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.
The reaction of histidine‐containing polypeptides with toxic and essential metals and the molecular mechanism of complexation has yet to be determined, particularly with respect to the conformational changes of the interacting macromolecules. Therefore, a system of oligopeptides containing histidine residues in various positions of Ala or Gly sequences has been designed and used in heavy metal comparatively binding experiments. The role of spacing residues (Gly and Ala repeats) in selecting the various conformations was investigated. The newly synthesized peptides and metal ion adducts have been characterized by Fourier transform infrared spectroscopy (FTIR) as well as electrospray ion trap mass spectrometry (ESI–MS) and circular dichroism (CD). The analysis of CD‐spectra of the four peptides in water revealed that the secondary structure depends much on the position of each amino acid in the peptide backbone. Our peptides system reveals various binding mechanisms of metal ions to peptides depending on the position of histidine residue and the corresponding conformations of Ala or Gly sequences. Biological and medical consequences of conformational changes of metal‐bound peptides are further discussed. Thus, the binding of heavy metals to four peptides may serve as a model system with respect to the conformational consequences of the metal addition on the amino acid repeats situated in prion protein. © 2010 Wiley Periodicals, Inc. Biopolymers 93:497–508, 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.
With the decline in productivity of drug‐development efforts, novel approaches to rational drug design are being introduced and developed. Naturally occurring and synthetic peptides are emerging as novel promising compounds that can specifically and efficiently modulate signaling pathways in vitro and in vivo. We describe sequence‐based approaches that use peptides to mimic proteins in order to inhibit the interaction of the mimicked protein with its partners. We then discuss a structure‐based approach, in which protein‐peptide complex structures are used to rationally design and optimize peptidic inhibitors. We survey flexible peptide docking techniques and discuss current challenges and future directions in the rational design of peptidic inhibitors. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 505–513, 2009. This article was originally published online as an accepted preprint. The “Published Online”date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

3.
Conformations of two pairs of dehydropeptides with the opposite configuration of the ΔPhe residue, Boc‐Gly‐ΔZPhe‐Gly‐Phe‐OMe ( Z‐ OMe ), Boc‐Gly‐ΔEPhe‐Gly‐Phe‐OMe ( E‐ OMe ), Boc‐Gly‐ΔZPhe‐Gly‐Phe‐p‐NA ( Z‐p‐ NA ), and Boc‐Gly‐ΔEPhe‐Gly‐Phe‐p‐NA ( E‐p‐ NA ) were compared on the basis of CD and NMR studies in MeOH, trifluoroethanol (TFE), MeCN, chloroform, and dimethylsulfoxide (DMSO). The CD results were used as the additional input data for the NMR‐based determination of the detailed solution conformations of the peptides. It was found that E‐ OMe is unordered and Z‐ OMe , Z‐p‐ NA , and E‐p‐ NA adopt the β‐turn conformation. There are two overlapping β‐turns in each of those peptides: type II and type III′ in Z‐ OMe and Z‐p‐ NA , and two type III in E‐p‐ NA . The ordered structure‐inducing properties of ΔZPhe and ΔEPhe in the peptides studied depend on the C‐terminal blocking group. In methyl esters, the ΔZPhe residue is a strong inducer of ordered conformations whereas the ΔEPhe one has no such properties. In p‐nitroanilides, both isomers of ΔPhe cause the peptides to adopt ordered structures to a similar extent. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1055–1064, 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   

4.
Collagen, which is used as a biomaterial, is the most abundant protein in mammals. We have previously reported that a dendrimer modified with collagen model peptides, (Gly‐Pro‐Pro)5, formed a collagen‐like triple‐helical structure, showing thermal reversibility. In this study, various collagen‐mimic dendrimers of different generations and at different binding ratios were synthesized, to investigate the relationship between the peptide clustering effect and the higher order structure formation. The formation of the higher order structure was influenced by the binding ratios of the peptide to the dendrimer, but was not influenced by the dendrimer generation. A spacer, placed between the dendrimer terminal group and the peptide, negatively contributed to the formation of the higher order structure. The collagen model peptides were also attached to poly(allylamine) (PAA) and poly‐L ‐lysine (poly(Lys)) to compare them with the collagen‐mimic dendrimers. The PAA‐based collagen‐mimic compound, bearing more collagen model peptides than the dendrimer, exhibited a thermally stable higher order structure. In contrast, this was not observed for the collagen‐mimic polymers based on poly(Lys). Therefore, dendrimers and vinyl polymers act as a scaffold for collagen model peptides and subsequently induce higher order structures. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 640–648, 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.
Recently, ubiquitin was suggested as a promising anti‐inflammatory protein therapeutic. We found that a peptide fragment corresponding to the ubiquitin50–59 sequence (LEDGRTLSDY) possessed the immunosuppressive activity comparable with that of ubiquitin. CD and NMR spectroscopies were used to determine the conformational preferences of LEDGRTLSDY in solution. The peptide mixture, obtained by pepsin digestion of ubiquitin, was even more potent than the intact protein. Although the peptide exhibited a well‐defined conformation in methanol, its structure was distinct from the corresponding 50–59 fragment in the native ubiquitin molecule. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 423–431, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

6.
The conformational study on Ac‐pSer‐Pro‐NHMe and Ac‐pThr‐Pro‐NHMe peptides has been carried out using hybrid density functional methods with the implicit solvation reaction field theory at the B3LYP/ 6‐311++G(d,p)//B3LYP/6‐31+G(d) level of theory in the gas phase and in solution (chloroform and water). For both pSer‐Pro and pThr‐Pro peptides in the gas phase and in chloroform, the most preferred conformation has the α‐helical structure for the pSer/pThr residue, the down‐puckered polyproline I structure for the Pro residue, and the cis prolyl peptide bond between the two residues, in which two hydrogen bonds between the phosphate oxygens with the backbone N? H groups seem to play a role. However, the trans conformations that have a single hydrogen bond of the phosphate oxygen with either of two backbone N? H groups become most preferred for both peptides in water. This is because the hydration free energy of the anionic oxygen of the phosphate group is expected to dramatically decrease for the cis conformation upon formation of the hydrogen bond with the backbone N? H groups. These calculated results are consistent with the observations by NMR and IR experiments, suggesting the existence of hydrogen bonds between the charged phosphoryl group and the backbone amide protons in solution. The calculated cis populations of 14.7 and 14.2% and rotational barriers of 19.87 and 20.57 kcal/mol to the cis‐to‐trans isomerization for pSer‐Pro and pThr‐Pro peptides in water, respectively, are consistent with the observed values for pSer‐Pro and pThr‐Pro containing peptides from NMR experiments. However, the hydrogen bond between the prolyl nitrogen and the following amide N? H group, which was suggested to be capable of catalyzing the prolyl isomerization, does not play a role in stabilizing the preferred transition state for the pSer/pThr‐Pro peptides in water. Instead, the amide hydrogen of the NHMe group is involved in a bifurcated hydrogen bond with the anionic oxygen and phosphoester oxygen of the phosphate group. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 330–339, 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   

7.
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   

8.
A fluorescently labeled 20‐residue polyglutamic acid (polyE) peptide 20 amino acid length polyglutamic acid (E20) was used to study structural changes which occur in E20 as it co‐aggregates with other unlabeled polyE peptides. Resonance energy transfer (RET) was performed using an o‐aminobenzamide donor at the N‐terminus and 3‐nitrotyrosine acceptor at the C‐terminus of E20. PolyE aggregates were not defined as amyloid, as they were nonfibrillar and did not bind congo red. Circular dichroism measurements indicate that polyE aggregation involves a transition from α‐helical monomers to aggregated β‐sheets. Soluble oligomers are also produced along with aggregates in the reaction, as determined through size exclusion chromatography. Time‐resolved and steady‐state RET measurements reveal four dominant E20 conformations: (1) a partially collapsed conformation (24 Å donor–acceptor distance) in monomers, (2) an extended conformation in soluble oligomers (>29 Å donor–acceptor distance), (3) a minor partially collapsed conformation (22 Å donor‐acceptor distance) in aggregates, and (4) a major highly collapsed conformation (13 Å donor–acceptor distance) in aggregates. These findings demonstrate the use of RET as a means of determining angstrom‐level structural details of soluble oligomer and aggregated states of proteins. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 299–317, 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   

9.
Here we describe the features of a peptide that was selected from the human immunodeficiency virus Type 1 (HIV‐1) Integrase (IN) peptide library which interacts with both, the viral Rev and IN proteins. Because of its ability to stimulate the IN enzymatic activity this peptide was designated INS (IN stimulatory). Modification of its amino acid sequence revealed that replacement of its C‐terminal lysine by glutamic acid (INS K188E) converts it from a stimulatory peptide to an inhibitory one. Both peptides promoted the dissociation of a previously described complex formed between Rev and IN whose formation results in IN inactivation. INS and INS K188E penetrated HIV‐1‐infected cells and caused stimulation and inhibition of viral genome integration, respectively. Using cultured cells infected with a ΔRev HIV revealed that INS can directly activate the viral IN. These results suggest that the stimulatory effect of INS in wild‐type virus‐infected cells is due to a dual effect: it dissociates the inactive Rev‐IN complex and directly activates the free IN. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 740–751, 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.
Our previous studies of the potential utility of the CαDα stretch frequency, ν(CD), as a tool for determining conformation in peptide systems (Mirkin and Krimm, J Phys Chem A 2004, 108, 10923–10924; 2007, 111, 5300–5303) dealt with the spectroscopic characteristics of isolated alanine peptides with αR, β, and polyproline II structures. We have now extended these ab initio calculations to include various explicit‐water environments interacting with such conformers. We find that the structure‐discriminating feature of this technique is in fact enhanced as a result of the conformation‐specific interactions of the bonding waters, in part due to our finding (Mirkin and Krimm, J Phys Chem B 2008, 112, 15268) that Cα? Dα…O(water) hydrogen bonds can be present in addition to those expected between water and the CO and NH of the peptide groups. In fact, ν(CD) is hardly affected by the latter bonding but can be shifted by up to 70 cm?1 by the former hydrogen bonds. We also discuss the factors that will have to be considered in developing the molecular dynamics (MD) treatment needed to satisfactorily take account of the influence of outer water layers on the structure of the first‐layer water molecules that hydrogen bond to the peptide backbone. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 791–800, 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.
12.
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   

13.
It has been reported that the complexes formed by hemin and some G‐quadruplexes can be developed as a new class of DNAzyme with peroxidase activity. This kind of DNAzyme has received a great deal of attention. But to date, the actual G‐quadruplex structure that can provide hemin with enhanced peroxidase activity is in doubt. Herein, the G‐quadruplex structure of CatG4, a 21‐nucleotide DNA oligomer which was previously reported to bind hemin and the resulting complex exhibiting enhanced peroxidase activity, was characterized by fluorescence and circular dichroism measurements. The results suggest that the catalytically active form of CatG4 may be a unimolecular parallel quadruplex rather than a unimolecular chair‐type antiparallel quadruplex or a multistranded parallel quadruplex. In addition, the fluorescence analysis of labeled oligonucleotides may be developed as a supplementary tool for the study of DNA conformations. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 331–339, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

14.
A hemoprotein‐based supramolecular polymer that has a covalently linked heme moiety on the protein surface has been constructed based on interprotein heme–heme pocket interactions of the chemically modified apocytochrome b562 ( 1 ‐H63C). The thermodynamic properties of the polymer have been investigated by means of size exclusion chromatography, UV–vis spectroscopy, and circular dichroism spectroscopy. The results indicate that, as with other synthetic systems reported so far, the 1 ‐H63C hemoprotein assembly is thermodynamically controlled in aqueous solution: the degree of polymerization is dependent on the 1 ‐H63C concentration and is modulated by the addition of the end‐capping units, native heme, and/or apocytochrome b562 mutant (apoH63C). These properties suggest a potential use for the hemoprotein self‐assembly in preparation of stimuli‐responsive functional nanobiomaterials. © 2008 Wiley Periodicals, Inc. Biopolymers 91: 194–200, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

15.
Poly‐L ‐proline has been used as a model system for various purposes over a period of more than 60 years. Its relevance among the protein/peptide community stems from its use as a reference system for determining the conformational distributions of unfolded peptides and proteins, its use as a molecular ruler, and from the pivotal role of proline residues in conformational transitions and protein–protein interactions. While several studies indicate that polyproline can aggregate and precipitate in aqueous solution, a systematic study of the aggregation process is still outstanding. We found, by means of UV‐circular dichroism and IR measurements, that polyproline is predominately monomeric at room temperature at millimolar concentrations. Upon heating, the polypeptide stays in its monomeric state until the temperature reaches a threshold of ca. 60°C. At higher temperatures, the peptide aggregates as a film on the inside surface of the employed cuvette. The process proceeds on a time scale of 103 s and can best be described by a bi‐exponential relaxation function. The respective CD and IR spectra are qualitatively different from the canonical spectra of polyproline in aqueous solution, and are indicative of a highly packed state. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 451–457, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

16.
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   

17.
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   

18.
Structures of (Pro‐Pro‐Gly)4‐Xaa‐Yaa‐Gly‐(Pro‐Pro‐Gly)4 (ppg9‐XYG) where (Xaa, Yaa) = (Pro, Hyp), (Hyp, Pro) or (Hyp, Hyp) were analyzed at high resolution using synchrotron radiation. Molecular and crystal structures of these peptides are very similar to those of the (Pro‐Pro‐Gly)9 peptide. The results obtained in this study, together with those obtained from related compounds, indicated the puckering propensity of the Hyp in the X position: (1) Hyp(X) residues involved in the Hyp(X):Pro(Y) stacking pairs prefer the down‐puckering conformation, as in ppg9‐OPG, and ppg9‐OOG; (2) Hyp(X) residues involved in the Hyp(X):Hyp(Y) stacking pairs prefer the up‐puckering conformation if there is no specific reason to adopt the down‐puckering conformation. Water molecules in these peptide crystals are classified into two groups, the 1st and 2nd hydration waters. Water molecules in the 1st hydration group have direct hydrogen bonds with peptide oxygen atoms, whereas those in the 2nd hydration group do not. Compared with globular proteins, the number of water molecules in the 2nd hydration shell of the ppg9‐XYG peptides is very large, likely due to the unique rod‐like molecular structure of collagen model peptides. In the collagen helix, the amino acid residues in the X and Y positions must protrude outside of the triple helix, which forces even the hydrophobic side chains, such as Pro, to be exposed to the surrounding water molecules. Therefore, most of the waters in the 2nd hydration shell are covering hydrophobic Pro side chains by forming clathrate structures. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 361–372, 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   

19.
Among the pathological hallmarks of Alzheimer's disease (AD) is the deposition of amyloid‐β (Aβ) peptides, primarily Aβ (1–40) and Aβ (1–42), in the brain as senile plaques. A large body of evidence suggests that cognitive decline and dementia in AD patients arise from the formation of various aggregated forms of Aβ, including oligomers, protofibrils and fibrils. Hence, there is increasing interest in designing molecular agents that can impede the aggregation process and that can lead to the development of therapeutically viable compounds. Here, we demonstrate the ability of the specifically designed α,β‐dehydroalanine (ΔAla)‐containing peptides P1 (K‐L‐V‐F‐ΔA‐I‐ΔA) and P2 (K‐F‐ΔA‐ΔA‐ΔA‐F) to inhibit Aβ (1–42) aggregation. The mechanism of interaction of the two peptides with Aβ (1–42) seemed to be different and distinct. Overall, the data reveal a novel application of ΔAla‐containing peptides as tools to disrupt Aβ aggregation that may lead to the development of anti‐amyloid therapies not only for AD but also for many other protein misfolding diseases. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 456–465, 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.
A sequence derived from the epithelial receptor tyrosine kinase Ros (pY2267) represents a high‐affinity binding partner for protein tyrosine phosphatase SHP‐1 and was recently used as lead structure to analyze the recognition requirements for the enzyme's N‐SH2 domain. Here, we focused on a set of peptides comprising C‐terminally extended linear and conformationally constrained side chain‐bridged cyclic N‐SH2 ligands based on the consensus sequence LxpYhxh(h/b)(h/b) (x = any amino acid, h = hydrophobic, and b = basic residue). Furthermore, the bivalent peptides described were designed to modulate the activity of SHP‐1 through binding to both, the N‐SH2 domain as well as an independent binding site on the surface of the catalytic domain (PTP domain). Consistent with previous experimental findings, surface plasmon resonance experiments revealed dissociation constants of most compounds in the low micromolar range. One peptide, EGLNpYc[KVD]MFPAPEEE? NH2, displayed favorable binding affinity, but reduced ability to stimulate SHP‐1. Docking experiments revealed that the binding of this ligand occurs in binding mode I, recently described to lead to an inhibited activation of SHP‐1. In summary, results presented in this study suggest that inhibitory N‐SH2 ligands of SHP‐1 may be obtained by designing bivalent compounds that associate with the N‐SH2 domain and simultaneously occupy a specific binding site on the PTP domain. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 102–112, 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