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 78 毫秒
1.
Superior to linear peptides in biological activities, cyclic peptides are considered to have great potential as therapeutic agents. To identify cyclic‐peptide ligands for therapeutic targets, phage‐displayed peptide libraries in which cyclization is achieved by the covalent conjugation of cysteines have been widely used. To resolve drawbacks related to cysteine conjugation, we have invented a phage‐display technique in which its displayed peptides are cyclized through a proximity‐driven Michael addition reaction between a cysteine and an amber‐codon‐encoded N?‐acryloyl‐lysine (AcrK). Using a randomized 6‐mer library in which peptides were cyclized at two ends through a cysteine–AcrK linker, we demonstrated the successful selection of potent ligands for TEV protease and HDAC8. All selected cyclic peptide ligands showed 4‐ to 6‐fold stronger affinity to their protein targets than their linear counterparts. We believe this approach will find broad applications in drug discovery.  相似文献   

2.
A systematic comparison between the grafting‐to (convergent) and grafting‐from (divergent) synthetic routes leading to cyclic peptide–polymer conjugates is described. The reversible addition–fragmentation chain transfer (RAFT) process was used to control the polymerizations and the couplings between cyclic peptide and polymer or RAFT agent were performed using N‐hydroxysuccinimide (NHS) active ester ligation. The kinetics of polymerization and polymer conjugation to cyclic peptides were studied for both grafting‐to and grafting‐from synthetic routes, using N‐acryloyl morpholine as a model monomer. The cyclic peptide chain transfer agent was able to mediate polymerization as efficiently as a traditional RAFT agent, reaching high conversion in the same time scale while maintaining excellent control over the molecular weight distribution. The conjugation of polymers to cyclic peptides proceeded to high conversion, and the nature of the carbon at the α‐position to the NHS group was found to play a crucial role in the reaction kinetics. The study was extended to a wider range of monomers, including hydrophilic and temperature responsive acrylamides, hydrophilic and hydrophobic acrylates, and hydrophobic and pH responsive methacrylates. Both approaches lead to similar peptide–polymer conjugates in most cases, while some exceptions highlight the advantages of one or the other method, thereby demonstrating their complementarity. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 1003–1011  相似文献   

3.
Constraining small peptides into specific secondary structures has been a major challenge in peptide ligand design. So far, the major solution for decreasing the conformational flexibility in small peptides has been cyclization. An alternative is the use of topological templates, which are able to induce and/or stabilize peptide secondary structures by means of covalent attachment to the peptide. Herein a multicomponent strategy and structural analysis of a new type of peptidosteroid architecture having the steroid as N‐substituent of an internal amide bond is reported. The approach comprises the one‐pot conjugation of two peptide chains (or amino acid derivatives) to aminosteroids by means of the Ugi reaction to give a unique family of N‐steroidal peptides. The conjugation efficiency of a variety of peptide sequences and steroidal amines, as well as their consecutive head‐to‐tail cyclization to produce chimeric cyclopeptide–steroid conjugates, that is, macrocyclic lipopeptides, was assessed. Determination of the three‐dimensional structure of an acyclic N‐steroidal peptide in solution proved that the bulky, rigid steroidal template is capable of both increasing significantly the conformational rigidity, even in a peptide sequence as short as five amino acid residues, and inducing a β‐turn secondary structure even in the all‐strans isomer. This report provides the first evidence of the steroid skeleton as β‐turn inducer in linear peptide sequences.  相似文献   

4.
Biomolecules express exquisite properties that are required for molecular recognition and self‐assembly on the nanoscale. These smart capabilities have developed through evolution and such biomolecules operate based on smart functions in natural systems. Recently, these remarkable smart capabilities have been utilized in not only biologically related fields, but also in materials science and engineering. A peptide‐screening technology that uses phage‐display systems has been developed based on this natural smart evolution for the generation of new functional peptide bionanomaterials. We focused on peptides that specifically bound to synthetic polymers. These polymer‐binding peptides were screened by using a phage‐display peptide library to recognize nanostructures that were derived from polymeric structural features and were utilized for possible applications as new bionanomaterials. We also focused on self‐assembling peptides with β‐sheet structures that formed nanoscale, fibrous structures for applications in new bottom‐up nanomaterials. Moreover, nanofiber‐binding peptides were also screened to introduce the desired functionalities into nanofibers without the need for additional molecular design. Our approach to construct new bionanomaterials that employ peptides will open up excellent opportunities for the next generation of materials science and technology.  相似文献   

5.
Macroporous epoxy cryogels can be used as an alternative for classical matrices in affinity chromatography. Due to the structural properties of cryogels, with pores of up to 100 μm, crude samples can be processed at high speed without previous manipulations such as clarification or centrifugation. Also, we previously used a peptide‐expressing M13 bacteriophage as an affinity ligand. These ligands show high specificity toward the target to be purified. Combination of both, leads to a relative cost‐effective one‐step chromatographic set‐up delivering a high purity sample (>95%), however, so far with limited capacity. To increase the binding capacity of the affinity columns, we now inserted spacers between the chromatographic matrix and the phage ligand. Both linear spacers, di‐amino‐alkanes (C2–C10), and branched polyethyleneimine spacers with different molecular weights (800 Da–10 kDa) were analyzed. Two types of peptide expressing phage ligands, a linear 15‐mer and a cyclic 6‐mer, were used for screening. Up to a tenfold increase in binding capacity was observed depending on the combination of phage ligand and spacer type.  相似文献   

6.
The synthesis, optical resolution, determination of absolute configuration and conformational preference, and spectroscopic characteristics of terminally protected (blocked) derivatives and short peptides of 2‐amino‐1,2,3,6‐tetrahydro‐6‐oxocyclopenta[c]fluorene‐2‐carboxylic acid (FlAib), a novel, rigid, chiral, cyclized Cα,α‐disubstituted glycine are described.  相似文献   

7.
A four‐membered oxygen ring (oxetane) can be readily grafted into native peptides and proteins through site‐selective bis‐alkylation of cysteine residues present as disulfides under mild and biocompatible conditions. The selective installation of the oxetane graft enhances stability and activity, as demonstrated for a range of biologically relevant cyclic peptides, including somatostatin, proteins, and antibodies, such as a Fab arm of the antibody Herceptin and a designed antibody DesAb‐Aβ against the human Amyloid‐β peptide. Oxetane grafting of the genetically detoxified diphtheria toxin CRM197 improves significantly the immunogenicity of this protein in mice, which illustrates the general utility of this strategy to modulate the stability and biological activity of therapeutic proteins containing disulfides in their structures.  相似文献   

8.
Peptides are often ideal ligands for diagnostic molecular imaging due to their ease of synthesis and tuneable targeting properties. However, labelling unmodified peptides with 18F for positron emission tomography (PET) imaging presents a number of challenges. Here we show the combination of photoactivated sodium decatungstate and [18F]‐N‐fluorobenzenesulfonimide effects site‐selective 18F‐fluorination at the branched position in leucine residues in unprotected and unaltered peptides. This streamlined process provides a means to directly convert native peptides into PET imaging agents under mild aqueous conditions, enabling rapid discovery and development of peptide‐based molecular imaging tools.  相似文献   

9.
Peptides are often ideal ligands for diagnostic molecular imaging due to their ease of synthesis and tuneable targeting properties. However, labelling unmodified peptides with 18F for positron emission tomography (PET) imaging presents a number of challenges. Here we show the combination of photoactivated sodium decatungstate and [18F]‐N‐fluorobenzenesulfonimide effects site‐selective 18F‐fluorination at the branched position in leucine residues in unprotected and unaltered peptides. This streamlined process provides a means to directly convert native peptides into PET imaging agents under mild aqueous conditions, enabling rapid discovery and development of peptide‐based molecular imaging tools.  相似文献   

10.
Matrix‐assisted laser desorption/ionization in‐source decay (MALDI‐ISD) induces N–Cα bond cleavage via hydrogen transfer from the matrix to the peptide backbone, which produces a c′/z? fragment pair. Subsequently, the z? generates z′ and [z + matrix] fragments via further radical reactions because of the low stability of the z?. In the present study, we investigated MALDI‐ISD of a cyclic peptide. The N–Cα bond cleavage in the cyclic peptide by MALDI‐ISD produced the hydrogen‐abundant peptide radical [M + 2H]+? with a radical site on the α‐carbon atom, which then reacted with the matrix to give [M + 3H]+ and [M + H + matrix]+. For 1,5‐diaminonaphthalene (1,5‐DAN) adducts with z fragments, post‐source decay of [M + H + 1,5‐DAN]+ generated from the cyclic peptide showed predominant loss of an amino acid with 1,5‐DAN. Additionally, MALDI‐ISD with Fourier transform‐ion cyclotron resonance mass spectrometry allowed for the detection of both [M + 3H]+ and [M + H]+ with two 13C atoms. These results strongly suggested that [M + 3H]+ and [M + H + 1,5‐DAN]+ were formed by N–Cα bond cleavage with further radical reactions. As a consequence, the cleavage efficiency of the N–Cα bond during MALDI‐ISD could be estimated by the ratio of the intensity of [M + H]+ and [M + 3H]+ in the Fourier transform‐ion cyclotron resonance spectrum. Because the reduction efficiency of a matrix for the cyclic peptide cyclo(Arg‐Gly‐Asp‐D‐Phe‐Val) was correlated to its tendency to cleave the N–Cα bond in linear peptides, the present method could allow the evaluation of the efficiency of N–Cα bond cleavage for MALDI matrix development. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

11.
Cyclo‐β‐tetrapeptides are known to adopt a conformation with an intramolecular transannular hydrogen bond in solution. Analysis of this structure reveals that incorporation of a β2‐amino‐acid residue should lead to mimics of ‘α‐peptidic β‐turns’ (cf. A, B, C ). It is also known that short‐chain mixed β/α‐peptides with appropriate side chains can be used to mimic interactions between α‐peptidic hairpin turns and G protein‐coupled receptors. Based on these facts, we have now prepared a number of cyclic and open‐chain tetrapeptides, 7 – 20 , consisting of α‐, β2‐, and β3‐amino‐acid residues, which bear the side chains of Trp and Lys, and possess backbone configurations such that they should be capable of mimicking somatostatin in its affinity for the human SRIF receptors (hsst1–5). All peptides were prepared by solid‐phase coupling by the Fmoc strategy. For the cyclic peptides, the three‐dimensional orthogonal methodology (Scheme 3) was employed with best success. The new compounds were characterized by high‐resolution mass spectrometry, NMR and CD spectroscopy, and, in five cases, by a full NMR‐solution‐structure determination (in MeOH or H2O; Fig. 4). The affinities of the new compounds for the receptors hsst1–5 were determined by competition with [125I]LTT‐SRIF28 or [125I] [Tyr10]‐CST14. In Table 1, the data are listed, together with corresponding values of all β‐ and γ‐peptidic somatostatin/Sandostatin® mimics measured previously by our groups. Submicromolar affinities have been achieved for most of the human SRIF receptors hsst1–5. Especially high, specific binding affinities for receptor hsst4 (which is highly expressed in lung and brain tissue, although still of unknown function!) was observed with some of the β‐peptidic mimics. In view of the fact that numerous peptide‐activated G protein‐coupled receptors (GPCRs) recognize ligands with turn structure (Table 2), the results reported herein are relevant far beyond the realm of somatostatin: many other peptide GPCRs should be ‘reached’ with β‐ and γ‐peptidic mimics as well, and these compounds are proteolytically and metabolically stable, and do not need to be cell‐penetrating for this purpose (Fig. 5).  相似文献   

12.
Formylglycine‐generating enzymes are of increasing interest in the field of bioconjugation chemistry. They catalyze the site‐specific oxidation of a cysteine residue to the aldehyde‐containing amino acid Cα‐formylglycine (FGly). This non‐canonical residue can be generated within any desired target protein and can subsequently be used for bioorthogonal conjugation reactions. The prototypic formylglycine‐generating enzyme (FGE) and the iron‐sulfur protein AtsB display slight variations in their recognition sequences. We designed specific tags in peptides and proteins that were selectively converted by the different enzymes. Combination of the different tag motifs within a single peptide or recombinant protein enabled the independent and consecutive introduction of two formylglycine residues and the generation of heterobifunctionalized protein conjugates.  相似文献   

13.
《Electroanalysis》2005,17(23):2163-2169
A thiol‐specific electroactive cross‐linker, N‐(2‐ethyl‐ferrocene)maleimide (Fc‐Mi), has been used to tag surface‐confined peptides containing cysteine residues or oligodeoxynucleotides (ODNs) whose 3′ ends have been modified with thiol groups. The peptides studied herein include both the oxidized and reduced forms of glutathione and a hexapeptide. Cyclic voltammograms (CVs) of the Fc‐Mi groups attached to the surfaces were used to quantify the total number of cysteine residues that are tagged and/or can undergo facile electron transfer reactions with the underlying electrodes. A quartz crystal microbalance was used in conjunction with CV to estimate the total number of cysteine groups labeled by Fc‐Mi per peptide molecule. By comparing to mass spectrometric studies, it is confirmed that not all of the Fc‐Mi linked to the cysteine groups can participate in the electron transfer reactions. The methodology is further extended to the determination of ODN samples in a sandwich assay wherein the thiol linker on the 3′ end can be tagged with Fc‐Mi. The analytical performance was evaluated through determinations of a complementary ODN target and targets with varying numbers of mismatching bases. ODN samples as low as 10 fmol can be detected. Such a low detection level is remarkable considering that no signal amplification scheme is involved in the current method. The approach is shown to be sequence‐ and/or structure‐specific and does not require sophisticated instrumentation and complex experimental procedure.  相似文献   

14.
This work reports a novel chlorooxime mediated modification of native peptides and proteins under physiologic conditions. This method features fast reaction kinetics (apparent k2=306±4 M−1s−1 for GSH) and exquisite selectivity for cysteine residues. This cysteine conjugation reaction can be carried out with just single-digit micromolar concentrations of the labeling reagent. The conjugates show high stability towards acid, base, and external thiol nucleophiles. A nitrile oxide species generated in situ is likely involved as the key intermediate. Furthermore, a bis-chlorooxime reagent is synthesized to enable facile Cys-Cys stapling in native peptides and proteins. This highly efficient cysteine conjugation and stapling was further implemented on bacteriophage to construct chemically modified phage libraries.  相似文献   

15.
We report a novel conjugation of N‐terminal cysteines (NCys) that proceeds with fast kinetics and exquisite selectivity, thereby enabling facile modification of NCys‐bearing proteins in complex biological milieu. This new NCys conjugation proceeds via a thiazolidine boronate (TzB) intermediate that results from fast (k2: ≈5000 m ?1 s?1) and reversible conjugation of NCys with 2‐formylphenylboronic acid (FPBA). We designed a FPBA derivative that upon TzB formation elicits intramolecular acyl transfer to give N‐acyl thiazolidines. In contrast to the quick hydrolysis of TzB, the N‐acylated thiazolidines exhibit robust stability under physiologic conditions. The utility of the TzB‐mediated NCys conjugation is demonstrated by rapid and non‐disruptive labeling of two enzymes. Furthermore, applying this chemistry to bacteriophage allows facile chemical modification of phage libraries, which greatly expands the chemical space amenable to phage display.  相似文献   

16.
1H‐NMR was previously used to analyze the interaction between peptides (E3 and R826) selected by phage display to target apoptotic cells and phospholipidic models of these cells. In order to avoid the use of apoptotic cells and to obtain a fast evaluation of the efficiency of the potential MRI contrast agents obtained by grafting these peptides and their scramble analogs on a paramagnetic gadolinium complex, their proton relaxometric behavior was investigated in the presence of micelles mimicking healthy and apoptotic cells. Their preferential interaction with 1,2‐dipalmitoyl‐sn‐glycero‐3‐phospho‐l ‐serine micelles mimicking apoptotic cells as compared with 1,2‐dipalmitoyl‐sn‐glycero‐3‐phosphocholine micelles modeling healthy cells was shown by nuclear magnetic relaxation dispersion profiles and the enhancement of the transverse proton relaxation rates at 60 MHz. The association constant values confirm the stronger interaction of the selected conjugated peptides (Ka Gd‐PMN‐E3(gadolinium 2,2′,2′′,2′′′‐[((4‐carboxy)pyridine‐2,6‐diyl)bis(methylenenitrilo)]‐tetrakis acetate) grafted with E3 peptide): 2.43 104 m ?1; Ka Gd‐DTPA‐R826(gadolinium ((1‐p‐isothiocyanatobenzyl)‐diethylenetriaminepentaacetate) grafted with R826 peptide): 2.91 104 m ?1) as compared with their conjugated scrambles (Ka Gd‐PMN‐E3sc(gadolinium 2,2′,2′′,2′′′‐[((4‐carboxy)pyridine‐2,6‐diyl)bis(methylenenitrilo)]‐tetrakis acetate) grafted with E3 scramble peptide): 0.18 104 m ?1; Ka Gd‐DTPA‐R826sc(gadolinium ((1‐p‐isothiocyanatobenzyl)‐diethylenetriaminepentaacetate) grafted with R826 scramble peptide): 0.32 104 m ?1) even if the conjugation of E3 and R826 seems to decrease their interaction. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

17.
Previous studies into the dissociation of [CuII(dien)peptide] . 2+ ions (dien = diethylenetriamine) have shown that NH‐containing auxiliary ligands do not favor the formation of [peptide] . + species; instead, they promote proton‐transfer reactions, especially for peptides containing basic amino residues. Formation of radical cationic tripeptides of the form GGX . + [GGX = glycylglycyl(residue X)] becomes feasible upon substituting the open‐chain tridentate ligand dien with its analogous cyclic ligand, 1,4,7‐triazacyclononane (9‐aneN3); i.e., from [CuII(9‐aneN3)GGX] . 2+ ions. Similar enhancements occur when using 1,4,7,10‐tetraoxacyclododecane (12‐crown‐4) in place of its open‐chain analog, 2,5,8,11‐tetraoxadecane (triglyme). We have demonstrated that a sterically encumbered auxiliary macrocyclic ligand within [CuII(L)GGX] . 2+ complex ions [where L = 9‐aneN3 or 12‐crown‐4] facilitates the formation of radical cationic peptides through gas‐phase fragmentation. We verified our experimental observations by examining the reactivities of a series of 19 tripeptides of the type GGX that differ only in the identity of their C‐terminal residue. The energy of the electron‐transfer reaction correlates well with the bond‐dissociation energy of the peptide–Cu(II) interaction; the presence of a constrained macrocyclic ligand weakens metal–peptide chelation through steric repulsion between the ligand and the peptide, and this situation may lead to more favorable radical cationic peptide formation. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

18.
Poly(ethylene imine) (PEI), a highly cationic polymer, is being used for deoxyribonucleic acid (DNA) complexation and delivery into cells. To enhance the cellular uptake of polymer/DNA complexes, arginine–glycine–aspartic acid (RGD) peptides have been conjugated to PEI with N‐succinimidyl 3‐(2‐pyridyldithio)propionate (SPDP). This coupling scheme creates a disulfide‐linked conjugate, the stability of which in the presence of thiols is uncertain. We have investigated the conjugation of an RGD peptide, glycine–arginine–glycine–aspartic acid–serine–proline–cysteine (GRGDSPC), to PEI with SPDP and subsequently assessed the stability of the conjugates in the presence of two thiol compounds, mercaptoethanol and cysteine. SPDP effectively controls the extent of GRGDSPC substitution on PEI. The conjugates, however, are readily cleaved in the presence of the thiols; the cleavage is rapid (~50% cleavage in 2–4 h) and inversely related to the degree of peptide substitution on the polymers. The peptide coupling is stable in the absence of thiols, and its cleavage is strongly dependent on the pH of the medium but not on the ionic strength of the medium. We conclude that RGD peptides coupled to PEI are labile in the presence of physiological concentrations of thiols, and this should be taken into account when such polymer–peptide conjugates are used for DNA delivery. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 6143–6156, 2004  相似文献   

19.
The low response rate and adaptive resistance of PD‐1/PD‐L1 blockade demands the studies on novel therapeutic targets for cancer immunotherapy. We discovered that a novel immune checkpoint TIGIT expressed higher than PD‐1 in many tumors especially anti‐PD‐1 resistant tumors. Here, mirror‐image phage display bio‐panning was performed using the d ‐enantiomer of TIGIT synthesized by hydrazide‐based native chemical ligation. d ‐peptide DTBP‐3 was identified, which could occupy the binding interface and effectively block the interaction of TIGIT with its ligand PVR. DTBP‐3 showed proteolytic resistance, tumor tissue penetrating ability, and significant tumor suppressing effects in a CD8+ T cell dependent manner. More importantly, DTBP‐3 could inhibit tumor growth and metastasis in anti‐PD‐1 resistant tumor model. This is the first d ‐peptide targeting TIGIT, which could serve as a potential candidate for cancer immunotherapy.  相似文献   

20.
We report a general and operationally simple method for the solid phase synthesis of α‐ketoamide peptides using standard Fmoc solid phase peptide synthesis. The method delivers deprotected peptide α‐ketoamides directly upon resin cleavage without any additional steps, and tolerates all side chain functional groups. A small collection of C‐terminal and internal α‐ketoamide peptides – including two reported protease inhibitors (HCV and SUB1) – were prepared in good yields. In addition, we demonstrate that our method serves as versatile platform for the convenient preparation of cyclic α‐ketoamide peptides, photocagged peptide α‐ketoamides, and fluorescently labeled peptides.  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