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 468 毫秒
1.
In order to explore whether hydrophobicity or charge at the 71th residue is important for chaperone‐like activity, wild type (wt) and two mutant αA‐crystallins, F71W and F71R, with an extra 11 residues including 6 histidines tagged at the N‐terminal were prepared. Chaperone‐like activity toward dithiotreitol (DTT) induced insulin aggregation showed the 11 extra residues had little effect. The activity for F71W was as active as the wild type, while F71R was much less active. For the wild type and F71W, the activity increased with the increase of molar quantity; surprisingly, for F71R it caused a drastic decline. ANS fluorescence measurements showed F71R had the highest surface hydrophobicity, suggesting positive charge at this site caused the conformational alteration leading to the increase of hydrophobicity and this hydrophobicity resulted in the formation of high molecular weight aggregates (> 2000 kDa) observed from the thermal stability study. The thermal stability study revealed F71W was the most thermally stable crystallin. Chaperone‐like activity studies at various temperatures suggested the monomer and oligomer, 650 kDa, were responsible for chaperone function and surface hydrophobicity was not the prerequisite for activity. Near‐UV CD studies showed β‐sheet was the major secondary structure for all three crystallins and F71R had a higher percentage of random coil.  相似文献   

2.
Huntington's disease is a genetic neurodegenerative disorder caused by an expansion in a polyglutamine domain near the N‐terminus of the huntingtin (htt) protein that results in the formation of protein aggregates. Here, htt aggregate structure has been examined using hydrogen–deuterium exchange techniques coupled with tandem mass spectrometry. The focus of the study is on the 17‐residue N‐terminal flanking region of the peptide that has been shown to alter htt aggregation kinetics and morphology. A top‐down sequencing strategy employing electron transfer dissociation is utilized to determine the location of accessible and protected hydrogens. In these experiments, peptides aggregate in a deuterium‐rich solvent at neutral pH and are subsequently subjected to deuterium–hydrogen back‐exchange followed by rapid quenching, disaggregation, and tandem mass spectrometry analysis. Electrospray ionization of the peptide solution produces the [M + 5H]5+ to [M + 10H]10+ charge states and reveals the presence of multiple peptide sequences differing by single glutamine residues. The [M + 7H]7+ to [M + 9]9+ charge states corresponding to the full peptide are used in the electron transfer dissociation analyses. Evidence for protected residues is observed in the 17‐residue N‐terminal tract and specifically points to lysine residues as potentially playing a significant role in htt aggregation. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

3.
4.
Matrix‐assisted laser desorption/ionization in‐source decay (MALDI‐ISD) is initiated by hydrogen transfer from matrix molecules to the carbonyl oxygen of peptide backbone with subsequent radical‐induced cleavage leading to c′/z? fragments pair. MALDI‐ISD is a very powerful method to obtain long sequence tags from proteins or to do de novo sequencing of peptides. Besides classical fragmentation, MALDI‐ISD also shows specific fragments for which the mechanism of formation enlightened the MALDI‐ISD process. In this study, the MALDI‐ISD mechanism is reviewed, and a specific mechanism is studied in details: the N‐terminal side of Cys residue (Xxx‐Cys) is described to promote the generation of c′ and w fragments in MALDI‐ISD. Our data suggest that for sequences containing Xxx‐Cys motifs, the N–Cα bond cleavage occurs following the hydrogen attachment to the thiol group of Cys side‐chain. The c?/w fragments pair is formed by side‐chain loss of the Cys residue with subsequent radical‐induced cleavage at the N–Cα bond located at the left side (N‐terminal direction) of the Cys residue. This fragmentation pathway preferentially occurs at free Cys residue and is suppressed when the cysteines are involved in disulfide bonds. Hydrogen attachment to alkylated Cys residues using iodoacetamide gives free Cys residue by the loss of ?CH2CONH2 radical. The presence of alkylated Cys residue also suppress the formation of c?/w fragments pair via the (Cβ)‐centered radical, whereas w fragment is still observed as intense signal. In this case, the z? fragment formed by hydrogen attachment of carbonyl oxygen followed side‐chain loss at alkylated Cys leads to a w fragment. Hydrogen attachment on peptide backbone and side‐chain of Cys residue occurs therefore competitively during MALDI‐ISD process. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

5.
Three 2′‐phenanthrenyl‐C‐deoxyribonucleosides with donor (phenNH2), acceptor (phenNO2), or no (phenH) substitution on the phenanthrenyl core were synthesized and incorporated into oligodeoxyribonucleotides. Duplexes containing either one or three consecutive phenR residues, which were located opposite each other, were formed. Within these residues, the phenR residues are expected to recognize each other through interstrand stacking interactions, in much the same way as described previously for biphenyl DNA. The thermal, thermodynamic, and fluorescence properties of such duplexes were determined by UV melting analysis and fluorescence spectroscopy. Depending on the nature of the substituent, the thermal stability of single‐modified duplexes can vary between ?2.7 to +11.3 °C in Tm and that of triple‐modified duplexes from +7.8 to +11.1 °C. Van′t Hoff analysis suggested that the observed higher thermodynamic stability in phenH‐ and phenNO2‐containing duplexes is of enthalpic origin. A single phenH or phenNO2 residue in a bulge position also stabilizes a corresponding duplex. If a phenNO2 residue is placed in a bulge position next to a base mismatch this can lead, in a sequence‐dependent manner, to duplex destabilization. The phenNO2 residue was found to be a highly efficient (10–100‐fold) quencher of phenH and phenNH2 fluorescence if placed in the opposite position to the fluorophores. When phenH and phenNH2 residues were placed opposite each other, efficient quenching of phenH and enhancement of phenNH2 fluorescence was found, which is an indicator for electron‐ or energy‐transfer processes between the aromatic units.  相似文献   

6.
Cyclic homologated amino acids are important building blocks for the construction of helical foldamers. N‐aminoazetidine‐2‐carboxylic acid (AAzC), an aza analogue of trans‐2‐aminocyclobutanecarboxylic acid (tACBC), displays a strong hydrazino turn conformational feature, which is proposed to act as an 8‐helix primer. tACBC oligomers bearing a single N‐terminal AAzC residue were studied to evaluate the ability of AAzC to induce and support an 8‐helix along the oligopeptide length. While tACBC homooligomers assume a dominant 12‐helix conformation, the aza‐primed oligomers preferentially adopt a stabilized 8‐helix conformation for an oligomer length up to 6 residues. The (formal) single‐atom exchange at the N terminus of a tACBC oligomer thus contributes to the sustainability of the 8‐helix, which resists the switch to a 12‐helix. This effect illustrates atomic‐level programmable design for fine tuning of peptide foldamer architectures.  相似文献   

7.
Inhibition of pathogenic protein aggregation may be an important and straightforward therapeutic strategy for curing amyloid diseases. Small‐molecule aggregation inhibitors of Alzheimer’s amyloid‐β (Aβ) are extremely scarce, however, and are mainly restricted to dye‐ and polyphenol‐type compounds that lack drug‐likeness. Based on the structure‐activity relationship of cyclic Aβ16–20 (cyclo‐[KLVFF]), we identified unique pharmacophore motifs comprising side‐chains of Leu2, Val3, Phe4, and Phe5 residues without involvement of the backbone amide bonds to inhibit Aβ aggregation. This finding allowed us to design non‐peptidic, small‐molecule aggregation inhibitors that possess potent activity. These molecules are the first successful non‐peptidic, small‐molecule aggregation inhibitors of amyloids based on rational molecular design.  相似文献   

8.
We have investigated the folding pathway of the 36‐residue villin headpiece subdomain (HP‐36) by action‐derived molecular dynamics simulations. The folding is initiated by hydrophobic collapse, after which the concurrent formation of full tertiary structure and α‐helical secondary structure is observed. The collapse is observed to be associated with a couple of specific native contacts contrary to the conventional nonspecific hydrophobic collapse model. Stable secondary structure formation after the collapse suggests that the folding of HP‐36 follows neither the framework model nor the diffusion‐collision model. The C‐terminal helix forms first, followed by the N‐terminal helix positioned in its native orientation. The short middle helix is shown to form last. Signs for multiple folding pathways are also observed. © 2009 Wiley Periodicals, Inc. J Comput Chem 2010  相似文献   

9.
The hydrogenated amorphous carbon nitride (a‐CNx:H) thin films were synthesized on the SS‐304 substrates using a dense plasma focus device. The a‐CNx:H thin films were synthesized using CH4/N2 admixture gas and 20 focus deposition shots on substrates placed at different distances from the anode top. X‐ray photoelectron spectroscopy and Raman analysis confirmed different C–N bonding in the a‐CNx:H thin films. A decrease in the N/C ratio as well as the sp3/sp2 ratio with an increase in the substrate distance has been observed. The higher amount of C–N formation for the film synthesized at 10 cm is observed which decreases with increasing distance. The X‐ray photoelectron spectroscopy and Raman analysis affirmed the C ≡ N presence in all the thin films synthesized at different distances. The morphology of the synthesized a‐CNx:H thin films showed nanoparticles and nanoparticle clusters formation at the surface. The hardness results showed comparatively lower hardness of the a‐CNx:H thin films due to the presence of C ≡ N. The C–N formation with lower amount of C ≡ N and a higher N/C ratio as well as a higher sp3/sp2 ratio for the films synthesized at 10 cm show reasonably higher hardness. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

10.
李英奇  乔秋瑞  杨晓婧  杨斌盛 《中国化学》2005,23(10):1361-1366
The interaction of gallium(Ⅲ) with the ligands containing phenolic group(s), such as salicylic acid, 8-hydroxyquinoline, N,N'-bis(2-hydroxybenzyl)ethylenediamine-N,N'diacetic acid (HBED), N,N'-ethylenebis[2-(o- hydroxyphenyl)glycine (EHPG), and ovotransferrin, was studied, respectively, by means of fluorescence in 0.01 mol/L Hepes at pH 7.4 and room temperature. Fluorescence intensity showed an increase when gallium(Ⅲ) was bound to 8-hydroxyquinoline and HBED. In contrast, it was decreased with the interaction of gallium(Ⅲ) with salicylic acid and EHPG. At pH 7.4, there was N…H-O type intramolecular hydrogen bond in the former, and the latter existed O…H-O type intramolecular hydrogen bond. Fluorescence titration of apoovotransferrin with gallium(Ⅲ) displayed that the fluorescence intensity was decreased at the N-terminal binding site, while enhanced at the C-terminal binding site. It can account for the O…H-O type intramolecular hydrogen bonds for the phenolic groups of Tyr92 and Tyr191 residues at the N-terminal binding site. And there are N…H-O type intramolecular hydrogen bonds for Tyr431 and Tyr524 residues at the C-terminal binding site. In addition, under the same conditions, the conditional binding constant of gallium(Ⅲ) with EHPG or HBED determined by fluorescence method is lg KGa-EHPG=19.18 or lg KGa-HBED= 19.08.  相似文献   

11.
Small oligomers of the amyloid β (Aβ) peptide, rather than the monomers or the fibrils, are suspected to initiate Alzheimer′s disease (AD). However, their low concentration and transient nature under physiological conditions have made structural investigations difficult. A method for addressing such problems has been developed by combining rapid fluorescence techniques with slower two‐dimensional solid‐state NMR methods. The smallest Aβ40 oligomers that demonstrate a potential sign of toxicity, namely, an enhanced affinity for cell membranes, were thus probed. The two hydrophobic regions (residues 10–21 and 30–40) have already attained the conformation that is observed in the fibrils. However, the turn region (residues 22–29) and the N‐terminal tail (residues 1–9) are strikingly different. Notably, ten of eleven known Aβ mutants that are linked to familial AD map to these two regions. Our results provide potential structural cues for AD therapeutics and also suggest a general method for determining transient protein structures.  相似文献   

12.
A new chemodosimeter based on pyridinium‐fused pyridinone iodide ( PI ) has been obtained through a “clean reaction” method. This compound can detect CN? in aqueous solution with a high selectivity and rapid response. The detection of CN? occurs through the nucleophilic attack of CN? on the C?N bond, which induces the destruction of the π‐conjugation on the pyridinium ring. Support of this detection mechanism was obtained by 1H NMR titration, HR‐MS, and DFT calculations. Upon the addition of 10 equivalents CN? to a solution of PI in THF/H2O (1:1, v/v), a 57‐fold enhancement in fluorescence intensity was observed at the maximum emission wavelength of 457 nm. Meanwhile, the maximum absorption wavelength was also blue‐shifted from 447 nm to 355 nm. Other common anions such as BF4?, PF6?, F?, Cl?, Br?, I?, H2PO4?, ClO4?, CH3COO?, NO2?, N3?, and SCN? had little effect on the detection of CN?. The response time of PI for CN? was less than 5 seconds. The detection limit was calculated to be 5.4×10?8 M , which is lower than the maximum permission concentration in drinking water (1.9 μM ) set by the World Health Organization (WHO).  相似文献   

13.
Glycine (Gly) is incorporated in roughly half of all known peptaibiotic (nonribosomally biosynthesized antibiotic peptides of fungal origin) sequences and is the residue with the greatest conformational flexibility. The conformational space of Aib (α‐aminoisobutyric acid) is severely restricted by the second methyl group attached to the Cα atom. Most of the crystal structures containing Aib are N‐terminal protected. Deprotection of the N‐ or C‐terminus of peptides may alter the hydrogen‐bonding scheme and/or the structure and may facilitate crystallization. The structure reported here for glycyl‐α‐aminoisobutyrylglycyl‐α‐aminoisobutyric acid tert‐butyl ester, C16H30N4O5, describes the first N‐terminal‐unprotected (Gly‐Aib)n peptide. The achiral peptide could form an intramolecular hydrogen bond between the C=O group of Gly1 and the N—H group of Aib4. This hydrogen bond is found in all tetrapeptides and N‐terminal‐protected tripeptides containing Aib, apart from one exception. In the present work, this hydrogen bond is not observed (N...O = 5.88 Å). Instead, every molecule is hydrogen bonded to six other symmetry‐related molecules with a total of eight hydrogen bonds per molecule. The backbone conformation starts in the right‐handed helical region (and the left‐handed helical region for the inverted molecule) and reverses the screw sense in the last two residues.  相似文献   

14.
Comparison of the crystal structures of two pentadehydropeptides containing ΔPhe residues, namely (Z,Z)‐N‐(tert‐butoxycarbonyl)glycyl‐α,β‐phenylalanylglycyl‐α,β‐phenylalanylglycine (or Boc0–Gly1–ΔZPhe2–Gly3–ΔZPhe4–Gly5–OH) methanol solvate, C29H33N5O8·CH4O, (I), and (E,E)‐N‐(tert‐butoxycarbonyl)glycyl‐α,β‐phenylalanylglycyl‐α,β‐phenylalanylglycine (or Boc0–Gly1–ΔEPhe2–Gly3–ΔEPhe4–Gly5–OH), C29H33N5O8, (II), indicates that the ΔZPhe residue is a more effective inducer of folded structures than the ΔEPhe residue. The values of the torsion angles ϕ and ψ show the presence of two type‐III′β‐turns at the ΔZPhe residues and one type‐II β‐turn at the ΔEPhe residue. All amino acids are linked trans to each other in both peptides. β‐Turns present in the peptides are stabilized by intramolecular 4→1 hydrogen bonds. Molecules in both structures form two‐dimensional hydrogen‐bond networks parallel to the (100) plane.  相似文献   

15.
By using aryl‐amination chemistry, a series of rodlike 1‐phenyl‐1H‐imidazole‐based liquid crystals (LCs) and related imidazolium‐based ionic liquid crystals (ILCs) has been prepared. The number and length of the C‐terminal chains (at the noncharged end of the rodlike core) and the length of the N‐terminal chain (on the imidazolium unit in the ILCs) were modified and the influence of these structural parameters on the mode of self‐assembly in LC phases was investigated by polarizing microscopy, differential scanning calorimetry, and X‐ray diffraction. For the single‐chain imidazole derivatives nematic phases (N) and bilayer SmA2 phases were found, but upon increasing the number of alkyl chains the LC phases were lost. For the related imidazolium salts LC phases were preserved upon increasing the number and length of the C‐terminal chains and in this series it leads to the phase sequence SmA–columnar (Col)–micellar cubic (CubI/Pm3n). Elongation of the N‐terminal chain gives the reversed sequence. Short N‐terminal chains prefer an end‐to‐end packing of the mesogens in which these chains are separated from the C‐terminal chains. Elongation of the N‐terminal chain leads to a mixing of N‐ and C‐terminal chains, which is accompanied by complete intercalation of the aromatic cores. In the smectic phases this gives rise to a transition from bilayer (SmA2) to monolayer smectic (SmA) phases. For the columnar and cubic phases the segregated end‐to‐end packing leads to core–shell aggregates. In this case, elongation of the N‐terminal chains distorts core–shell formation and removes CubI and Col phases in favor of single‐layer SmA phases. Hence, by tailoring the length of the N‐terminal chain, a crossover from taper‐shaped to polycatenar LC tectons was achieved, which provides a powerful tool for control of self‐assembly in ILCs.  相似文献   

16.
Carbon nitride (CNx) bilayer films with Ti and TiN interlayer were synthesized by cathode arc technique at various nitrogen pressures (PN2). The dependences of microstructure and bonding composition of the films on the PN2 and interlayer were analyzed by Raman spectroscopy and X‐ray photoelectron spectroscopy. Microstructure evolution consisting of the ordering and size of Csp2 clusters, the faction of N–sp3/N–sp2 bonds and graphite‐like/pyridine‐like configurations was dominated by PN2, interlayer and annealing. The results showed that Ti and TiN interlayer decrease the atomic ratio of N/C and increase clustering Csp2. High PN2 induces the formation of C ≡ N and C ? N bonds, the increase of sp2‐bonding content and the growth of Csp2 clusters. A large part of nitrogen atoms are coordinated with sp2‐hybridized carbon (minimum 71% for annealed CNx monolayer). TiN/CNx bilayer had a higher content of pyridine‐like configuration. Morphological characteristics of CNx monolayer and bilayer mainly depend on the surface character (roughness and surface energy) of the sublayer. The internal stress in the as‐deposited Ti/CNx bilayer is smaller, but it after annealing is higher than that of CNx monolayer and TiN/CNx bilayer. These results may be of interest for studying the CNx films with controlled bonding composition and expected engineering properties. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

17.
Self‐assembly of hydrophobic molecules into aqueous colloids contradicts common chemical intuition, but has been achieved through hydration of [Fe(CO){CO(CH2)5CH3}(Cp)(PPh3)] (FpC6). FpC6 has no surface activity, no NMR signals in D2O and no critical aggregation concentration (CAC) in H2O. The molecule, however, contains both acyl and terminal CO groups that are prone to being hydrated. By adding water to a solution in THF, self‐assembly of FpC6 can be initiated through water–carbonyl interactions (WCIs) with the highly polarized acyl CO groups. This aggregation subsequently enhances the hydration of the acyl CO groups and also induces the WCI of otherwise unhydrated terminal CO groups. The resultant metal–carbonyl aggregates have been proved to be bilayer vesicles with iron complexes exposed towards water and alkyl chains forming inner walls (MCsomes). These MCsomes show high structure integration upon dilution due to the hydrophobic nature of the building blocks. The highly polarized CO groups on the surface of the MCsomes result in a negative zeta potential (?65 mV) and create a local electric field, which significantly enhances the IR absorption of CO groups by more than 100‐fold. This is the first discovery of aggregation‐induced self‐enhanced IR absorption (AI‐SRIRA) without the assistant of external dielectric substrates. Highly integrated MCsomes are, therefore, promising as a novel group of materials, for example, for IR‐based sensing and imaging.  相似文献   

18.
《化学:亚洲杂志》2017,12(14):1796-1806
A new class of twinned amphiphiles was developed by conjugating a pair of hydrophilic head groups from mPEG chains (M n: 350 or 1000) and a pair of hydrophobic segments from linear alkyl chains (C11 or C18) through a novel spacer synthesized from glycerol and p ‐hydroxybenzoic acid. The aggregation phenomena of the amphiphiles were proven by DLS and fluorescence experiments, whereas size and morphology of the aggregates were evaluated by cryo‐TEM. The measurements proved the formation of globular, thread‐like or rod‐like micelles as well as planar double‐layer assemblies, depending on the amphiphile's molecular structure. The applicability of these non‐ionic amphiphilic systems as nanocarriers for hydrophobic guest molecules was demonstrated by encapsulating a hydrophobic dye, Nile Red, and a hydrophobic drug, Nimodipine. The transport capacity results for both Nimodipine and Nile Red prove them as a promising candidate for drug delivery.  相似文献   

19.
The biological activity of antibiotic peptaibols has been linked to their ability to aggregate, but the structure–activity relationship for aggregation is not well understood. Herein, we report a systematic study of a class of synthetic helical oligomer (foldamer) composed of aminoisobutyric acid (Aib) residues, which mimic the folding behavior of peptaibols. NMR spectroscopic analysis was used to quantify the dimerization constants in solution, which showed hydrogen‐bond donors at the N terminus promoted aggregation more effectively than similar modifications at the C terminus. Elongation of the peptide chain also favored aggregation. The geometry of aggregation in solution was investigated by means of titrations with [D6]DMSO and 2D NOE NMR spectroscopy, which allowed the NH protons most involved in intermolecular hydrogen bonds in solution to be identified. X‐ray crystallography studies of two oligomers allowed a comparison of the inter‐ and intramolecular hydrogen‐bonding interactions in the solid state and in solution and gave further insight into the geometry of foldamer–foldamer interactions. These solution‐based and solid‐state studies indicated that the preferred geometry for aggregation is through head‐to‐tail interactions between the N and C termini of adjacent Aib oligomers.  相似文献   

20.
《Electroanalysis》2017,29(12):2906-2912
The aggregation of amyloid‐β peptide (Aβ) is believed to play a crucial role in the Alzheimer's disease (AD) pathogenesis and is considered as a therapeutic target for treating AD. The Aβ electrooxidation via a Tyr‐10 residue, sensitive to a depletion of a pool of Aβ monomers and oligomers in the course of Aβ aggregation, may be employed for testing natural and synthetic organic compounds (including short peptides) potentially able to inhibit the pathological Aβ aggregation (antiaggregants). In the present work, using the known peptide antiaggregant RGKLVFFGR‐NH2 (OR2) and its scrambled variant KGLRVGFRF‐NH2 as a control, we demonstrate that the electrochemical method based on electrooxidation of an Aβ42 Tyr‐10 residue, when combined with methods allowing for the evaluation of the Aβ42 aggregate structure and size, can provide essential information regarding the antiaggregant impact on Aβ42 aggregation. Electrochemical measurements were performed using square wave voltammetry on carbon screen printed electrodes whereas the Aβ42 aggregate structure and size were analyzed by means of the conventional thioflavin T (ThT) based fluorescence assay and dynamic light scattering. While inhibiting Aβ42 fibrillation as manifested by the unchanged level of ThT fluorescence, the OR2 peptide antiaggregant had no effect on the decrease of Aβ42 electrooxidation current in the course of Aβ42 aggregation. These observations suggest that OR2 does not stop the aggregation but redirects it into a pathway where amorphous rather than fibrillar aggregates are formed. Hence, the direct electrochemistry appears to offer a simple and cost‐effective approach for probing potential peptide antiaggregants, which is complementary to methods based on detecting Aβ aggregates.  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