首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Abstract: Peptide bond bioisosteres, such as hydroxyethylamine (Hea), have frequently been used to stabilize metabolically labile peptide bonds in peptidomimetic drug design in an effort to increase the oral bioavailability of drug candidates. However, the impact of the peptide bond bioisosteres on the cell permeation characteristics of peptidomimetics is not well understood, particularly with respect to the effects on the substrate activity for proteins that can restrict (e.g. P‐glycoprotein, P‐gp) or facilitate (e.g. the oligopeptide transporter, OPT) intestinal mucosal permeation of peptidomimetics. In this study, terminally free and terminally modified (N‐acetylated and C‐amidated) peptidomimetics of H‐Ala‐Phe‐OH and H‐Ala‐Phe‐Ala‐OH with the Ala‐Phe peptide bonds replaced by Hea bioisosteres were synthesized. Transport characteristics of these peptidomimetics were investigated using Caco‐2 cell monolayers as an in vitro model of the intestinal mucosa. The study showed that the Hea bioisostere stabilized the peptidomimetics to protease metabolism in Caco‐2 cells. All terminally free peptidomimetics showed significant affinity and substrate activity for OPT. The affinity and substrate activity for OPT were stereoselective for peptidomimetics containing an S,S‐configuration for the two adjacent chiral centers related to the Hea bioisostere. Three of the four terminally modified peptidomimetics showed significant substrate activity for P‐gp and, interestingly, the substrate activity for P‐gp was also stereoselective; however, it was in favor of an R,R‐configuration for the two adjacent chiral centers related to the Hea bioisostere.  相似文献   

2.
Abstract: Different types of dipeptide building units containing N‐ or C‐terminal arginine were prepared for synthesis of the backbone cyclic analogues of the peptide hormone bradykinin (BK: Arg‐Pro‐Pro‐Gly‐Phe‐Ser‐Pro‐Phe‐Arg). For cyclization in the N‐terminal sequence N‐carboxyalkyl and N‐aminoalkyl functionalized dipeptide building units were synthesized. In order to avoid lactam formation during the condensation of the N‐terminal arginine to the N‐alkylated amino acids at position 2, the guanidino function has to be deprotected. The best results were obtained by coupling Z‐Arg(Z)2‐OH with TFFH/collidine in DCM. Another dipeptide building unit with an acylated reduced peptide bond containing C‐terminal arginine was prepared to synthesize BK‐analogues with backbone cyclization in theC‐terminus. To achieve complete condensation to the resin and to avoid side reactions during activation of the arginine residue, this dipeptide unit was formed on a hydroxycrotonic acid linker. HYCRAM? technology was applied using the Boc‐Arg(Alloc)2‐OH derivative and the Fmoc group to protect the aminoalkyl function. The reduced peptide bond was prepared by reductive alkylation of the arginine derivative with the Boc‐protected amino aldehyde, derived from Boc‐Phe‐OH. The best results for condensation of the branching chain to the reduced peptide bond were obtained using mixed anhydrides. Both types of dipeptide building units can be used in solid‐phase synthesis in the same manner as amino acid derivatives.  相似文献   

3.
Abstract: The synthesis of conformationally restricted dipeptidic moieties 4‐amino‐1,2,4,5‐tetrahydro‐2‐benzazepin‐3‐one (Aba)‐Gly ([(4S)‐amino‐3‐oxo‐1,2,4,5‐tetrahydro‐1H‐2‐benzazepin‐2‐yl]‐acetic acid) and 8‐hydroxy‐4‐amino‐1,2,4,5‐tetrahydro‐2‐benzazepin‐3‐one (Hba)‐d ‐Ala ([(4S)‐amino‐8‐hydroxy‐3‐oxo‐1,2,4,5‐tetrahydro‐benzo[c]azepin‐2‐yl]‐propionic acid) was based on a synthetic strategy that uses an oxazolidinone as an N‐acyliminium precursor. Introducing these Aba scaffolds into the N‐terminal tetrapeptide of dermorphin (H‐Tyr‐d ‐Ala‐Phe‐Gly‐Tyr‐Pro‐Ser‐NH2)‐induced remarkable shifts in affinity and selectivity towards the opioid μ‐ and δ‐receptors. This paper provides the synthesis and biological in vitro and in vivo evaluation of constricted analogues of the N‐terminal tetrapeptide H‐Tyr‐d ‐Ala‐Phe‐Gly‐NH2, which is the minimal subunit of dermorphin needed for dermorphin‐like opiate activity.  相似文献   

4.
Abstract: Solid‐phase synthesis and aminolysis cleavage conditions were optimized to obtain N‐ and C‐terminally protected hydrophobic peptides with both high quality and yield. Uncharged ‘WALP’ peptides, consisting of a central (Leu‐Ala)n repeating unit (where n = 5, 10.5 or 11.5) flanked on both sides by Trp ‘anchors’, and gramicidin A (gA) were synthesized using 9‐fluorenylmethoxycarbonyl chemistry from either Wang or Merrifield resins. For WALP peptides, the N‐terminal amino acid was capped by coupling N‐acetyl‐ or N‐formyl‐Ala or ‐Gly to the peptide/resin or by formylation of the completed peptide/resin with para‐nitrophenylformate (p‐NPF). N‐Terminal acetyl‐ or formyl‐Ala racemized when coupled as an HOBt‐ester to the resin‐bound peptide, but not when the peptide was formylated with p‐NPF. Racemization was avoided at the last step by completing the peptide with acetyl‐ or formyl‐Gly. For both WALP peptides and gA, cleavage conditions using ethanolamine or ethylenediamine were optimized as functions of solvent, time, temperature and resin type. For WALP peptides, maximum yields of highly pure peptide were obtained by cleavage with 20% ethanolamine or ethylenediamine in 80% dichloromethane for 48 h at 24°C. N‐Acetyl‐protected WALP peptides consistently gave higher yields than those protected with N‐formyl. For gA, cleavage with 20% ethanolamine or ethylenediamine in 80% dimethylformamide for 48 h at 24°C gave excellent results. For both WALP peptides and gA, decreasing the cleavage time to 4 h and increasing the temperature to 40–55°C resulted in significantly lower yields. The inclusion of hexafluoroisopropanol in the cleavage solvent mixture did not improve yields for either gA or WALP peptides.  相似文献   

5.
Abstract: The melanocortin receptor (MCR) pathway has been identified as participating in several physiologically important pathways including pigmentation, energy homeostasis, inflammation, obesity, hypertension, and sexual function. All the endogenous MCR agonists contain a core His‐Phe‐Arg‐Trp sequence identified as important for receptor molecular recognition and stimulation. Several structure–activity studies using the Ac‐His‐d ‐Phe‐Arg‐Trp‐NH2 tetrapeptide template have been performed in the context of modifying N‐terminal ‘capping’ groups and amino acid constituents. Herein, we report the synthesis and pharmacologic characterization of modified Xaa‐d ‐Phe‐Arg‐Trp‐NH2 (Xaa = His or Phe) melanocortin tetrapeptides (N‐site selective methylation, permethylation, or amide bond reduction) at the mouse MC1, MC3, MC4 and MC5 receptors. The modified peptides generated in this study resulted in equipotent or reduced MCR potency when compared with control ligands. The reduced amide bond analog of the Phe‐d ‐Phe‐Arg‐Trp‐NH2 peptide converted its agonist activity into an antagonistic at the central mMC3 and mMC4 receptors involved in the regulation of energy homeostasis, while retaining full agonist activity at the peripheral MC1 and MC5 receptors.  相似文献   

6.
Abstract: This study details a series of conditions that may be applied to ensure ‘safe’ incorporation of cysteine with minimal racemization during automated or manual solid‐phase peptide synthesis. Earlier studies from our laboratories [Han et al. (1997) J. Org. Chem. 62 , 4307–4312] showed that several common coupling methods, including those exploiting in situ activating agents such as N‐[(dimethylamino)‐1H‐1,2,3‐triazolo[4,5‐b]pyridin‐1‐ylmethylene]‐N‐methylmethanaminium hexafluorophosphate N‐oxide (HATU), N‐[1H‐benzotriazol‐1‐yl)‐(dimethylamino)methylene]‐N‐methylmethanaminium hexafluorophosphate N‐oxide (HBTU), and (benzotriazol‐1‐yl‐N‐oxy)tris(dimethylamino)phosphonium hexafluorophosphate (BOP) [all in the presence of N‐methylmorpholine (NMM) or N,N‐diisopropylethylamine (DIEA) as a tertiary amine base], give rise to unacceptable levels (i.e. 5–33%) of cysteine racemization. As demonstrated on the tripeptide model H‐Gly‐Cys‐Phe‐NH2, and on the nonapeptide dihydrooxytocin, the following methods are recommended: O‐pentafluorophenyl (O‐Pfp) ester in DMF; O‐Pfp ester/1‐hydroxybenzotriazole (HOBt) in DMF; N,N′‐diisopropylcarbodiimide (DIPCDI)/HOBt in DMF; HBTU/HOBt/2,4,6‐trimethylpyridine (TMP) in DMF (preactivation time 3.5–7.0 min in all of these cases); and HBTU/HOBt/TMP in CH2Cl2/DMF (1:1) with no preactivation. In fact, several of the aforementioned methods are now used routinely in our laboratory during the automated synthesis of analogs of the 58‐residue protein bovine pancreatic trypsin inhibitor (BPTI). In addition, several highly hindered bases such as 2,6‐dimethylpyridine (lutidine), 2,3,5,6‐tetramethylpyridine (TEMP), octahydroacridine (OHA), and 2,6‐di‐tert‐butyl‐4‐(dimethylamino)pyridine (DB[DMAP]) may be used in place of the usual DIEA or NMM to minimize cysteine racemization even with the in situ coupling protocols.  相似文献   

7.
Abstract: Previous research has shown that the immunomodulatory peptide α‐melanocyte‐stimulating hormone (α‐MSH) and its carboxy‐terminal tripeptide KPV (Lys‐Pro‐Val α‐MSH11?13) have antimicrobial influences. By inserting a Cys‐Cys linker between two units of KPV, we designed the dimer [Ac‐CKPV]2 that showed excellent candidacidal effects in pilot tests and was the subject of further investigations. [Ac‐CKPV]2 was active against azole‐resistant Candida spp. Therefore, the molecule appeared a promising candidate for therapy of fungal infections and was the subject of a structural study. 1H‐NMR and restrained mechanic and dynamic calculations suggest that the peptide adopts an extended backbone structure with a β‐turn‐like structure. These results open a pathway to development of additional novel compounds that have candidacidal effects potentially useful against clinical infections.  相似文献   

8.
To develop PSA peptide substrates with improved specificity and plasma stability from the known substrate sequence glutaryl‐Hyp‐Ala‐Ser‐Chg‐Gln, systematic replacements of the N‐terminal segment with D‐retro‐inverso‐peptides were performed with the incorporation of 7‐amino‐4‐methylcoumarin (7‐AMC) after Gln for convenient fluorometric determination and ranking of the PSA substrate activity. The D‐retro‐inverso‐peptide conjugates with P2‐P5 D‐amino acid substitutions were moderate but poorer PSA substrates as compared to the original peptide, suggesting that inversion of the amide bonds and/or incorporation of the additional atom as in the urea linker adversely affected PSA binding. However, P5 substitution of Hyp with Ser showed significant improvements in PSA cleavage rate; the resulting AMC conjugate, glutaryl‐Ser‐Ala‐Ser‐Chg‐Gln‐AMC ( 11 ), exhibited the fastest PSA cleavage rate of 351 pmol/min/100 nmol PSA. In addition, GABA←mGly‐Ala‐Ser‐Chg‐Gln‐AMC (conjugate 6 ) was the second best PSA substrate and released 7‐AMC at a rate of 225 pmol/min/100 nmol PSA as compared to 171 pmol/min/100 nmol PSA for the control conjugate glutaryl‐Hyp‐Ala‐Ser‐Chg‐Gln‐AMC. Incubations of selected AMC conjugates with mouse and human plasma revealed that GABA←D‐Ser‐ψ[NH‐CO‐NH]‐Ala‐Ser‐Chg‐Gln‐AMC ( 5 ) and GABA←mGly‐Ala‐Ser‐Chg‐Gln‐AMC ( 6 ) were most stable to non‐PSA‐mediated proteolysis. Our results suggest that the PSA specificity of glutaryl‐Hyp‐Ala‐Ser‐Chg‐Gln is improved with Ser and mGly substitutions of Hyp at the P5.  相似文献   

9.
Abstract: The crystal structures of the peptides, Boc‐Leu‐Trp‐Val‐OMe ( 1) , Ac‐Leu‐Trp‐Val‐OMe ( 2a and 2b), Boc‐Leu‐Phe‐Val‐OMe ( 3 ), Ac‐Leu‐Phe‐Val‐OMe ( 4 ), and Boc‐Ala‐Aib‐Leu‐Trp‐Val‐OMe ( 5 ) have been determined by X‐ray diffraction in order to explore the nature of interactions between aromatic rings, specifically the indole side chain of Trp residues. Peptide 1 adopts a type I β‐turn conformation stabilized by an intramolecular 4→1 hydrogen bond. Molecules of 1 pack into helical columns stabilized by two intermolecular hydrogen bonds, Leu(1)NH…O(2)Trp(2) and IndoleNH…O(1)Leu(1). The superhelical columns further pack into the tetragonal space group P43 by means of a continuous network of indole–indole interactions. Peptide 2 crystallizes in two polymorphic forms, P21 ( 2a ) and P212121 ( 2b ). In both forms, the peptide backbone is extended, with antiparallel β‐sheet association being observed in crystals. Extended strand conformations and antiparallel β‐sheet formation are also observed in the Phe‐containing analogs, Boc‐Leu‐Phe‐Val‐OMe ( 3 ) and Ac‐Leu‐Phe‐Val‐OMe ( 4 ). Peptide 5 forms a short stretch of 310‐helix. Analysis of aromatic–aromatic and aromatic–amide interactions in the structures of peptides, 1 , 2a , 2b are reported along with the examples of 14 Trp‐containing peptides from the Cambridge Crystallographic Database. The results suggest that there is no dramatic preference for a preferred orientation of two proximal indole rings. In Trp‐containing peptides specific orientations of the indole ring, with respect to the preceding and succeeding peptide units, appear to be preferred in β‐turns and extended structures.  相似文献   

10.
Abstract: In order to make clear the structural role of the C‐terminal amide group of endomorphin‐2 (EM2, H‐Tyr‐Pro‐Phe‐Phe‐NH2), an endogenous µ‐receptor ligand, in the biological function, the solution conformations of endomorphin‐2 and its C‐terminal free acid (EM2OH, H‐Tyr‐Pro‐Phe‐Phe‐OH), studied using two‐dimensional 1H NMR measurements and molecular modeling calculations, were compared. Both peptides were in equilibrium between the cis and trans isomers around the Tyr‐Pro ω bond in a population ratio of ≈ 1 : 2. The lack of significant temperature and concentration dependence of NH protons suggested that the NMR spectra reflected the conformational features of the respective molecules themselves. Fifty possible 3D structures for the each isomer were generated by the dynamical simulated annealing method under the proton?proton distance constraints derived from the ROE cross‐peaks. These energy‐minimized conformers, which were all in the φ torsion angles estimated from JNHCαH coupling constants within ± 30°, were then classified in groups one or two according to the folding backbone structures. All trans and cis EM2 conformers adopt an open conformation in which their extended backbone structures are twisted at the Pro2–Phe3 moiety. In contrast, the trans and cis conformers of EM2OH show conformational variation between the ‘bow’‐shaped extended and folded backbone structures, although the cis conformers of its zwitterionic form are refined into the folded structure of the close disposition of C‐ and N‐terminal groups. These results indicate clearly that the substitution of carboxyl group for C‐terminal amide group makes the peptide flexible. The conformational requirement for µ‐receptor activation has been discussed based on the active form proposed for endomorphin‐1 and by comparing conformational features of EM2 and EM2OH.  相似文献   

11.
Abstract: A series of three homologous dimethyldiamides Ac‐ΔAla‐NMe2, Ac‐l ‐Ala‐NMe2 and Ac‐dl ‐Ala‐NMe2 has been synthesized and the structures of these amides determined from single‐crystal X‐ray diffraction data. To learn more about the conformational preferences of compounds studied, the fully relaxed (φ–ψ) conformational energy maps in vacuo (AM1) of Ac‐ΔAla‐NMe2 and Ac‐l ‐Ala‐NMe2 were obtained, and the calculated minima reoptimized with the DFT/B3LYP/6–31G** method. The crystal‐state results have been compared with the literature data. Ac‐ΔAla‐NMe2 and other α,β‐dehydroamino acid dimethyldiamides, Ac‐ΔXaa‐NMe2 adopt the conservative conformation of the torsion angles φ, ψ = ~ ?45°, ~130°, which are located in the high‐energy region (region H) of Ramachandran diagram. Ac‐l ‐Ala‐NMe2 and Ac‐dl ‐Ala‐NMe2, as well as other saturated amino acid dimethylamides Ac‐l /dl ‐Xaa‐NMe2, present common peptide structures, and no conformational preferences are observed. Molecular packing of the amides analysed reveals two general hydrogen‐bonded motifs. Dehydro and dl ‐species are paired into centrosymmetric dimers, and l ‐compounds form catemers. However, Ac‐ΔAla‐NMe2 and Ac‐DL ‐Ala‐NMe2 constitute exceptions: their molecules also link into catemers.  相似文献   

12.
13.
Abstract: Protein farnesyltransferase (PFTase) catalyzes the attachment of a geranylazide (C10) or farnesylazide (C15) moiety from the corresponding prenyldiphosphates to a model peptide substrate, N‐dansyl‐Gly‐Cys‐Val‐Ile‐Ala‐OH. The rates of incorporation for these two substrate analogs are comparable and approximately twofold lower than that using the natural substrate farnesyl diphosphate (FPP). Reaction of N‐dansyl‐Gly‐Cys(S‐farnesylazide)‐Val‐Ile‐Ala‐OH with 2‐diphenylphosphanylbenzoic acid methyl ester then gives a stable alkoxy‐imidate linked product. This result suggests future generations whereby azide groups introduced using this enzymatic approach are functionalized using a broad range of azide‐reactive reagents. Thus, chemistry has been developed that could be used to achieve highly specific peptide and protein modification. The farnesylazide analog may be useful in certain biological studies, whereas the geranylazide group may be more useful for general protein modification and immobilization.  相似文献   

14.
Abstract: Single N‐methyl amino acid‐containing peptides related to the central hydrophobic region β16–20 (Lys‐Leu‐Val‐Phe‐Phe) of the β‐amyloid protein are able to reduce the cytotoxicity of natural β1–42 in PC12 cell cultures. N‐methyl phenylalanine analogs yield statistically significant increments in cell viability (Student's t‐test < 0.01%) and are nontoxic in the same assay. These promising results indicate that these peptide molecules could be a starting point for the development of potential therapeutic compounds for the treatment of Alzheimer's disease.  相似文献   

15.
Abstract: In order to extend the use of proteases to organic synthesis and seek the rules of enzymatic reactions in organic media, we focused on unnatural substrates for proteases to form amide bonds. In this paper, the study of unnatural substrates containing d ‐amino acid residue, which act as acyl acceptors as well as acyl donors for proteases in organic media, is reported. Dermorphin is a heptapeptide (H‐Tyr‐d ‐Ala‐Phe‐Gly‐Tyr‐Pro‐Ser‐NH2) with potent analgesic activity. The N‐terminal tetrapeptide is the minimum sequence that retains dermorphin activity, and is selected as the model compound in our study. Two dermorphin‐(1–4) derivatives, Boc‐Tyr‐d ‐Ala‐Phe‐Gly‐N2H2Ph and Boc‐Tyr‐d ‐Ala‐Phe‐Gly‐NH2, which contained a d ‐amino acid residue, were synthesized by proteases in organic media for the first time. The synthesis of these two dermorphin‐(1–4) derivatives could be catalyzed by subtilisin with Boc‐Tyr‐d ‐Ala‐OCH2CF3 as an acyl donor substrate in AcOEt. The synthesis of dermorphin‐(1–2) derivative Boc‐Tyr‐d ‐Ala‐N2H2Ph was catalyzed by α‐chymotrypsin in different organic solvents and d ‐Ala‐N2H2Ph was used as an acyl acceptor substrate. Factors influencing the above enzymatic reactions were systematically studied.  相似文献   

16.
Abstract: A series of cyclic, disulfide‐ or dithioether‐containing tetrapeptides based on previously reported potent μ‐ and δ‐selective analogs has been explored with the aim of improving their poor affinity to the κ‐opioid receptor. Specifically targeted were modifications of tetrapeptide residues 3 and 4, as they presumably interact with residues from transmembrane helices 6 and 7 and extracellular loop 3 that differ among the three receptors. Accordingly, tetrapeptides were synthesized with Phe3 replaced by aliphatic (Gly, Ala, Aib, Cha), basic (Lys, Arg, homo‐Arg), or aromatic sides chains (Trp, Tyr, p‐NH2Phe), and with d ‐Pen4 replaced by d ‐Cys4, and binding affinities to stably expressed μ‐, δ‐, and κ‐receptors were determined. In general, the resulting analogs failed to exhibit appreciable affinity for the κ‐receptor, with the exception of the tetrapeptide Tyr‐c[d ‐Cys‐Phe‐d ‐Cys]‐NH2, cyclized via a disulfide bond, which demonstrated high binding affinity toward all opioid receptors (Kiμ = 1.26 nm , Kiδ = 16.1 nm , Kiκ = 38.7 nm ). Modeling of the κ‐receptor/ligand complex in the active state reveals that the receptor‐binding pocket for residues 3 and 4 of the tetrapeptide ligands is smaller than that in the μ‐receptor and requires, for optimal fit, that the tripeptide cycle of the ligand assume a higher energy conformation. The magnitude of this energy penalty depends on the nature of the fourth residue of the peptide (d ‐Pen or d ‐Cys) and correlates well with the observed κ‐receptor binding affinity.  相似文献   

17.
In vivo metabolism of the radiolabelled receptor‐specific peptides has been described; however, information regarding the pharmacokinetic behaviour of the degradation products within the body is very scarce. The present study was designed to obtain new knowledge on the disposition and elimination of low‐molecular radiometabolites of receptor‐specific peptides in the organism and to reveal the potential involvement of selected membrane transport mechanisms in the cellular uptake of radiometabolites, especially in the kidney. The study compared pharmacokinetics of two radiometabolites: a final metabolite of somatostatin analogues, 177Lu‐DOTA‐DPhe, and a tripeptide metabolite of 177Lu‐DOTA‐minigastrin 11, 177Lu‐DOTA‐DGlu‐Ala‐Tyr. Their pharmacokinetics was compared with that of respective parent 177Lu‐radiopeptide. Both radiometabolites exhibited relative rapid clearing from most body tissues in rats in vivo along with predominant renal excretion. The long‐term renal retention of the smaller radiometabolite 177Lu‐DOTA‐DPhe was lower than that of 177Lu‐DOTA‐DGlu‐Ala‐Tyr. An uptake of 177Lu‐DOTA‐DPhe by human renal influx transporter organic cation transporter 2 was found in vitro in a cellular model. The study brings the first experimental data on the in vivo pharmacokinetics of radiometabolites of receptor‐specific somatostatin and gastrin analogues. The found results may indicate a negative correlation between the degree of decomposition of the parent peptide chain and the renal retention of the metabolite.  相似文献   

18.
Abstract: Protecting groups in N‐ and C‐terminal positions play a decisive role in the conformational preference of smaller peptides. Conformational analysis of tetrapeptide derivatives containing Ala, Ile and Gly residues was performed. Peptide 1 , Boc‐Ala‐Ile‐Ile‐Gly‐OMe (Boc: tert‐butyloxycarbonyl) has a predominantly helical turn conformation in all the alcoholic solvents studied, whereas in the solid state it has a β‐sheet conformation. In contrast, peptide 2 , Ac‐Ala‐Ile‐Ile‐Gly‐OMe (Ac: acetyl) has a random coil conformation in solution. The FTIR spectrum of peptide 1 shows a lower frequency of urethane carbonyl, indicating involvement of the carbonyl group in hydrogen bonding in the helical turn.  相似文献   

19.
Abstract: The glycopeptide hormone catfish somatostatin (somatostatin‐22) has the amino acid sequence H‐Asp‐Asn‐Thr‐Val‐Thr‐Ser‐Lys‐Pro‐Leu‐Asn‐Cys‐Met‐Asn‐Tyr‐Phe‐Trp‐Lys‐Ser‐Arg‐Thr‐Ala‐Cys‐OH; it includes a cyclic disulfide connecting the two Cys residues, and the major naturally occurring glycoform contains d ‐GalNAc and d ‐Gal O‐glycosidically linked to Thr5. The linear sequence was assembled smoothly starting with an Fmoc‐Cys(Trt)‐PAC‐PEG‐PS support, using stepwise Fmoc solid‐phase chemistry. In addition to the nonglycosylated peptide, two glycosylated forms of somatostatin‐22 were accessed by incorporating as building blocks, respectively, NαFmoc‐Thr(Ac3‐α‐D‐GalNAc)‐OH and Nα‐Fmoc‐Thr(Ac4‐β‐D‐Gal‐(1→3)‐Ac2‐α‐D‐GalNAc)‐OH. Acidolytic deprotection/cleavage of these peptidyl‐resins with trifluoroacetic acid/scavenger cocktails gave the corresponding acetyl‐protected glycopeptides with free sulfhydryl functions. Deacetylation, by methanolysis in the presence of catalytic sodium methoxide, was followed by mild oxidation at pH 7, mediated by Nαdithiasuccinoyl (Dts)‐glycine, to provide the desired monomeric cyclic disulfides. The purified peptides were tested for binding affinities to a panel of cloned human somatostatin receptor subtypes; in several cases, presence of the disaccharide moiety resulted in 2‐fold tighter binding.  相似文献   

20.
Abstract: During a study aimed at generating a bispecific molecule between BN antagonist (d ‐Trp6,Leu13‐ψ[CH2NH]‐Phe14)BN6‐14 (Antag1) and mAb22 (anti‐FcγRI), we attempted to cross‐link the two molecules by introducing a thiol group into Antag1 via 2‐iminothiolane (2‐IT, Traut's reagent). We found that reaction of Antag1 with 2‐IT, when observed using HPLC, affords two products, but that the later eluting peptide is rapidly transformed into the earlier eluting peptide. To understand what was occurring we synthesized a model peptide, d ‐Trp‐Gln‐Trp‐NH2 (TP1), the N‐terminal tripeptide of Antag1. Reaction of TP1 with 2‐IT for 5 min gave products 1a and 3a ; the concentration of 1a decreased with reaction time, whereas that of 3a increased. Thiol 1a , the expected Traut product, was identified by collecting it in a vial containing N‐methylmaleimide and then isolating the resultant Michael addition product 2a , which was confirmed by mass spectrometry. Thiol 1a is stable at acidic pH, but is unstable at pH 7.8, cyclizes and loses NH3 to give N‐TP1‐2‐iminothiolane ( 3a ), ES‐MS (m/z) [602.1 (M+H)+], as well as regenerating TP1. Repeat reaction with Antag1 and 2‐IT allowed us to isolate N‐Antag1–2‐iminothiolane ( 3b ), FAB‐MS (m/z) [1212.8 (M+H)+] and trap the normal Traut product 1b as its N‐methylmaleimide Michael addition product 2b , ES‐MS (m/z) [1340.8 (M+H)+]. Thiol 1b is also stable at acidic pH, but when neutralized is unstable and cyclizes, forming 3b and Antag1.  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