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.
Electronic and vibrational circular dichroism are often used to determine the secondary structure of proteins, because each secondary structure has a unique spectrum. Little is known about the vibrational circular dichroic spectroscopic features of the β‐hairpin. In this study, the VCD spectral features of a decapeptide, YYDPETGTWY (CLN025), which forms a stable β‐hairpin that is stabilized by intramolecular weakly polar interactions and hydrogen bonds were determined. Molecular dynamics simulations and ECD spectropolarimetry were used to confirm that CLN025 adopts a β‐hairpin in water, TFE, MeOH, and DMSO and to examine differences in the secondary structure, hydrogen bonds, and weakly polar interactions. CLN025 was synthesized by microwave‐assisted solid phase peptide synthesis with Nα‐Fmoc protected amino acids. The VCD spectra displayed a (?,+,?) pattern with bands at 1640 to 1656 cm?1, 1667 to 1687 cm?1, and 1679 to 1686 cm?1 formed by the overlap of a lower frequency negative couplet and a higher frequency positive couplet. A maximum IR absorbance was observed at 1647 to 1663 cm?1 with component bands at 1630 cm?1, 1646 cm?1, 1658 cm?1, and 1675 to 1680 cm?1 that are indicative of the β‐sheet, random meander, either random meander or loop and turn, respectively. These results are similar to the results of others, who examined the VCD spectra of β‐hairpins formed by DPro‐Xxx turns and indicated that observed pattern is typical of β‐hairpins. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 442–450, 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   

2.
Riboswitch regulation of gene expression requires ligand‐mediated RNA folding. From the fluorescence lifetime distribution of bound 2‐aminopurine ligand, we resolve three RNA conformers (Co, Ci, Cc) of the liganded G‐ and A‐sensing riboswitches from Bacillus subtilis. The ligand binding affinities, and sensitivity to Mg2+, together with results from mutagenesis, suggest that Co and Ci are partially unfolded species compromised in key loop‐loop interactions present in the fully folded Cc. These data verify that the ligand‐bound riboswitches may dynamically fold and unfold in solution, and reveal differences in the distribution of folded states between two structurally homologous purine riboswitches: Ligand‐mediated folding of the G‐sensing riboswitch is more effective, less dependent on Mg2+, and less debilitated by mutation, than the A‐sensing riboswitch, which remains more unfolded in its liganded state. We propose that these sequence‐dependent RNA dynamics, which adjust the balance of ligand‐mediated folding and unfolding, enable different degrees of kinetic discrimination in ligand binding, and fine‐tuning of gene regulatory mechanisms. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 953–965, 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   

3.
This study is aimed at elucidating the structure of a novel T‐cell adhesion inhibitor, cyclo(1,8)‐CPRGGSVC using one‐ and two‐dimensional (2D) 1H NMR and molecular dynamics (MD) simulation. The peptide is derived from the sequence of its parent peptide cIBR (cyclo(1,12)‐PenPRGGSVLVTGC), which is a fragment of intercellular adhesion molecule‐1 (ICAM‐1). Our previous results show that the cyclo(1,8)‐CPRGGSVC peptide binds to the LFA‐1 I‐domain and inhibits heterotypic T‐cell adhesion, presumably by blocking the LFA‐1/ICAM‐1 interactions. The structure of the peptide was determined using NMR and MD simulation in aqueous solution. Our results indicate that the peptide adopts type‐I β‐turn conformation at the Pro2‐Arg3‐Gly4‐Gly5 (PRGG) sequence. The β‐turn structure at the PRGG motif is well conserved in cIBR peptide and ICAM‐1 receptor, which suggests the importance of the PRGG motif for the biological activity of cyclo(1,8)‐CPRGGSVC peptide. Meanwhile, the Gly5‐Ser6‐Val7‐Cys8‐Cys1 (GSVCC) sequence forms a “turn‐like” random coil structure that does not belong to any structured motif. Therefore, cyclo(1,8)‐CPRGGSVC peptide has only one structured region at the PRGG sequence, which may play an important role in the binding of the peptide to the LFA‐1 I‐domain. The conserved β‐turn conformation of the PRGG motif in ICAM‐1, cIBR, and cyclo(1,8)‐CPRGGSVC peptides can potentially be used to design peptidomimetics. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 633–641, 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 modes of binding of 5′‐[4‐(aminoiminomethyl)phenyl]‐[2,2′‐Bifuran]‐5‐carboximidamide (DB832) to multi‐stranded DNAs: human telomere quadruplex, monomolecular R‐triplex, pyr/pur/pyr triplex consisting of 12 T*(T·A) triplets, and DNA double helical hairpin were studied. The optical adsorption of the ligand was used for monitoring the binding and for determination of the association constants and the numbers of binding sites. CD spectra of DB832 complexes with the oligonucleotides and the data on the energy transfer from DNA bases to the bound DB832 assisted in elucidating the binding modes. The affinity of DB832 to the studied multi‐stranded DNAs was found to be greater (Kass ≈ 107M?1) than to the duplex DNA (Kass ≈ 2 × 105M?1). A considerable stabilizing effect of DB832 binding on R‐triplex conformation was detected. The nature of the ligand tight binding differed for the studied multi‐stranded DNA depending on their specific conformational features: recombination‐type R‐triplex demonstrated the highest affinity for DB832 groove binding, while pyr/pur/pyr TTA triplex favored DB832 intercalation at the end stacking contacts and the human telomere quadruplex d[AG3(T2AG3)3] accommodated the ligand in a capping mode. Additionally, the pyr/pur/pyr TTA triplex and d[AG3(T2AG3)3] quadruplex bound DB832 into their grooves, though with a markedly lesser affinity. DB832 may be useful for discrimination of the multi‐sranded DNA conformations and for R‐triplex stabilization. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 8–20, 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   

5.
Glucose acts as a β‐cell stimulus factor and leads to cellular responses that involve a large amount of biomolecule formation, relocation, and transformation. We hypothesize that information about these changes can be obtained in real‐time by laser tweezers Raman spectroscopy. To test this hypothesis, repeated measurements designs in accordance with the application of Raman spectroscopy detection were used in the current experiment. Single rat β‐cells were measured by Raman spectroscopy in 2.8 mmol/l glucose culture medium as a basal condition. After stimulation with high glucose (20 mmol/l), the same cells were measured continuously. Each cell was monitored over a total time span of 25 min, in 5 min intervals. During this period of time, cells were maintained at an appropriate temperature controlled by an automatic heater, to provide near‐physiological conditions. It was found that some significant spectral changes induced by glucose were taking place during the stimulation time course. The most noticeable changes were the increase of spectral intensity at the 1002, 1085, 1445, and 1655 cm?1 peaks, mainly corresponding to protein and lipid. We speculate that these changes might have to do with β‐cell protein and lipid synthesis. Using laser tweezers Raman spectroscopy in combination with glucose stimulation, optical spectral information from rat β‐cells was received and analyzed. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 587–594, 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.
As a member of intrinsically unstructured protein family, β‐casein (β‐CN) contains relatively high amount of prolyl residues, adopts noncompact and flexible structure and exhibits chaperone‐like activity in vitro. Like many chaperones, native β‐CN does not contain cysteinyl residues and exhibits strong tendencies for self‐association. The chaperone‐like activities of three recombinant β‐CNs wild type (WT) β‐CN, C4 β‐CN (with cysteinyl residue in position 4) and C208 β‐CN (with cysteinyl residue in position 208), expressed and purified from E. coli, which, consequently, lack the phosphorylated residues, were examined and compared with that of native β‐CN using insulin and alcohol dehydrogenase as target/substrate proteins. The dimers (β‐CND) of C4‐β‐CN and C208 β‐CN were also studied and their chaperone‐like activities were compared with those of their monomeric forms. Lacking phosphorylation, WT β‐CN, C208 β‐CN, C4 β‐CN and C4 β‐CND exhibited significantly lower chaperone‐like activities than native β‐CN. Dimerization of C208 β‐CN with two distal hydrophilic domains considerably improved its chaperone‐like activity in comparison with its monomeric form. The obtained results demonstrate the significant role played by the polar contributions of phosphorylated residues and N‐terminal hydrophilic domain as important functional elements in enhancing the chaperone‐like activity of native β‐CN. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 623–632, 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   

7.
We tested directly the differences in the aggregation kinetics of three important β amyloid peptides, the full‐length Aβ1‐42, and the two N‐terminal truncated and pyroglutamil modified Aβpy3‐42 and Aβpy11‐42 found in different relative concentrations in the brains in normal aging and in Alzheimer disease. By following the circular dichroism signal and the ThT fluorescence of the solution in phosphate buffer, we found substantially faster aggregation kinetics for Aβpy3‐42. This behavior is due to the particular sequence of this peptide, which is also responsible for the specific oligomeric aggregation states, found by TEM, during the fibrillization process, which are very different from those of Aβ1‐42, more prone to fibril formation. In addition, Aβpy3‐42 is found here to have an inhibitory effect on Aβ1‐42 fibrillogenesis, coherently with its known greater infective power. This is an indication of the important role of this peptide in the aggregation process of β‐peptides in Alzheimer disease. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 861–873, 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   

8.
While end capping in α‐helices is well understood, the concept of capping a β‐hairpin is a relatively recent development; to date, favorable Coulombic interactions are the only example of sidechains at the termini influencing the overall stability of a β‐hairpin. While cross‐strand hydrophobic residues generally provide hairpin stabilization, particular when flanking the turn region, those remote from this location appear to provide little stabilization. While probing for an optimal residue at a hydrogen bond position near the terminus of a designed β‐hairpin a conservative, hydrophobic, V → I mutation was observed to not only result in a significant change in fold population but also effected major changes in the structuring shifts at numerous sites in the peptide. Mutational studies reveal that there is an interaction between the sidechain at this H‐bonded site and the sidechain at the C‐terminal non‐H‐bonded site of the hairpin. This interaction, which appears to be hydrophobic in character, requires a highly twisted hairpin structure. Modifications at the C‐terminal site, for example an E → A mutation (ΔΔGU = 6 kJ/mol), have profound affects on fold structure and stability. The data suggests that this may be a case of hairpin end capping by the formation of a hydrophobic cluster. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 557–564, 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.
A 34‐residue α/β peptide [IG(28–61)], derived from the C‐terminal part of the B3 domain of the immunoglobulin binding protein G from Streptoccocus, was studied using CD and NMR spectroscopy at various temperatures and by differential scanning calorimetry. It was found that the C‐terminal part (a 16‐residue‐long fragment) of this peptide, which corresponds to the sequence of the β‐hairpin in the native structure, forms structure similar to the β‐hairpin only at T = 313 K, and the structure is stabilized by non‐native long‐range hydrophobic interactions (Val47–Val59). On the other hand, the N‐terminal part of IG(28–61), which corresponds to the middle α‐helix in the native structure, is unstructured at low temperature (283 K) and forms an α‐helix‐like structure at 305 K, and only one helical turn is observed at 313 K. At all temperatures at which NMR experiments were performed (283, 305, and 313 K), we do not observe any long‐range connectivities which would have supported packing between the C‐terminal (β‐hairpin) and the N‐terminal (α‐helix) parts of the sequence. Such interactions are absent, in contrast to the folding pathway of the B domain of protein G, proposed recently by Kmiecik and Kolinski (Biophys J 2008, 94, 726–736), based on Monte‐Carlo dynamics studies. Alternative folding mechanisms are proposed and discussed. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 469–480, 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   

10.
Modified internucleotide linkage featuring the C3′‐O‐P‐CH2‐O‐C4″ phosphonate grouping as an isosteric alternative to the phosphodiester C3′‐O‐P‐O‐CH2‐C4″ bond was studied in order to learn more on its stereochemical arrangement, which we showed earlier to be of prime importance for the properties of the respective oligonucleotide analogues. Two approaches were pursued: First, the attempt to prepare the model dinucleoside phosphonate with 13C‐labeled CH2 group present in the modified internucleotide linkage that would allow for a more detailed evaluation of the linkage conformation by NMR spectroscopy. Second, the use of ab initio calculations along with molecular dynamics (MD) simulations in order to observe the most populated conformations and specify main structural elements governing the conformational preferences. To deal with the former aim, a novel synthesis of key labeled reagent (CH3O)2P(O)13CH2OH for dimer preparation had to be elaborated using aqueous 13C‐formaldehyde. The results from both approaches were compared and found consistent. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 514–529, 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.
Mucins form a group of heavily O‐glycosylated biologically important glycoproteins that are involved in a variety of biological functions, including modulating immune response, inflammation, and adhesion. Mucins are also involved in cancer and metastasis and often express diagnostic cancer antigens. Recently, a modified porcine submaxillary mucin (Tn‐PSM) containing GalNAcα1‐O‐Ser/Thr residues was shown to bind to soybean agglutinin (SBA) with ~106‐fold enhanced affinity relative to GalNAcα1‐O‐Ser, the pancarcinoma carbohydrate antigen. In this study, dynamic force spectroscopy is used to investigate molecular pairs of SBA and Tn‐PSM. A number of force jumps that demonstrate unbinding or rebinding events were observed up to a distance equal to 2.0 μm, consistent with the length of the mucin chain. The unbinding force increased from 103 to 402 pN with increasing force loading rate. The position of the activation barrier in the energy landscape of the interaction was 0.1 nm. The lifetime of the SBA–TnPSM complex in the absence of applied force was determined to be in the range 1.3–1.9 s. Kinetic parameters describing the rate of dissociation of other sugar lectin interactions are in the range 3.3 × 10?3–2.5 × 10?3 s. The long lifetime of the SBA‐TnPSM complex is compatible with a binding model in which lectin molecules “bind and jump” from α‐GalNAc residue to α‐GalNAc residue along the polypeptide chain of Tn‐PSM before dissociating. These findings have important implications for the molecular recognition properties of mucins. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 719–728, 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   

12.
Mei Zhang 《Biopolymers》2010,93(2):121-131
Recently, we isolated and purified a neutral polysaccharide (PGN) from edible fungus Pleurotus geestanus. Its structure was characterized by a range of physical–chemical methods, including high performance anion exchange chromatography, uronic acid, and protein analyses, size exclusion chromatography with ultraviolet, refractive index and light scattering detectors, and nuclear magnetic resonance. Our results revealed that PGN is a novel β‐(1→3)‐D ‐glucan with glucose attached to every other sugar residues at Position 6 in the backbone. It has a degree of branching of 1/2. Such structure is different from typical β‐(1→3)‐D ‐glucans schizophyllan and lentinan in which DB is 1/3 and 2/5, respectively. Rheological study showed a very interesting melting behavior of PGN in water solution: heating PGN in water leads to two transitions, in the range of 8–12.5°C and 25–60°C, respectively. The melting behavior and conformational changes were characterized by rheometry, micro‐differential scan calorimetry, atomic force microscopy, static and dynamic light scattering at different temperatures. The first heating‐induced transition corresponds to the disintegration of polymer bundles into small helical clusters, resembling the heating‐induced dissociation of SPG in water at 7°C; the second one might correspond to the dissociation of helical strands to individual chains. The ability of PGN to undergo a conformation/viscosity transition in water upon heating is very valuable to immobilize cells or enzymes or therapeutic DNA/RNA, which makes PGN a potentially useful biomaterial. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 121–131, 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   

13.
14.
Alkaline tropoelastin solutions (pH 11) were optically clear at low temperatures, but a firm gel formed when the temperature was raised to 37°C. Reversion to a clear solution took place if the temperature was lowered to below 20°C within less than 2 h, but not if 37°C was maintained for several hours. The precipitated elastin‐like hydrogel thus formed did not visually redissolve at low temperatures. Tropoelastin hydrogel was stable to subsequent washings with alkaline solution at 37°C, but at 4°C some hydrogel redissolved showing that association is at least partly reversible. Washing the hydrogel with neutral 8M urea solution at 4°C dissolved less than 10% of tropoelastin in 24 h. We characterized this phenomenon by combining temperature‐controlled light microscopy analysis, 1H NMR spectroscopy (temperature, diffusion, and relaxation time studies), and UV‐absorption‐based concentration measurements. The self‐association of tropoelastin at pH 11 is due to hydrophobic interactions in an emulsion‐like system in which the spherules coalesce in a manner like a water‐based latex paint that forms a durable hydrophobic sheet as water and the organic solvent evaporate. In the present case, the sedimentation and entanglement of the tropoelastin porous sheets means that reverse dissolution is a kinetically slow process. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 321–330, 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   

15.
Tropomyosin (Tm) is a dimeric coiled‐coil protein that polymerizes through head‐to‐tail interactions. These polymers bind along actin filaments and play an important role in the regulation of muscle contraction. Analysis of its primary structure shows that Tm is rich in acidic residues, which are clustered along the molecule and may form sites for divalent cation binding. In a previous study, we showed that the Mg2+‐induced increase in stability of the C‐terminal half of Tm is sensitive to mutations near the C‐terminus. In the present report, we study the interaction between Mg2+ and full‐length Tm and smaller fragments corresponding to the last 65 and 26 Tm residues. Although the smaller Tm peptide (Tm259‐284(W269)) is flexible and to large extent unstructured, the larger Tm220‐284(W269) fragment forms a coiled coil in solution whose stability increases significantly in the presence of Mg2+. NMR analysis shows that Mg2+ induces chemical shift perturbations in both Tm220‐284(W269) and Tm259‐284(W269) in the vicinity of His276, in which are located several negatively charged residues. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 583–590, 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   

16.
Fourier transform infrared (FT‐IR) spectroscopy combined with 2D correlation spectroscopy has been used to offer some information about stability and structure of some soluble elastins. Temperature has been chosen as the perturbation to monitor the infrared behavior of various soluble elastins, namely, α‐elastin p, α‐elastin, and k‐elastin. In the 3800–2700 cm?1 region, the H‐containing groups were analyzed. The bonded hydroxyls are found to decrease prior to the NH‐related hydrogen bonds and also to the conformational reorganization of hydrocarbon chains. The transition temperatures were evaluated and they were found to agree with those obtained from DSC data. The FTIR spectra and their 2nd derivatives denote that α‐ elastins exhibited amide‐I, ‐II and ‐III bands at 1656, 1539 and 1236 cm?1, respectively, while in k‐elastin these bands were found at 1652 cm?1 for amide I, 1540 cm?1 for amide II and 1248 cm?1 for amide III. The macroscopic IR finger‐print method, which combines: general IR spectra, secondary derivative spectra, and 2D‐IR correlation spectra, is useful to discriminate different elastins. Thus using the differences of the position and intensity of the bands from “fingerprint region” of studied elastins, which include the peaks assigned to C?O, C? C groups from α‐helix, β‐turn, and the peaks assigned to the amide groups, it is possible to identify and discriminate elastins from each others. Furthermore, the pattern of 2D‐IR correlation spectra under thermal perturbation, allow their direct identification and discrimination. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1072–1084, 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.
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   

18.
The loss of metal homeostasis and the toxic effect of metal ion are important events in neurodegenerative and age‐related diseases, such as Alzheimer's disease (AD). For the first time, we investigated the impacts of mercury(II) ions on the folding and aggregation of Alzheimer's tau fragment R2 (residues 275‐305: VQIIN KKLDL SNVQS KCGSK DNIKH VPGGGS), corresponding to the second repeat unit of the microtubule‐binding domain, which was believed to be pivotal to the biochemical properties of full tau protein. By ThS fluorescence assay and electron microscopy, we found that mercury(II) dramatically promoted heparin‐induced aggregation of R2 at an optimum molar ratio of 1: 2 (metal: protein), and the resulting R2 filaments became smaller. Isothermal titration calorimetry (ITC) experiment revealed that the strong coordination of mercury(II) with R2 was an enthalpy‐controlled, entropy‐decreased thermodynamic process. The exceptionally large magnitude of heat release (ΔH1 = ?34.8 Kcal mol?1) suggested that the most possible coordinating site on the R2 peptide chain was the thiol group of cysteine residue (Cys291), and this was further confirmed by a control experiment using Cys291 mutated R2. Circular dichroism spectrum demonstrated that this peptide underwent a significant conformational change from random coil to β‐turn structure upon its binding to mercury(II) ion. This study was undertaken to better understand the mechanism of tau aggregation, and evaluate the possible role of mercury(II) in the pathogenesis of AD. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1100–1107, 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   

19.
We report here, the design and synthesis of a positional scanning synthetic combinatorial library for the identification of novel peptide ligands targeted against the cancer‐specific epidermal growth factor tyrosine kinase receptor mutation variant III (EGFRvIII). This receptor is expressed in several kinds of cancer, in particular, ovarian, glioblastomas, and breast cancer, but not in normal tissue. The library consisted of six individual positional sublibraries in the format, H‐O1–6XXXXX‐NH2, O being one of the 19 proteinogenic amino acids (cysteine omitted) and X an equimolar mixture of these. The library consisted of 114 mixtures in total. Using a biotin‐streptavidin assay, the binding of each sublibrary to NR6M, NR6W‐A, and NR6 cells was tested. These cells express EGFRvIII, EGFR, and neither of the receptors, respectively. The result from each sublibrary was examined to identify the most active amino acid residue at each position. On the basis of this knowledge, eight peptides were synthesized and tested for binding to EGFRvIII. We identified one peptide, H‐FALGEA‐NH2, that showed more selective binding to the mutated receptor than the EGFRvIII specific peptide PEPHC1. This study demonstrates the value of using mixture‐based combinatorial positional scanning libraries for the identification of novel peptide ligands targeted against the cancer‐specific EGFRvIII. Our best candidate H‐FALGEA‐NH2 will be radioactively labeled and evaluated as an imaging agent for positron emission tomography investigation for diagnosis, staging, and monitoring of therapy of various types of cancer. © 2008 Wiley Periodicals, Inc. Biopolymers 91: 201–206, 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   

20.
Salmon calcitonin (sCT) was selected as a model protein drug for investigating its intrinsic thermal stability and conformational structure in the solid and liquid states by using a Fourier transform infrared (FT‐IR) microspectroscopy with or without utilizing thermal analyzer. The spectral correlation coefficient (r) analysis between two second‐derivative IR spectra was applied to quantitatively estimate the structural similarity of sCT in the solid state before and after different treatments. The thermal FT‐IR microspectroscopic data clearly evidenced that sCT in the solid state was not effected by temperature and had a thermal reversible property during heating–cooling process. Moreover, the high r value of 0.973 or 0.988 also evidenced the structural similarity of solid‐state sCT samples before and after treatments. However, sCT in H2O exhibited protein instability and thermal irreversibility after incubation at 40°C. The temperature‐induced conformational changes of sCT in H2O was occurred to transform the α‐helix/random coil structures to β‐sheet structure and also resulted in the formation of intramolecular and intermolecular β‐sheet structures. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 200–207, 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   

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