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.
Two new rigid bi‐aromatic linkers for synthesis of peptide arrays by SPOT methodology were obtained from cellulose treated with 2,4‐dichloro‐6‐methoxy‐1,3,5‐triazine. Reaction with m‐phenylenediamine gave non‐cleavable TYPE I linker which enabled attachment of the peptides via resistant to harsh reaction conditions amide, ether, and amine bonds. Reaction with 3‐Fmoc‐aminobenzoic acid followed by thermal isomerization of the intermediate “superactive” ester producing an amide‐like bond gave TYPE II linker that was very stable during peptide synthesis. However, the peptide was cleavable, with fragment of the linker, in the presence of 1 M LiOH solution. The uniform loading of the cellulose and efficient synthesis of the peptide array was achieved by using N‐(4,6‐dimethoxy‐1,3,5‐triazin‐1‐yl)‐N‐methylmorpholinium 4‐toluenesulfonate as the coupling reagent.  相似文献   

2.
To develop an optimal attractant for Monochamus saltuarius (Gebler) (Coleoptera: Cerambycidae), the synergistic effects of a few potential attractants (ethanol and α‐pinene as host‐plant volatiles, and ipsenol and ipsdienol as bark beetle pheromones) were tested in a pine forest combined with 2‐(1‐undecyloxy)‐1‐ethanol (monochamol), the aggregation pheromone of Monochamus species, for two consecutive years, 2014 and 2015. Total number of catches was 65 and 33 in 2014 and 2015, respectively. Ethanol or ethanol + monochamol (a base blend) were not attractive to M. saltuarius with no difference from the control. Addition of α‐pinene and ipsdienol to the base blend did not significantly increase catches. However, ipsenol was significantly synergistic to the base blend in attracting M. saltuarius in 2014, and the blend (ipsenol + base blend) attracted meaningfully higher numbers of M. saltuarius in 2015. Our study illustrates the potential for monochamol and ipsenol baits for monitoring and trapping of M. saltuarius in the field.  相似文献   

3.
Gliadin, a protein present in wheat, rye, and barley, undergoes incomplete enzymatic degradation during digestion, producing an immunogenic 33‐mer peptide, LQLQPF(PQPQLPY)3PQPQPF. The special features of 33‐mer that provoke a break in its tolerance leading to gliadin sensitivity and celiac disease remains elusive. Herein, it is reported that 33‐mer gliadin peptide was not only able to fold into polyproline II secondary structure but also depending on concentration resulted in conformational transition and self‐assembly under aqueous condition, pH 7.0. A 33‐mer dimer is presented as one initial possible step in the self‐assembling process obtained by partial electrostatics charge distribution calculation and molecular dynamics. In addition, electron microscopy experiments revealed supramolecular organization of 33‐mer into colloidal nanospheres. In the presence of 1 mM sodium citrate, 1 mM sodium borate, 1 mM sodium phosphate buffer, 15 mM NaCl, the nanospheres were stabilized, whereas in water, a linear organization and formation of fibrils were observed. It is hypothesized that the self‐assembling process could be the result of the combination of hydrophobic effect, intramolecular hydrogen bonding, and electrostatic complementarity due to 33‐mer's high content of proline and glutamine amino acids and its calculated nonionic amphiphilic character. Although, performed in vitro, these experiments have revealed new features of the 33‐mer gliadin peptide that could represent an important and unprecedented event in the early stage of 33‐mer interaction with the gut mucosa prior to onset of inflammation. Moreover, these findings may open new perspectives for the understanding and treatment of gliadin intolerance disorders. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 96–106, 2014.  相似文献   

4.
The rotational strengths and the robustness values of amide‐I and amide‐II vibrational modes of For(AA)nNHMe (where AA is Val, Asn, Asp, or Cys, n = 1–5 for Val and Asn; n = 1 for Asp and Cys) model peptides with α‐helix and β‐sheet backbone conformations were computed by density functional methods. The robustness results verify empirical rules drawn from experiments and from computed rotational strengths linking amide‐I and amide‐II patterns in the vibrational circular dichroism (VCD) spectra of peptides with their backbone structures. For peptides with at least three residues (n ≥ 3) these characteristic patterns from coupled amide vibrational modes have robust signatures. For shorter peptide models many vibrational modes are nonrobust, and the robust modes can be dependent on the residues or on their side chain conformations in addition to backbone conformations. These robust VCD bands, however, provide information for the detailed structural analysis of these smaller systems. Chirality 27:625–634, 2015 © 2015 Wiley Periodicals, Inc.  相似文献   

5.
AA‐NADase from Agkistrodon acutus venom is a unique multicatalytic enzyme with both NADase and AT(D)Pase activities. Among all identified NADases, only AA‐NADase contains Cu(II) and has disulfide‐bond linkages between two peptide chains. The effects of the reduction of the disulfide‐bonds and Cu(II) in AA‐NADase by small‐molecule reductants on its NADase and ADPase activities have been investigated by polyacrylamide gel electrophoresis, high performance liquid chromatography, electron paramagnetic resonance spectroscopy and isothermal titration calorimetry. The results show that AA‐NADase has six disulfide‐bonds and fifteen free cysteine residues. L‐ascorbate inhibits AA‐NADase on both NADase and ADPase activities through the reduction of Cu(II) in AA‐NADase to Cu(I), while other reductants, dithiothreitol, glutathione and tris(2‐carboxyethyl)phosphine inhibit both NADase and ADPase activities through the reduction of Cu(II) to Cu(I) and the cleavage of disulfide‐bonds in AA‐NADase. Apo‐AA‐NADase can recover its NADase and ADPase activities in the presence of 1 mM Zn(II). However, apo‐AA‐NADase does not recover any NADase or ADPase activity in the presence of 1 mM Zn(II) and 2 mM TCEP. The multicatalytic activity relies on both disulfide‐bonds and Cu(II), while Cu(I) can not activate the enzyme activities. AA‐NADase is probably only active as a dimer. The inhibition curves for both ADPase and NADase activities by each reductant share a similar trend, suggesting both ADPase and NADase activities probably occur at the same site. In addition, we also find that glutathione and L‐ascorbate are endogenous inhibitors to the multicatalytic activity of AA‐NADase. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 141–149, 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   

6.
The Raman spectra of poly(L -lysine) with various structures, ionized poly(L -glutamic acid), and deuterated N-methylacetamides have been observed using visible and the 257.3-nm laser lines as the light source. Most of the Raman bands with significantly enhanced intensities in the uv-excited spectra of the polymers have been assigned to the vibrations associated with the C?O and C–N stretching modes, the amide I, II, III, I′, II′, and III′, with reference to the results obtained for simple amide molecules including the deuterated N-methylacetamides. Several amide frequencies have been newly identified and the structures of the polymers have been discussed through the comparison of the Raman and ir amide frequencies.  相似文献   

7.
The conformational preferences of blocked alanine dipeptide (ADP), Ac‐Ala‐NHMe, in aqueous solution were studied using vibrational circular dichroism (VCD) together with density functional theory (DFT) calculations. DFT calculations of three most representative conformations of ADP surrounded by six explicit water molecules immersed in a dielectric continuum have proven high sensitivity of amide III VCD band shape that is characteristic for each conformation of the peptide backbone. The polyproline II (PII) and αR conformation of ADP are associated with a positive VCD band while β conformation has a negative VCD band in amide III region. Knowing this spectral characteristic of each conformation allows us to assign the experimental amide III VCD spectrum of ADP. Moreover, the amide III region of the VCD spectrum was used to determine the relative populations of conformations of ADP in water. Based on the interpretation of the amide III region of VCD spectrum we have shown that dominant conformation of ADP in water is PII which is stabilized by hydrogen bonded water molecules between CO and NH groups on the peptide backbone. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 814–818, 2014.  相似文献   

8.
Knowledge of the fold class of a protein is valuable because fold class gives an indication of protein function and evolution. Fold class can be accurately determined from a crystal structure or NMR structure, though these methods are expensive, time-consuming, and inapplicable to all proteins. In contrast, vibrational spectra [infra-red, Raman, or Raman optical activity (ROA)] are rapidly obtained for proteins under wide range of biological molecules under diverse experimental and physiological conditions. Here, we show that the fold class of a protein can be determined from Raman or ROA spectra by converting a spectrum into data of 10 cm−1 bin widths and applying the random forest machine learning algorithm. Spectral data from 605 and 1785 cm−1 were analyzed, as well as the amide I, II, and III regions in isolation and in combination. ROA amide II and III data gave the best performance, with 33 of 44 proteins assigned to one of the correct four top-level structural classification of proteins (SCOP) fold class (all α, all β, α and β, and disordered). The method also shows which spectral regions are most valuable in assigning fold class.  相似文献   

9.
In addition to the well‐known Ca2+ sensor calmodulin, plants possess many calmodulin‐like proteins (CMLs) that are predicted to have specific roles in the cell. Herein, we described the biochemical and biophysical characterization of recombinant Arabidopsis thaliana CML14. We applied isothermal titration calorimetry to analyze the energetics of Ca2+ and Mg2+ binding to CML14, and nuclear magnetic resonance spectroscopy, together with intrinsic and ANS‐based fluorescence, to evaluate the structural effects of metal binding and metal‐induced conformational changes. Furthermore, differential scanning calorimetry and limited proteolysis were used to characterize protein thermal and local stability. Our data demonstrate that CML14 binds one Ca2+ ion with micromolar affinity (Kd ~ 12 µM) and the presence of 10 mM Mg2+ decreases the Ca2+ affinity by ~5‐fold. Although binding of Ca2+ to CML14 increases protein stability, it does not result in a more hydrophobic protein surface and does not induce the large conformational rearrangement typical of Ca2+ sensors, but causes only localized structural changes in the unique functional EF‐hand. Our data, together with a molecular modelling prediction, provide interesting insights into the biochemical properties of Arabidopsis CML14 and may be useful to direct additional studies aimed at understanding its physiological role.  相似文献   

10.
A highly efficient and reproducible two-step liquid—liquid ion-pair extraction technique for the isolation of morphine from biological samples is described. A rapid normal phase high-performance liquid chromatographic procedure coupled with amperometric electrochemical detection has also been developed for subsequent quantification of morphine. Extraction involves the disruption of brain tissue or plasma in methanol, centrifugation, evaporation and reconstitution in ethyl acetate containing 10 mM di-(2-ethylhexyl) phosphoric acid, a liquid cation-exchanger, and back-extraction into 170 mM orthophosphoric acid. An acidic eluent consisting of acetonitrile—76 mM orthophosphoric acid—ammonia buffer (pH 3.0) (15:85, v/v) in combination with a strong cation-exchange column allows complete separation of morphine and the internal standard, nalbuphine. The limit of detection for morphine is 1.3 ng on-column.  相似文献   

11.
The peripheral anionic site (PAS) of acetylcholinesterase (AChE) is involved in amyloid beta (Aβ) peptides aggregation of Alzheimer's disease (AD). AChE exhibits an aryl acylamidase (AAA) activity along with the well known esterase activity. Numerous studies have reported the beneficiary effect of metal chelators in AD treatment. Hence, a comparative study on the effect of metal chelators on both the esterase and AAA activity of AChE globular (G4 and G1) molecular forms was performed. The inhibitory effect of 1,10‐phenanthroline was high towards AChE esterase activity. The corresponding IC50 values for esterase activity of G4 and G1‐form was 190 µM and 770 µM and for AAA activity it was 270 µM and 2.74 mM, respectively. Kinetic studies on both forms of AChE show that 1,10‐phenanthroline inhibits esterase in competitive and AAA activity in non‐competitive manner. Protection studies further revealed that the nature of 1,10‐phenanthroline inhibition on AChE is through its direct binding to protein rather than its metal chelation property. Molecular docking studies shows orientation of 1,10‐phenathroline in the PAS through hydrophobic interactions with the PAS residues (Trp286, Tyr124 and Tyr341) and hydrogen bonding with Phe295. Further molecular dynamics simulation of “hAChE‐1,10‐phenanthroline” complex revealed that both hydrogen and hydrophobic interaction contribute equally for 1,10‐phenanthroline binding to hAChE. Such an interaction of 1,10‐phenanthroline on PAS may hinder “AChE‐Aβ peptide” complex formation. Hence, 1,10‐phenanthroline can act as a lead molecule for developing drug(s) against AD ailment with dual functions namely, anti‐cholinesterase and anti‐amyloid aggregation potency in a single chemical entity. Proteins 2013. Proteins 2013; 81:1179–1191. © 2013 Wiley Periodicals, Inc.  相似文献   

12.
The storage of protein/peptide hormones within subcellular compartments and subsequent release are crucial for their native function, and hence these processes are intricately regulated in mammalian systems. Several peptide hormones were recently suggested to be stored as amyloids within endocrine secretory granules. This leads to an apparent paradox where storage requires formation of aggregates, and their function requires a supply of non-aggregated peptides on demand. The precise mechanism behind amyloid formation by these hormones and their subsequent release remain an open question. To address this, we examined aggregation and fibril reversibility of a cyclic peptide hormone somatostatin (SST)-14 using various techniques. After proving that SST gets stored as amyloid in vivo, we investigated the role of native structure in modulating its conformational dynamics and self-association by disrupting the disulfide bridge (Cys3–Cys14) in SST. Using two-dimensional NMR, we resolved the initial structure of somatostatin-14 leading to aggregation and further probed its conformational dynamics in silico. The perturbation in native structure (S-S cleavage) led to a significant increase in conformational flexibility and resulted in rapid amyloid formation. The fibrils formed by disulfide-reduced noncyclic SST possess greater resistance to denaturing conditions with decreased monomer releasing potency. MD simulations reveal marked differences in the intermolecular interactions in SST and noncyclic SST providing plausible explanation for differential aggregation and fibril reversibility observed experimentally in these structural variants. Our findings thus emphasize that subtle changes in the native structure of peptide hormone(s) could alter its conformational dynamics and amyloid formation, which might have significant implications on their reversible storage and secretion.  相似文献   

13.
The bacterial Sec pathway is responsible for the translocation of secretory preproteins. During the later stages of transport, the membrane‐embedded signal peptidase I (SPase I) cleaves the signal peptide from a preprotein. We used tryptophan fluorescence spectroscopy of a soluble, catalytically active E. coli SPase I Δ2‐75 enzyme to study its dynamic conformational changes while in solution and when interacting with lipids and signal peptides. We generated four single Trp SPase I Δ2‐75 mutants, W261, W284, W300, and W310. Based on fluorescence quenching experiments, W300 and W310 were found to be more solvent accessible than W261 and W284 in the absence of ligands. W300 and W310 inserted into lipids, consistent with their location at the enzyme's proposed membrane‐interface region, while the solvent accessibilities of W261, W284, and W300 were modified in the presence of signal peptide, suggesting propagation of structural changes beyond the active site in response to peptide binding. The signal peptide binding affinity for the enzyme was measured via FRET experiments and the Kd determined to be 4.4 μM. The location of the peptide with respect to the enzyme was also established; this positioning is crucial for the peptide to gain access to the enzyme active site as it emerges from the translocon into the membrane bilayer. These studies reveal enzymatic structural changes required for preprotein proteolysis as it interacts with its two key partners, the signal peptide and membrane phospholipids. Proteins 2014; 82:596–606. © 2013 Wiley Periodicals, Inc.  相似文献   

14.
A comparative study of substrate specificity of bovine duodenal proteinases—chymotrypsin-like duodenase (ChlD) and dual-specificity duodenase (dsD)—was carried out using oligopeptide substrates (human proinsulin, glucagon, melittin, angiotensinogen fragment 1-14). ChlD displayed mainly chymotrypsin-like properties towards these substrates, hydrolyzing peptide bonds carboxy-terminally to bulky aliphatic or aromatic residues. In melittin, ChlD additionally cleaved peptide bonds after Thr and Ser residues. Dual-specificity duodenase (dsD) significantly restricted its specificity to only trypsin-like or only chymotrypsin-like or displayed full activity, combining both specificities, depending on substrate. Both ChlD and dsD efficiently hydrolyzed a single peptide bond (Phe8–His9) in angiotensinogen fragment 1-14. The kinetic parameters of angiotensinogen fragment 1-14 cleavage by ChlD and dsD were determined (k cat/K m = 80,500 M-1·sec-1 for ChlD and 103,000 M-1·sec-1 for dsD).  相似文献   

15.
S. L. Hsu  W. H. Moore  S. Krimm 《Biopolymers》1976,15(8):1513-1528
Raman spectra of native and solubilized feather keratin have been obtained, and the amide I and amide III regions have been analyzed by band resolution techniques. The amide I region of the native form indicates that at least 64% of the protein has an antiparallel chain pleated sheet structure, the remainder being unordered. For the solubilized keratin all of the protein is in an unordered state. The amide III region is not as easily analyzed into component contributions. Normal vibration analyses on N-acetyl-L -alanine-N-methylamide support the conclusion that the amide III region is not as satisfactory as the amide I region in characterizing unordered structures. Even in the latter case caution must be used, since the observed amide I band is an average over the conformational distribution in the particular unordered system.  相似文献   

16.
Polarization‐resolved Raman microspectroscopy with near‐infrared laser excitation was applied to intact human hair in order to non‐invasively investigate the conformation and orientation of the polypeptide chains. By varying the orientation of the hair shaft relative to the polarization directions of the laser/analyzer, a set of four polarized Raman spectra is obtained; this allows to simultaneously determine both the secondary structure of hair proteins and the orientation of the polypeptide strands relative to the axis of the hair shaft. For the amide I band, results from a quantitative analysis of the polarized Raman spectra are compared with theoretically expected values for fibers with uniaxial symmetry. Based on the polarization behavior of the amide I band and further vibrational bands, a partial ordering of α‐helical polypeptide strands parallel to the hair shaft can be concluded. We suggest that this microspectroscopic approach may be used for human hair diagnostics by detecting structural or orientational alterations of keratins. (© 2008 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim)  相似文献   

17.
The Raman spectrum of chemically denatured lysozyme was studied. The denaturants studied included dimethyl sulfoxide, LiBr, guanidine · HCl, sodium dodecyl sulfate, and urea. Previous studies have shown that the amide I and amide III regions of the Raman spectrum are sensitive to the nature of the hydrogen bond involving the amide group. The intensity of the amide III band at 1260 cm?1 (assigned to strongly hydrogen-bonded α-helix structure) relative to the intensity of the amide III band near 1240 cm?1 (assigned to less strongly hydrogen-bonded groups) is used as a parameter for comparison with other physical parameters used to assess denaturation. The correlation between this Raman parameter and denaturation as evidenced by enzyme activity and viscosity measurements is good, leading to the conclusion that the amide III Raman spectrum is useful for assessing the degree of denaturation. The Raman spectrum clearly depends on the type of denaturant employed, suggesting that there is not one unique denatured state for lysozyme. The data, as interpreted, place constraints on the possible models for lysozyme denaturation. One of these is that the simple two-state model does not seem consistent with the observed Raman spectral changes.  相似文献   

18.
Melittin, the main hemolytic component of honeybee venom, is unfolded in an aqueous environment and folds into an α‐helical conformation in a lipid environment. Membrane fluidity is known to affect the activity and structure of melittin. By combining two structurally sensitive optical methods, circular dichroism (CD) and deep‐ultraviolet resonance Raman spectroscopy (dUVRR), we have identified distinct structural fluctuations in melittin correlated with increased and decreased 1,2‐dimyristoyl‐sn‐glycero‐3‐phosphocholine bilayer fluidities. CD spectra have reduced intensity at temperatures above 22°C and high concentrations of the cholesterol analog 5α‐cholestan‐3β‐ol indicating distortions in the α‐helical structure under these conditions. No increase in the amide S is observed in the temperature‐dependent dUVRR spectra, suggesting an increase in 310‐helical structure with increasing temperatures above 22°C. However, incorporation of 25 mol% 5α‐cholestan‐3β‐ol resulted in a small increase in the amide S intensity indicating partial unfolding of melittin. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 895–902, 2014.  相似文献   

19.
In this study, the transition process of the helix‐sense inversion of poly(β‐phenethyl‐L‐aspartate) was investigated by Raman scattering and 2‐dimensional correlation spectroscopy. Temperature‐dependent Raman spectra were obtained during the helix‐sense inversion. The results of 2‐dimensional correlation analysis in the spectral regions of 1600‐1800 and 3200‐3400 cm?1 showed that the intensity changes of the side‐chain ester C═O stretching bands occurred prior to those of amide A and amide I bands in the unwinding process of αR‐helix on heating. The sequential order of the intensity changes for amide A, amide I, and the side‐chain ester C═O stretching bands during the inversion process was determined. It was found that the conformation change of the side chain occurred prior to that of the main chain for the αR‐helix on heating. Thus, we concluded that the transformation of the backbone chain from right‐handed to left‐handed is triggered by the conformational change of the side chains.  相似文献   

20.
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