polymorphism, amyloid formation and toxicity. We propose that mutations in polyQ/N may be favored by natural selection thus making prion domains likely by-products of the evolution of polyQ/N.  相似文献   

18.
The aggregation behavior of peptides Ac‐VQIVYK‐amide (AcPHF6) and Ac‐QIVYK‐amide (AcPHF5) from the amyloidogenic protein tau was examined by atomic force microscopy (AFM) and fluorescence microscopy. Although AcPHF5 did not show enhancement of thioflavin T (ThT) fluorescence in aqueous buffer, distinct aggregates were discernible when peptide was dissolved in organic solvents such as methanol (MeOH), trifluoroethanol (TFE), and hexafluoroisopropanol (HFIP) dried on mica and examined by AFM. Self‐association was evident even though the peptide did not have the propensity to form secondary structures in the organic solvents. In dried films, the peptide adopts predominantly β‐conformation which results in the formation of distinct aggregates. ThT fluorescence spectra and fluorescence images indicate the formation of fibrils when AcPHF6 solutions in organic solvents were diluted into buffer. AcPHF6 had the ability to organize into fibrillar structures when AFM samples were prepared from peptide dissolved in MeOH, TFE, HFIP, and also when diluted into buffer. AcPHF6 showed propensity for β‐structure in aqueous buffer. In MeOH and TFE, AcPHF6 showed helical and β‐structure. Morphology of the fibrils was dependent on peptide conformation in the organic solvents. The structures observed for AcPHF6 are formed rapidly and long incubation periods in the solvents are not necessary. The structures with varying morphologies observed for AcPHF5 and AcPHF6 appear to be mediated by surfaces such as mica and the organic solvents used for dissolution of the peptides. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Reversed‐phase high‐pressure liquid chromatography analysis and purification of three hydrophobic, aggregation‐prone peptides, composed mainly of the transmembrane (TM) sequence, were performed using elution systems containing 2,2,2‐trifluoroethanol (TFE). The addition of 10–16% TFE to a common mobile phase, such as a water/acetonitrile/propanol (PrOH) or a water/PrOH/formic acid system, markedly improved the chromatographic separation of these peptides. The superior performance of TFE‐containing systems in separating peptides over water/PrOH/formic acid systems [Bollhagen R. et al., J. Chromatogr. A, 1995; 711 : 181–186.] clearly demonstrated that adding TFE to the mobile phase is one of best methods for TM‐peptide purification. Characterization of the potential side reactions using MALDI and ESI‐LIT/Orbitrap mass spectrometry indicated that prolonged incubation of peptides in a mixture of TFE–formic acid possibly induces O‐formylation of the Ser residue and N‐formylation of the N‐terminus of peptides. The conditions for selective removal of the formyl groups from TM peptides were also screened. We believe that these results will expand our ability to analyze and prepare hydrophobic, aggregation‐prone TM peptides and proteins. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
At least nine inherited neurodegenerative diseases, including Huntington's, are caused by poly(L-glutamine) (polyGln, polyQ) expansions > 35-40 repeats in widely or ubiquitously expressed proteins. Except for their expansions, these proteins have no sequence homologies, and their functions mostly remain unknown. Although each disease is characterized by a distinct pathology specific to a subset of neuronal cells, the formation of neuronal intranuclear aggregates containing protein with an expanded polyQ is the hallmark and common feature to most polyQ disorders. The neurodegeneration is thought to be caused by a toxic gain of function that occurs at the protein level and depends on the length of the expansion: Longer repeats cause earlier age of onset and more severe symptoms. To address whether there is a structural difference between polyQ having < 40 versus > 40 residues, we undertook an X-ray fiber diffraction study of synthetic polyQ peptides having varying numbers of residues: Ac-Q8-NH2, D2Q15K2, K2Q28K2, and K2Q45K2. These particular lengths bracket both the range of normalcy (9-36 repeats) and the pathological (45 repeats), and therefore could be indicative of the structural changes expected in expanded polyQ domains. Contrary to expectations of different length-dependent morphologies, we accounted for all the X-ray patterns by slablike, beta-sheet structures, approximately 20 A thick in the beta-chain direction, all having similar monoclinic lattices. Moreover, the slab thickness indicates that K2Q45K2, rather than forming a water-filled nanotube, must form multiple reverse turns.  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