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.
Overexpression of gonadotropin‐releasing hormone (GnRH) receptor in many tumors but not in normal tissues makes it possible to use GnRH analogs as targeting peptides for selective delivery of cytotoxic agents, which may help to enhance the uptake of anticancer drugs by cancer cells and reduce toxicity to normal cells. The GnRH analogs [d ‐Cys6, desGly10, Pro9‐NH2]‐GnRH, [d ‐Cys6, desGly10, Pro9‐NHEt]‐GnRH, and [d ‐Cys6, α‐aza‐Gly10‐NH2]‐GnRH were conjugated with doxorubicin (Dox), respectively, through N‐succinimidyl‐3‐maleimidopropionate as a linker to afford three new GnRH‐Dox conjugates. The metabolic stability of these conjugates in human serum was determined by RP‐HPLC. The antiproliferative activity of the conjugates was examined in GnRH receptor‐positive MCF‐7 human breast cancer cell line by MTT assay. The three GnRH‐Dox conjugates showed improved metabolic stability in human serum in comparison with AN‐152. The antiproliferative effect of conjugate II ([d ‐Cys6, desGly10, Pro9‐NHEt]‐GnRH‐Dox) on MCF‐7 cells was higher than that of conjugate I ([d ‐Cys6, desGly10, Pro9‐NH2]‐GnRH‐Dox) and conjugate III ([d ‐Cys6, α‐aza‐Gly10‐NH2]‐GnRH‐Dox), and the cytotoxicity of conjugate II against GnRH receptor‐negative 3T3 mouse embryo fibroblast cells was decreased in comparison with free Dox. GnRH receptor inhibition test suggested that the antiproliferative activity of conjugate II might be due to the cellular uptake mediated by the targeting binding of [d ‐Cys6‐des‐Gly10‐Pro9‐NHEt]‐GnRH to GnRH receptors. Our study indicates that targeting delivery of conjugate II mediated by [d ‐Cys6‐des‐Gly10‐Pro9‐NHEt]‐GnRH is a promising strategy for chemotherapy of tumors that overexpress GnRH receptors.  相似文献   

2.
We report the conformational analysis by 1H‐NMR in DMSO and computer simulations involving distance geometry and molecular dynamics simulations of peptoid analogs of the cyclic hexapeptide c‐[Phe11‐Pro6‐Phe7‐d ‐Trp8‐Lys9‐Thr10] L‐363,301 (the numbering refers to the positions in native somatostatin). The compounds c‐[Phe11‐Nphe6‐Nal7‐d ‐Trp8‐Lys9‐Thr10] ( Nphe 6 ‐ Nal 7 analog 1 ), c‐[Nal11‐Nphe6‐Phe7‐d ‐Trp8‐Lys9‐Thr10] ( Nal 11 ‐ Nphe 6 analog 2 ) and c‐[Phe11‐Nnal6‐Phe7‐d ‐Trp8‐Lys9‐Thr10] ( Nnal 6 analog 3 ), where Nphe denotes N‐benzylglycine and Nnal denotes N‐(1‐naphthylmethyl)glycine, are subjected to SAR studies in order to investigate the influence of the bulky naphthyl aromatic ring on the conformation. The Nal 11 ‐ Nphe 6 and Nphe 6 ‐Nal 7 analogs exhibit potent binding to the hsst2, hsst3 and hsst5 receptors, whereas the Nnal 6 analog has decreased binding affinity to all receptors but is more selective towards the hsst2 than the other two analogs and L‐363,301. The conformational search employing distance geometry, energy minimization and molecular dynamic simulations gives insight into the conformational flexibility of these analogs. The molecules adopt both cis and trans orientations of the peptide bond between residues 11 and 6. The cis isomers of these analogs adopt type II′ β‐turns with d ‐Trp in the i+1 position and type VIa β‐turns with the cis peptide bond between residues 6 and 11. The results of free and distance restrained molecular dynamics simulations at 300 K indicate that the Nphe 6 ‐Nal 7 and Nal 11 ‐Nphe 6 compounds adopt a preferred backbone conformation which can be described as ‘folded’ about residues 7 and 10. The Nnal 6 analog, which binds less effectively to the hsst receptors, has a more flexible backbone structure than the Nal 11 ‐Nphe 6 and Nphe 6 ‐Nal 7 analogs and prefers a ‘flat’ structure with regard to the orientations about Phe7 and Thr10 during molecular dynamics simulations. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
The aim of this work was to examine the bioactivity and the conformational behavior of some gomesin (Gm) analogues in different environments that mimic the biological membrane/water interface. Thus, manual peptide synthesis was performed by the solid-phase method, antimicrobial activity was evaluated by a liquid growth inhibition assay, and conformational studies were performed making use of several spectroscopic techniques: CD, fluorescence and EPR. [TOAC1]-Gm; [TOAC1, Ser2,6,11,15]-Gm; [Trp7]-Gm; [Ser2,6,11,15, Trp7]-Gm; [Trp9]-Gm; and [Ser2,6,11,15, Trp9]-Gm were synthesized and tested. The results indicated that incorporation of TOAC or Trp caused no significant reduction of antimicrobial activity; the cyclic analogues presented a β-hairpin conformation similar to that of Gm. All analogues interacted with negatively charged SDS both above and below the detergent's critical micellar concentration (cmc). In contrast, while Gm and [TOAC1]-Gm required higher LPC concentrations to bind to micelles of this zwitterionic detergent, the cyclic Trp derivatives and the linear derivatives did not seem to interact with this membrane-mimetic system. These data corroborate previous results that suggest that electrostatic interactions with the lipid bilayer of microorganisms play an important role in the mechanism of action of gomesin. Moreover, the results show that hydrophobic interactions also contribute to membrane binding of this antimicrobial peptide.  相似文献   

4.
Vasopressin and nonmammalian hormone vasotocin are known to increase the water permeability of mammalian collecting ducts, frog skin and the urinary bladder. Neurohypophysial nonapeptides have also been shown to interfere with the regulation of renal ion transport. The subject of this study was a search for vasopressin and vasotocin analogues with selective effects on renal water, sodium and potassium excretion. During this study, we synthesised the following peptides: 13 vasotocin analogues modified at positions 4 (Thr or Arg), 7 (Gly or Leu) and 8 (d ‐Arg, Lys or Glu); 4 vasopressin analogues modified at positions 4 and 8; and 9 peptides shortened or extended at the C‐terminal or with substitutions for Gly‐NH2. Most of these peptides had mercaptopropionic acid (Mpa) instead of Cys in position 1. The effects of these nonapeptides on renal water, sodium and potassium transport were evaluated in in vivo experiments using Wistar rats. Some nonapeptides possessed antidiuretic, natriuretic and kaliuretic activities ([Mpa1]‐arginine vasotocin, [Mpa1, homoArg8]‐vasotocin, [Mpa1, Thr4]‐arginine vasotocin and [Mpa1, Arg4]‐arginine vasopressin). Substitutions at positions 4 and 8 increased the selectivity of peptide actions. The antidiuretic [d ‐Arg8]‐vasotocin analogues had no effects on sodium excretion. [Mpa1, Arg4]‐arginine vasotocin was antidiuretic and kaliuretic but not natriuretic. [Mpa1, Glu8]‐oxytocin had weak natriuretic activity without any effects on water and potassium transport. In accordance with the data obtained, synthesised vasotocin analogues could be good candidates for pharmaceuticals selectively regulating renal sodium and potassium transport, which is of clinical importance. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
Many reports have shown that antimicrobial peptides exhibit anticancer abilities. Gomesin (Gm) exhibits potent cytotoxic activity against cancer cells by a membrane pore formation induced after well-orchestrated intracellular mechanisms. In this report, the replacements of the Cys by Ser or Thr, and the use D-amino acids in the Gm structure were done to investigate the importance of the resistance to degradation of the molecule with its cytotoxicity. [Thr2,6,11,15]-Gm, and [Ser2,6,11,15]-Gm exhibits low cytotoxicity, and low resistance to degradation, and after 24 h are present in localized area near to the membrane. Conversely, the use of D-amino acids in the analogue [D-Thr2,6,11,15]-D-Gm confers resistance to degradation, increases its potency, and maintained this peptide spread in the cytosol similarly to what happens with Gm. Replacements of Cys by Thr and Gln by L- or D-Pro ([D-Thr2,6,11,15, Pro9]-D-Gm, and [Thr2,6,11,15, D-Pro9]-Gm), which induced a similar β-hairpin conformation, also increase their resistance to degradation, and cytotoxicity, but after 24 h they are not present spread in the cytosol, exhibiting lower cytotoxicity in comparison to Gm. Additionally, chloroquine, a lysosomal enzyme inhibitor potentiated the effect of the peptides. Furthermore, the binding and internalization of peptides was determined, but a direct correlation among these factors was not observed. However, cholesterol ablation, which increase fluidity of cellular membrane, also increase cytotoxicity and internalization of peptides. β-hairpin spatial conformation, and intracellular localization/target, and the capability of entry are important properties of gomesin cytotoxicity.  相似文献   

6.
The multiphosphorylated tryptic peptide αs1‐casein(59–79) has been shown to be antigenic with anti‐casein antibodies. In an approach to determine the amino acyl residues critical for antibody binding we undertook an epitope analysis of the peptide using overlapping synthetic peptides. With αs1‐casein(59–79) as the adsorbed antigen in a competitive ELISA only two of five overlapping synthetic peptides at 1 mM significantly inhibited binding of the anti‐casein antibodies. Peptides Glu‐Ser(P)‐Ile‐Ser(P)‐Ser(P)‐Ser(P)‐Glu‐Glu and Ile‐Val‐Pro‐Asn‐Ser(P)‐Val‐Glu‐Glu inhibited antibody binding by 20.0±3.6% and 60.3±7.9%, respectively. The epitope of Glu63‐Ser(P)‐Ile‐Ser(P)‐Ser(P)‐Ser(P)‐Glu‐Glu70 was further localised to the phosphoseryl cluster as the peptide Ser(P)‐Ser(P)‐Ser(P) significantly inhibited binding of the anti‐casein antibodies to αs1‐casein(59–79) by 29.5±7.4%. Substitution of Ser(P)75 with Ser75 in the second inhibitory peptide Ile‐Val‐Pro‐Asn‐Ser(P)75‐Val‐Glu‐Glu also abolished inhibition of antibody binding to αs1‐casein (59–79) demonstrating that Ser(P)75 is also a critical residue for recognition by the antibodies. These data show that the phosphorylated residues in the cluster sequence ‐Ser(P)66‐Ser(P)‐Ser(P)68 and in the sequence ‐Pro73‐Asn‐Ser(P)‐Val‐Glu77‐ are critical for antibody binding to αs1‐casein(59–79) and further demonstrate that a highly phosphorylated segment of a protein can be antigenic. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
The natural product cyclic peptide stylissatin A ( 1a ) was reported to inhibit nitric oxide production in LPS‐stimulated murine macrophage RAW 264.7 cells. In the current study, solid‐phase total synthesis of stylissatin A was performed by using a safety‐catch linker and yielded the peptide with a trans‐Phe7‐Pro6 linkage, whereas the natural product is the cis rotamer at this position as evidenced by a marked difference in NMR chemical shifts. In order to preclude the possibility of 1b being an epimer of the natural product, we repeated the synthesis using d ‐allo‐Ile in place of l ‐Ile and a different site for macrocyclization. The resulting product (d ‐allo‐Ile2)‐stylissatin A ( 1c ) was also found to have the trans‐Phe7‐Pro6 peptide conformations like rotamer 1b . Applying the second route to the synthesis of stylissatin A itself, we obtained stylissatin A natural rotamer 1a accompanied by rotamer 1b as the major product. Rotamers 1a , 1b , and the epimer 1c were separable by HPLC, and 1a was found to match the natural product in structure and biological activity. Six related analogs 2–7 of stylissatin A were synthesized on Wang resin and characterized by spectral analysis. The natural product ( 1a ), the rotamer ( 1b ), and (d ‐allo‐Ile2)‐stylissatin A ( 1c ) exhibited significant inhibition of NO.. Further investigations were focused on 1b , which also inhibited proliferation of T‐cells and inflammatory cytokine IL‐2 production. The analogs 2–7 weakly inhibited NO. production, but strongly inhibited IL‐2 cytokine production compared with synthetic peptide 1b . All analogs inhibited the proliferation of T‐cells, with analog 7 having the strongest effect. In the analogs, the Pro6 residue was replaced by Glu/Ala, and the SAR indicates that the nature of this residue plays a role in the biological function of these peptides. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The anti‐plasmodial activity of conformationally restricted analogs of angiotensin II against Plasmodium gallinaceum has been described. To observe activity against another Plasmodium species, invasion of red blood cells by Plasmodium falciparum was analyzed. Analogs restricted with lactam or disulfide bridges were synthesized to determine their effects and constraints in the peptide–parasite interaction. The analogs were synthesized using tert‐butoxycarbonyl and fluoromethoxycarbonyl solid phase methods, purified by liquid chromatography, and characterized by mass spectrometry. Results indicated that the lactam bridge restricted analogs 1 (Glu‐Asp‐Arg‐Orn ‐Val‐Tyr‐Ile‐His‐Pro‐Phe) and 3 (Asp‐Glu‐Arg‐Val‐Orn ‐Tyr‐Ile‐His‐Pro‐Phe) showed activity toward inhibition of ring formation stage of P. falciparum erythrocytic cycle, preventing invasion in about 40% of the erythrocytes. The disulfide‐bridged analog 10 (Cys‐Asp‐Arg‐Cys ‐Val‐Tyr‐Ile‐His‐Pro‐Phe) was less effective yet significant, showing a 25% decrease in infection of new erythrocytes. In all cases, the peptides presented no pressor activity, and hydrophobic interactions between the aromatic and alkyl amino acid side chains were preserved, a factor proven important in efficacy against P. gallinaceum. In contrast, hydrophilic interactions between the Asp1 carboxyl and Arg2 guanidyl groups proved not to be as important as they were in the case of P. gallinaceum, while interactions between the Arg2 guanidyl and Tyr4 hydroxyl groups were not important in either case. The β‐turn conformation was predominant in all of the active peptides, proving importance in anti‐plasmodial activity. This approach provides insight for understanding the importance of each amino acid residue on the native angiotensin II structure and a new direction for the design of potential chemotherapeutic agents. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Anoplin is a short natural cationic antimicrobial peptide which is derived from the venom sac of the solitary wasp, Anoplius samariensis. Due to its short sequence G1LLKR5IKT8LL‐NH2, it is ideal for research tests. In this study, novel analogs of anoplin were prepared and examined for their antimicrobial, hemolytic activity, and proteolytic stability. Specific substitutions were introduced in amino acids Gly1, Arg5, and Thr8 and lipophilic groups with different lengths in the N‐terminus in order to investigate how these modifications affect their antimicrobial activity. These cationic analogs exhibited higher antimicrobial activity than the native peptide; they are also nontoxic at their minimum inhibitory concentration (MIC) values and resistant to enzymatic degradation. The substituted peptide GLLKF5IKK8LL‐NH2 exhibited high activity against Gram‐negative bacterium Zymomonas mobilis (MIC = 7 µg/ml), and the insertion of octanoic, decanoic, and dodecanoic acid residues in its N‐terminus increased the antimicrobial activity against Gram‐positive and Gram‐negative bacteria (MIC = 5 µg/ml). The conformational characteristics of the peptide analogs were studied by circular dichroism. Structure activity studies revealed that the substitution of specific amino acids and the incorporation of lipophilic groups enhanced the amphipathic α‐helical conformation inducing better antimicrobial effects. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
The in vivo and in vitro potency of native and modified forms of gonadotropin releasing hormone (GnRH) to release luteotropic hormone (LH) was studied in sea bass Dicentrarchus labrax in particular the hypothalamic fish‐specific sea bream GnRH form (sbGnRH) and the general mesoencephalic form chicken GnRH‐II (cGnRH‐II). The potencies of the natives and their analogs (GnRHas) were referred to that of [D‐Ala6, Pro9Net]‐mGnRHa (LHRHa) at equivalent doses. Analogs of the native peptides [D‐Arg6, Pro9Net]‐cGnRH‐II, [D‐Ala6, Pro9Net]‐cGnRH‐II, [D‐Trp6, Pro9Net]‐sbGnRH and [D‐Ala6, Pro9Net]‐sbGnRH were effective in inducing in vivo LH release (at 15 µg kg?1 body mass), exhibiting longer lasting activity than their corresponding native forms. Injection of sbGnRH and cGnRH‐II provoked a small but significant peak of circulating LH at 1·5 h after treatment (a.t.) decreasing down to basal levels at 4 h a.t. [D‐Arg6, Pro9Net]‐cGnRH‐II, [D‐Ala6, Pro9Net]‐cGnRH‐II and [D‐Ala6, Pro9Net]‐mGnRHa evoked a higher and a more sustained elevation of LH, peaking at 12 h a.t. and returning to basal levels between 48 and 72 h a.t. [D‐Trp6, Pro9Net]‐sbGnRH and [D‐Ala6, Pro9Net]‐sbGnRH also induced a significant surge of LH in plasma at 4 h a.t. turning to the basal levels at 24 h a.t. These rises, however, were of less amplitude and duration than the observed after treatment with cGnRH‐II analogs and [D‐Ala6, Pro9Net]‐mGnRHa. The in vitro stimulation of dispersed pituitary cells with the different native and modified forms of GnRH resulted in a dose‐dependent increase in the quantity of LH released at 24 h a.t. [D‐Arg6, Pro9Net]‐cGnRH‐II and [D‐Ala6, Pro9Net]‐cGnRH‐II induced the highest response of LH in vitro release followed by salmon GnRH (sGnRH), [D‐Ala6, Pro9Net]‐mGnRHa and [D‐Trp6, Pro9Net]‐sbGnRH. The lowest activity was exhibited by sbGnRH. Collectively, the in vitro biological activity (compared by their EC50) can be ordered as follows: [D‐Arg6, Pro9Net]‐cGnRH‐II > [D‐Ala6, Pro9Net]‐cGnRH‐II > sGnRH > [D‐Ala6, Pro9Net]‐mGnRHa > [D‐Trp6, Pro9Net]‐sbGnRH > [D‐Ala6, Pro9Net]‐sbGnRH > cGnRH‐II > sbGnRH.  相似文献   

11.
Abstract: In [3H]myristic acid-prelabeled Chinese hamster ovary cells stably expressing the rat NK1 tachykinin receptor, the selective NK1 agonist [Pro9]substance P ([Pro9]SP) time and concentration dependently stimulated the formation of [3H]phosphatidylethanol in the presence of ethanol. This [Pro9]SP-induced activation of phospholipase D (PLD) was blocked by NK1 receptor antagonists and poorly or not mimicked by NK2 and NK3 agonists, respectively. In confirmation of previous observations, [Pro9]SP also stimulated the hydrolysis of phosphoinositides, the release of arachidonic acid, and the formation of cyclic AMP (cAMP). All these [Pro9]SP-evoked responses could be mimicked by aluminum fluoride, but they remained unaffected in cells pretreated with pertussis toxin, suggesting that a Gi/Go protein is not involved in these different signaling pathways. The activation of PLD by [Pro9]SP was sensitive to external calcium and required an active protein kinase C because the inhibition of this kinase (Ro 31-8220) or its down-regulation (long-term treatment with a phorbol ester) abolished the response. In contrast, a cAMP-dependent process was not involved in the activation of PLD because the [Pro9]SP-evoked response was neither affected by Rp-8-bromoadenosine 3′,5′-cyclic monophosphorothioate nor mimicked by cAMP-generating compounds (cholera toxin or forskolin) or by 8-bromo-cyclic AMP. A functional coupling of NK1 receptors to PLD was also demonstrated in the human astrocytoma cell line U 373 MG stimulated by SP or [Pro9]SP. These results suggest that PLD activation could be an additional signaling pathway involved in the mechanism of action of SP in target cells expressing NK1 receptors.  相似文献   

12.
In contrast with the common belief that all the amino acid residues in higher organisms are l ‐forms, d ‐amino acid residues have been recently detected in various aging tissues. Aspartic acid (Asp) residues are known to be the most prone to stereoinvert via cyclic imide intermediate. Although the glutamic acid (Glu) is similar in chemical structure to Asp, little has been reported to detect d ‐Glu residues in human proteins. In this study, we investigated the mechanism of the Glu‐residue stereoinversion catalyzed by water molecules using B3LYP/6‐31+G(d,p) density functional theory calculations. We propose that the Glu‐residue stereoinversion proceeds via a cyclic imide intermediate, i.e., glutarimide (GI). All calculations were performed by using a model compound in which a Glu residue was capped with acetyl and methylamino groups on the N‐ and C‐termini, respectively. We found that two water molecules catalyze the three steps involved in the GI formation: iminolization, cyclization, and dehydration. The activation energy required for the Glu residue to form a GI intermediate was estimated to be 32.3 kcal mol?1, which was higher than that of the experimental Asp‐residue stereoinversion. This calculation result suggests that the Glu‐residue stereoinversion is not favored under the physiological condition.  相似文献   

13.
Saha I  Shamala N 《Biopolymers》2012,97(1):54-64
The covalent linkage between the side‐chain and the backbone nitrogen atom of proline leads to the formation of the five‐membered pyrrolidine ring and hence restriction of the backbone torsional angle ? to values of ?60 °± 30° for the L ‐proline. Diproline segments constitute a chain fragment with considerably reduced conformational choices. In the current study, the conformational states for the diproline segment ( L Pro‐ L Pro) found in proteins has been investigated with an emphasis on the cis and trans states for the Pro‐Pro peptide bond. The occurrence of diproline segments in turns and other secondary structures has been studied and compared to that of Xaa‐Pro‐Yaa segments in proteins which gives us a better understanding on the restriction imposed on other residues by the diproline segment and the single proline residue. The study indicates that PII–PII and PII–α are the most favorable conformational states for the diproline segment. The analysis on Xaa‐Pro‐Yaa sequences reveals that the Xaa‐Pro peptide bond exists preferably as the trans conformer rather than the cis conformer. The present study may lead to a better understanding of the behavior of proline occurring in diproline segments which can facilitate various designed diproline‐based synthetic templates for biological and structural studies. © 2011 Wiley Periodicals, Inc. Biopolymers 97: 54–64, 2012.  相似文献   

14.
Extensive conformational analysis of a series of β‐alkyl substituted cyclopeptides—cyclo(Pro1–Xaa2–Nle3–Ala4–Nle5–Pro6–Xaa7–Nle8–Ala9–Nle10) and cyclo[Pro1–Xaa2–Nle3–(Cys4– Nle5–Pro6–Xaa7–Nle8–Cys9)–Nle10] as well as their corresponding unsubstituted core structures cyclo(Pro1–Xaa2–Ala3–Ala4–Ala5–Pro6–Xaa7–Ala8–Ala9–Ala10) and cyclo(Pro1–Xaa2–Ala3–Cys4– Ala5–Pro6–Xaa7–Ala8–Cys9–Ala10) has been performed employing both the ECEPP/2 and the MAB force fields (Xaa = Gly, L ‐Ala, D ‐Ala, Aib, and D ‐Pro). Results show that (a) possible three‐dimensional structures of the cyclo(Pro1–Gly2–Lys3–Ala4–Lys5–Pro6–Gly7–Lys8–Ala9–Lys10) molecule are not limited to a single extended “rectangular” conformation with all Lys side chains oriented at the same side of the molecule; (b) conformational equilibrium in monocyclic analogues obtained by replacements of conformationally flexible Gly residues for L ‐Ala, D ‐Ala, Aib, or D ‐Pro is not significantly shifted towards the target “rectangular” conformational type; and (c) introduction of disulfide bridges between positions 4 and 9 is a very powerful way to stabilize the target conformations in the resulting bicyclic molecules. These findings form the basis for further design of rigidified regioselectively addressable functionalized templates with many application areas ranging from biostructural to diagnostic purposes. © 1999 John Wiley & Sons, Inc. Biopoly 50: 361–372, 1999  相似文献   

15.
The crucial step of folding of recombinant proteins presents serious challenges to obtaining the native structure. This problem is exemplified by insulin‐like growth factor (IGF)‐I which when refolded in vitro produces the native three‐disulfide structure, an alternative structure with mispaired disulfide bonds and other isomeric forms. To investigate this phenomenon we have examined the refolding properties of an analog of IGF‐I which contains a 13‐amino acid N‐terminal extension and a charge mutation at position 3 (Long‐ [Arg3]IGF‐I). Unlike IGF‐I, which yields 45% of the native structure and 24% of the alternative structure when refolded in vitro, Long‐[Arg3]GF‐I yields 85% and 10% of these respective forms. To investigate the interactions that affect the refolding of Long‐[Arg3]IGF‐I and IGF‐I, we acid‐trapped folding intermediates and products for inclusion in a kinetic analysis of refolding. In addition to non‐native intermediates, three native‐like intermediates were identified, that appear to have a major role in the in vitro refolding pathway of Long‐[Arg3]IGF‐I; a single‐disulfide Cys18–Cys61 intermediate, an intermediate with Cys18–Cys61 and Cys6–Cys48 disulfide bonds and another with Cys18–Cys61 and Cys47–Cys52 disulfide bonds. Furthermore, from our kinetic analysis we propose that the Cys18‐Cys61, Cys6‐Cys48 intermediate forms the native structure, not by the direct formation of the last (Cys47‐Cys52) disulfide bond, but by rearrangement via the Cys18–Cys61 intermediate and a productive Cys18–Cys61, Cys47–Cys52 intermediate. In this pathway, the last disulfide bond to form involves Cys6 and Cys48. Finally, we apply this pathway to IGF‐I and conclude that the divergence in the in vitro folding pathway of IGF‐I is caused by non‐native interactions involving Glu3 that stabilize the alternative structure. © 1999 John Wiley & Sons, Inc. Biotechnol Bioeng 62: 693–703, 1999.  相似文献   

16.
Cyclization of bioactive peptides, utilizing functional groups serving as natural pharmacophors, is often accompanied with loss of activity. The backbone cyclization approach was developed to overcome this limitation and enhance pharmacological properties. Backbone cyclic peptides are prepared by the incorporation of special building units, capable of forming amide, disulfide and coordinative bonds. Urea bridge is often used for the preparation of cyclic peptides by connecting two amine functionalized side chains. Here we present urea backbone cyclization as an additional method for the preparation of backbone cyclic peptide libraries. A straightforward method for the synthesis of crystalline Fmoc‐Nα [ω‐amino(Alloc)‐alkyl] glycine building units is presented. A set of urea backbone cyclic Glycogen Synthase Kinase 3 analogs was prepared and assessed for protein kinase B inhibition as anticancer leads. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
C‐2 dimethylated/unmethylated thiazolidine‐4‐carboxylic acid and C‐2 dimethylated oxazolidine‐4‐carboxylic acid were introduced into the insect kinin core pentapeptide in place of Pro3, yielding three new analogues. NMR analysis revealed that the peptide bond of Phe2‐pseudoproline (ΨPro)3 is practically 100% in cis conformation in the case of dimethylated pseudoproline‐containing analogues, about 50% cis for the thiazolidine‐4‐carboxylic acid analogue and about 33% cis for the parent Pro3 peptide. The diuretic activities are consistent with the population of cis conformation of the Phe2‐ΨPro3/Pro3 peptide bonds, and the results confirm a cis Phe‐Pro bond as bioactive conformation. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
Syntheses are described of the nociceptin (1–13) amide [NC(1–13)-NH2] and of several analogues in which either one or both the phenylalanine residues (positions 1 and 4), the arginine residues (positions 8 and 12) and the alanine residues (positions 7 and 11) have been replaced by N-benzyl-glycine, N-(3-guanidino-propyl)-glycine and β-alanine, respectively. The preparation is also described of NC(1–13)-NH2 analogues in which either galactose or N-acetyl-galactosamine are β-O-glycosidically linked to Thr5 and/or to Ser10. Preliminary pharmacological experiments on mouse vas deferens preparations showed that Phe4, Thr5, Ala7 and Arg8 are crucial residues for OP4 receptor activation. Manipulation of Phe1 yielded peptides endowed with antagonist activity but [Nphe1] NC(1–13)-NH2 acted as an antagonist still possessing weak agonist activity. Introduction of the βAla residue either in position 7 or 11 of the [Nphe1] NC(1–13)-NH2 sequence, abolished any residual agonist activity and [Nphe1, βAla7] NC(1–13)-NH2 and [Nphe1, βAla11] NC(1–13)-NH2 acted as competitive antagonists only. Modification of both Ala7 and Ala11 abolished the antagonist activity of [Nphe1]NC(1–13)-NH2 probably by hindering receptor binding. Changes at positions 10 and 11 gave analogues still possessing agonist activity. [Ser(βGal)10] NC(1–13)-NH2 displayed an activity comparable with that of NC(1–13)-NH2, [Ser(βGalNAc)10] NC(1–13)-NH2 and [βAla11] NC(1–13)-NH2 were five and 10 times less active, respectively.The α-amino acid residues are of the l-configuration. Standard abbreviations for amino acid derivatives and peptides are according to the suggestions of the IUPAC-IUB Commission on Biochemical Nomeclature (1984), Eur. J. Biochem. 138, 9–37. Abbreviations listed in the guide published in (2003), J. Peptide Sci. 9, 1–8 are used without explanation.  相似文献   

19.
Abstract

High-field nuclear magnetic resonance measurements were carried out on substance P fragments SP4–11 [pGlu5]-SP5–11 and [pGlu6]SP6–11 both at 400 and at 500 MHz. A spectral simulation was carried out on two of these peptides and the coupling constants were interpreted in terms of the conformations. The JNH-CHa coupling constants are all ~8 Hz, with the exception of glycine, indicating no preferred conformation for the backbone. For the amino acids other than p-Glu, a comparison of the coupling constant data suggests the same relative rotamer populations for the side chains. Proton longitudinal relaxation time data were measured for all three peptides and support the above conclusions.  相似文献   

20.
A pair of l ‐leucine (l ‐Leu) and d ‐leucine (d ‐Leu) was incorporated into α‐aminoisobutyric acid (Aib) peptide segments. The dominant conformations of four hexapeptides, Boc‐l ‐Leu‐Aib‐Aib‐Aib‐Aib‐l ‐Leu‐OMe (1a), Boc‐d ‐Leu‐Aib‐Aib‐Aib‐Aib‐l ‐Leu‐OMe (1b), Boc‐Aib‐Aib‐l ‐Leu‐l ‐Leu‐Aib‐Aib‐OMe (2a), and Boc‐Aib‐Aib‐d ‐Leu‐l ‐Leu‐Aib‐Aib‐OMe (2b), were investigated by IR, 1H NMR, CD spectra, and X‐ray crystallographic analysis. All peptides 1a,b and 2a,b formed 310‐helical structures in solution. X‐ray crystallographic analysis revealed that right‐handed (P) 310‐helices were present in 1a and 1b and a mixture of right‐handed (P) and left‐handed (M) 310‐helices was present in 2b in their crystalline states. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  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