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.
Via a Mannich reaction involving a dibenzyliminium species and the titanium enolates of Evans' chiral acylated oxazolidinones the β2‐amino acids (R)‐ and (S)‐Fmoc‐β2homovaline and (R)‐Fmoc‐β2homoleucine are synthesized. These building blocks were used, in combination with commercially available α‐ and β3‐amino acids, for the synthesis of the cyclo‐(αβ3αβ2α)2 peptide 2 and the cyclo‐(αβ2αβ3α)2 peptides 3 – 5 . The peptides 2 – 5 were screened for their ability to inhibit a small panel of Gram‐negative and Gram‐positive bacterial strains.  相似文献   

2.
β‐Amino acids containing hybrid peptides and β‐peptides show great potential as peptidomimetics. In this paper we describe the synthesis and affinity toward the µ‐ and δ‐opioid receptors of β‐peptides, analogues of Leu‐enkephalin, deltorphin I, dermorphin and α,β‐hybrides, analogues of deltorphin I. Substitution of α‐amino acid residues with β3homo‐amino acid residues, in general resulted in decrease of affinity to opioid receptors. However, the incorporation β3h‐D ‐Ala in position 2 or β3hPhe in position 3 of deltorphin I resulted in potent and selective ligand for δ‐opioid receptor. The NMR studies of β‐deltorphin I analogue suggest that conformational motions in the central part of the peptide backbone are partially restricted and some conformational preferences can be expected. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
Aggregation, orientation and dynamics of transmembrane helices are of relevance for protein function and transmembrane signaling. To explore the interactions of transmembrane helices and the interdependence of peptide structure and lipid composition of the membranes, β‐peptides were explored as model transmembrane domains. Various hydrophobic β‐peptide sequences were synthesized by solid phase peptide synthesis. Conformational analyses of β‐peptide helices were performed in organic solvents (methanol and 2,2,2‐trifluoroethanol) and in large unilamellar liposomes (dimyristoylphosphatidylcholine, dipalmitoylphosphatidylcholine and dioleoylphosphatidylcholine) indicating 12‐ and 14‐helix conformations, depending on β3‐amino acid sequences. The intrinsic tryptophan fluorescence of β3‐homotryptophan units inserted in the center or near the end of the sequence was used to verify the membrane insertion of the β‐peptides. A characteristic blue shift with peripheral β3‐homotryptophan compared with β‐peptides with central tryptophan served as indication for a transmembrane orientation of the β‐peptides within the lipid bilayer. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
To examine the roles of TGFβ isoforms on corneal morphogenesis, the eyes of mice that lack TGFβs were analyzed at different developmental stages for cell proliferation, migration and apoptosis, and for expression patterns of keratin 12, lumican, keratocan and collagen I. Among the three Tgfb−/− mice, only Tgfb2−/− mice have abnormal ocular morphogenesis characterized by thin corneal stroma, absence of corneal endothelium, fusion of cornea to lens (a Peters'-like anomaly phenotype), and accumulation of hyaline cells in vitreous. In Tgfb2−/− mice, fewer keratocytes were found in stroma that has a decreased accumulation of ECM; for example, lumican, keratocan and collagen I were greatly diminished. The absence of TGFβ2 did not compromise cell proliferation, nor enhance apoptosis. The thinner stroma resulting from decreased ECM synthesis may account for the decreased cell number in the stroma of Tgfb2 null mice. Keratin 12 expression was not altered in Tgfb2−/− mice, implicating normal corneal type epithelial differentiation. Delayed appearance of macrophages in ocular tissues was observed in Tgfb2−/− mice. Malfunctioning macrophages may account for accumulation of cell mass in vitreous of Tgfb2 null mice.  相似文献   

5.
3β-Hydroxysteroid dehydrogenase (3β-HSD)/Δ5→4-isomerase activity in steroidogenic tissues is required for the synthesis of biologically active steroids. Previously, by use of dehydroepiandrosterone (3β-hydroxy-5-androsten-17-one, DHEA) as substrate, it was established that in addition to steroidogenic tissues 3β-HSD/Δ5→4-isomerase activity also is expressed in extraglandular tissues of the human fetus. In the present study, we attempted to determine whether the C-5,C-6-double bond of DHEA serves to influence 3β-HSD activity. For this purpose, we compared the efficiencies of a 3β-hydroxy-5-ene steroid (DHEA) and a 3β-hydroxy-5α-reduced steroid (5α-androstane-3β,17β-diol, 5α-A-diol) as substrates for the enzyme. The apparent Michaelis constant (Km) for 5α-A-diol in midtrimester placenta, fetal liver, and fetal skin tissues was at least one order of magnitude higher than that for DHEA, viz the apparent Km of placental 3β-HSD for 5α-A-diol was in the range of 18 to 40 μmol/l (n = 3) vs 0.45 to 4 μmol/l for DHEA (n = 3); for the liver enzyme, 17 μmol/l for 5α-A-diol and 0.60 μmol/l for DHEA, and for the skin enzyme 14 and 0.18 μmol/l, respectively. Moreover, in 13 human fetal tissues evaluated the maximal velocities obtained with 5α-A-diol as substrate were higher than those obtained with DHEA. A similar finding in regard to Kms and rates of product formation was obtained by use of purified placental 3β-HSD with DHEA, pregnenolone, and 3β-hydroxy-5α-androstan-17-one (epiandrosterone) as substrates: the Km of 3β-HSD for DHEA was 2.8 μmol/l, for pregnenolone 1.9 μmol/l, and for epiandrosterone 25 μmol/l. The specific activity of the purified enzyme with pregnenolone as substrate was 27 nmol/mg protein·min and, with epiandrosterone, 127 nmol/mg protein·min. With placental homogenate as the source of 3β-HSD, DHEA at a constant level of 5 μmol/l behaved as a competitive inhibitor when the radiolabeled substrate, [3H]5α-A-diol, was present in concentrations of 20 to 60 μmol/l, but a lower substrate concentrations the inhibition was of the mixed type; similar results were obtained with [3H]DHEA as the substrate at variable concentrations in the presence of a fixed concentration of 5α-A-diol (40 μmol/l). These findings are indicative that both steroids bind to a common site on the enzyme, however, the binding affinity for these steroids appear to differ markedly as suggested by the respective Kms. Studies of inactivation of purified placental 3β-HSD/Δ5→4-isomerase by an irreversible inhibitor, viz 5,10-secoestr-4-yne-3,10,17-trione, were suggestive that the placental protein adopts different conformations depending on whether the steroidal substrate has a 5α-configuration, e.g. epiandrosterone, or a C-5,C-6-double bond e.g. DHEA or pregnenolone. The lower rates of product formation obtained with placenta and fetal tissues by use of 3β-hydroxy-5-ene steroids as substrates when compared with those obtained with 3β-hydroxy-5α-reduced steroids may be explained by a combination of factors, including: (i) inhibition of 3β-HSD activity by end products of metabolism of 3β-hydroxy-5-ene steroids, e.g. 4-androstene-3,17-dione formed with DHEA as substrate; (ii) higher binding affinity of the enzyme for 3β-hydroxy-5-ene steroids—and possibly for their 3-oxo-5-ene metabolites; (iii) lack of a requirement for the isomerization step with 5α-reduced steroids as substrates, and (iv) the possible presence in fetal tissues of an enzyme with 3β-HSD activity only (i.e. no Δ5→4-isomerase).  相似文献   

6.
Sequential tritylation, benzoylation, and detritylation of methyl 3-deoxy-3-fluoro-β- -galactopyranoside gave crystalline methyl 2,4-di-O-benzoyl-3-deoxy-3-fluoro-β- -galactopyranoside (9), which was used as the initial nucleophile in the synthesis of the target oligosaccharide (16). Treatment of 9 with 2,3,4-tri-O-benzoyl-6-O-bromoacetyl-α- -galactopyranosyl bromide gave the corresponding disaccharide derivative 13, having a selectively removable blocking group at O-6′. Debromoacetylation of 13 afforded the disaccharide nucleophile 14 which, when treated with 2,4,6-tri-O-benzoyl-3-deoxy-3-fluoro-α- -galactopyranosyl bromide, gave the fully protected trisaccharide 15. Debenzoylation of 15 gave the title glycoside 16. Condensation reactions were performed with silver trifluoromethane-sulfonate as a promoter in the presence of sym-collidine under base-deficient conditions, and gave excellent yields of the desired β-(trans)-products. Analyses of the 1H- and 13C-n.m.r. spectra, as well as determination of the JCF and JHF coupling constants, were made by using various one- and two-dimensional n.m.r. techniques.  相似文献   

7.
The present work is devoted to the synthesis, conformational analysis, and stereodynamic study of aza‐β3‐cyclodipeptides. This pseudopeptidic ring shows E/Z hydrazide bond isomerism, eight‐membered ring conformation, and chirotopic nitrogen atoms, all of which are elements that are prone to modulate the ring shape. The (E,E) twist boat conformation observed in the solid state by X‐ray diffraction is also the ground conformation in solution, and emerges as the lowest in energy when using quantum chemical calculations. The relative configuration associated with ring chirality and with the two nitrogen chiral centers is governed by steric crowding and adopts the (P)SNSN/(M)RNRN combination which projects side chains in equatorial position. The nitrogen pyramidal inversion (NPI) at the two chiral centers is correlated with the ring reversal. The process is significantly hindered as was shown by VT‐NMR experiments run in C2D2Cl4, which did not make it possible to determine the barrier to inversion. Finally, these findings make it conceivable to resolve enantiomers of aza‐β3‐cyclodipeptides by modulating the backbone decoration appropriately. Chirality 25:341–349, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

8.
The 4‐methoxybenzyloxymethyl (MBom) group was introduced at the Nπ‐position of the histidine (His) residue by using a regioselective procedure, and its utility was examined under standard conditions used for the conventional and the microwave (MW)‐assisted solid phase peptide synthesis (SPPS) with 9‐fluorenylmethyoxycarbonyl (Fmoc) chemistry. The Nπ‐MBom group fulfilling the requirements for the Fmoc strategy was found to prevent side‐chain‐induced racemization during incorporation of the His residue even in the case of MW‐assisted SPPS performed at a high temperature. In particular, the MBom group proved to be a suitable protecting group for the convergent synthesis because it remains attached to the imidazole ring during detachment of the protected His‐containing peptide segments from acid‐sensitive linkers by treatment with a weak acid such as 1% trifluoroacetic acid in dichloromethane. We also demonstrated the facile synthesis of Fmoc‐His(π‐MBom)‐OH with the aid of purification procedure by crystallization to effectively remove the undesired τ‐isomer without resorting to silica gel column chromatography. This means that the present synthetic procedure can be used for large‐scale production without any obstacles. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Ammonium 2,6-anhydro-3-deoy- -glycero- -talo-octonate (1), a potent inhibitor of the enzyme CMP-KDO synthetase, its C-2 epimer 2, and the methyl β-(3) and α-glycoside (4) of KDO were studied by 1H- and 13C-n.m.r. spectroscopy. Compound 1 was also analysed by X-ray crystallography. Each compound adopted a 5C2 chair conformation with the side chain equatorial. The preponderant side-chain conformation of 1 in solution was the same as that in the crystal and was stabilised by an intramolecular hydrogen bond from HO-8 to the carboxylate group. This hydrogen bond appeared to be present also in 3. However, the side-chain conformation of 2 and 4 was different from that in 1 and 3. The metal-ion-binding properties, determined on the basis of the line-broadening effects of Mn2+ on the 13C-n.m.r. signals, showed that the carboxylate group was involved in the binding with O-8 in 1 and 3 and with O-6 and O-8 in 2 and 4.  相似文献   

10.
Besides hydrolyzing different membrane phospholipids, plant phospholipases D and molecular species of their byproducts phosphatidic acids (PLDs/PAs) are involved in diverse cellular events such as membrane‐cytoskeleton dynamics, hormone regulation and biotic and/or abiotic stress responses at cellular or subcellular levels. Among the 12 Arabidopsis PLD genes, PLDζ1 and PLDζ2 uniquely possess Ca2+‐independent phox (PX) and pleckstrin (PH) homology domains. Here, we report that mutants deficient in these PLDs, pldζ1 and pldζ2, show differential sensitivities to hypoxia stimulus. In the present study, we used protoplasts of wild type and mutants and compared the hypoxia‐induced changes in the levels of three major signaling mediators such as cytoplasmic free calcium [Ca2+cyt.], hydrogen peroxide (H2O2) and PA. The concentrations of cytosolic Ca2+ and H2O2 were determined by fluorescence microscopy and the fluorescent dyes Fura 2‐AM and CM‐H2DCFDA, specific for calcium and H2O2, respectively, while PA production was analyzed by an enzymatic method. The study reveals that AtPLDζ1 is involved in reactive oxygen species (ROS) signaling, whereas AtPLDζ2 is involved in cytosolic Ca2+ signaling pathways during hypoxic stress. Hypoxia induces an elevation of PA level both in Wt and pldζ1, while the PA level is unchanged in pldζ2. Thus, it is likely that AtPLDζ2 is involved in PA production by a calcium signaling pathway, while AtPLDζ1 is more important in ROS signaling.  相似文献   

11.
N‐(tert‐butyloxycarbonyl) or N‐(9‐fluorenylmethoxycarbonyl) dipeptides with C‐terminal (Z)‐α,β‐didehydrophenylalanine (?ZPhe), (Z)‐α,β‐didehydrotyrosine (?ZTyr), (Z)‐α,β‐didehydrotryptophan (?ZTrp), (Z)‐α,β‐didehydromethionine (?ZMet), (Z)‐α,β‐didehydroleucine (?ZLeu), and (Z/E)‐α,β‐didehydroisoleucine (?Z/EIle) were synthesised from their saturated analogues via oxidation of intermediate 2,5‐disubstituted‐oxazol‐5‐(4H)‐ones (also known as azlactones) with pyridinium tribromide followed by opening of the produced unsaturated oxazol‐5‐(4H)‐one derivatives in organic‐aqueous solution with a catalytic amount of trifluoroacetic acid or by a basic hydrolysis. In all cases, a very strong preference for Z isomers of α,β‐didehydro‐α‐amino acid residues was observed except of the ΔIle, which was obtained as the equimolar mixture of Z and E isomers. Reasons for the (Z)‐stereoselectivity and the increased stability of the aromatic α,β‐didehydro‐α‐amino acid residue oxazol‐5‐(4H)‐ones over the corresponding aliphatic ones are also discussed. It is the first use of such a procedure to synthesise peptides with the C‐terminal unsaturated residues and a peptide with 2 consecutive ΔPhe residues. This approach is very effective especially in the synthesis of peptides with aliphatic α,β‐didehydro‐α‐amino acid residues that are difficult to obtain by other methods. It allowed the first synthesis of the ?Met residue. It is also more cost‐effective and less laborious than other synthesis protocols. The dipeptide building blocks obtained were used in the solid‐phase synthesis of model peptides on a polystyrene‐based solid support. Peptides containing aromatic α,β‐didehydro‐α‐amino acid residues were obtained with PyBOP or TBTU as a coupling agent with good yields and purities. In the case of aliphatic α,β‐didehydro‐α‐amino acid residues, a good efficiency was achieved only with DPPA as a coupling agent.  相似文献   

12.
To obtain general rules of peptide design using α,β-dehydro-residues, a sequence with two consecutive ΔPhe-residues, Boc-L -Val-ΔPhe–ΔPhe- L -Ala-OCH3, was synthesized by azlactone method in solution phase. The peptide was crystallized from its solution in an acetone/water mixture (70:30) in space group P61 with a=b=14.912(3) Å, c= 25.548(5) Å, V=4912.0(6) Å3. The structure was determined by direct methods and refined by a full matrix least-squares procedure to an R value of 0.079 for 2891 observed [I?3σ(I)] reflections. The backbone torsion angles ?1=?54(1)°, ψ1= 129(1)°, ω1=?177(1)°, ?2 =57(1)°, ψ2=15(1)°, ω2 =?170(1)°, ?3=80(1)°, ψ3 =7(2)°, ω3=?177(1)°, ?4 =?108(1)° and ψT4=?34 (1)° suggest that the peptide adopts a folded conformation with two overlapping β-turns of types II and III′. These turns are stabilized by two intramolecular hydrogen bonds between the CO of the Boc group and the NH of ΔPhe3 and the CO of Val1 and the NH of Ala4. The torsion angles of ΔPhe2 and ΔPhe3 side chains are similar and indicate that the two ΔPhe residues are essentially planar. The folded molecules form head-to- tail intermolecular hydrogen bonds giving rise to continuous helical columns which run parallel to the c-axis. This structure established the formation of two β-turns of types II and III′ respectively for sequences containing two consecutive ΔPhe residues at (i+2) and (i+3) positions with a branched β-carbon residue at one end of the tetrapeptide.  相似文献   

13.
Tuberculosis is still affecting millions of people worldwide, and new resistant strains of Mycobacterium tuberculosis are being found. It is therefore necessary to find new compounds for treatment. In this paper, we report the synthesis and in vitro testing of peptidyl β‐aminoboronic acids and β‐aminoboronates with anti‐tubercular activity. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
Condensation of 2,4,6-tri-O-acetyl-3-deoxy-3-fluoro-α- -galactopyranosyl bromide (3) with methyl 2,3,4-tri-O-acetyl-β- -galactopyranoside (4) gave a fully acetylated (1→6)-β- -galactobiose fluorinated at the 3′-position which was deacetylated to give the title disaccharide. The corresponding trisaccharide was obtained by reaction of 4 with 2,3,4-tri-O-acetyl-6-O-chloroacetyl-α- -galactopyranosyl bromide (5), dechloroacetylation of the formed methyl O-(2,3,4-tri-O-acetyl-6-O-chloroacetyl-β- -galactopyranosyl)-(1→6)- 2,3,4-tri-O-acetyl-β- -galactopyranoside to give methyl O-(2,3,4-tri-O-acetyl-β- -galactopyranosyl)-(1→6)-2,3,4-tri-O-acetyl-β- -galactopyranoside (14), condensation with 3, and deacetylation. Dechloroacetylation of methyl O-(2,3,4-tri-O-acetyl-6-O-chloroacetyl-β- -galactopyranosyl)-(1→6)-O-(2,3,4-tri-O-acetyl- β- -galactopyranosyl)-(1→6)-2,3,4-tri-O-acetyl-β- -galactopyranoside, obtained by condensation of disaccharide 14 with bromide 5, was accompanied by extensive acetyl migration giving a mixture of products. These were deacetylated to give, crystalline for the first time, the methyl β-glycoside of (1→6)-β- -galactotriose in high yield. The structures of the target compounds were confirmed by 500-MHz, 2D, 1H- and conventional 13C- and 19F-n.m.r. spectroscopy.  相似文献   

15.
The 17‐phenyl PGF analogue bimatoprost ( 10a ) is the most efficacious ocular hypotensive agent currently available for the treatment of glaucoma or ocular hypertension. A novel convergent synthesis of 13,14‐en‐15‐ol prostamideF analogues was developed employing Julia–Lythgoe olefination of the structurally advanced phenylsulfone (+)‐(5Z)‐ 15 with an enantiomerically pure aldehyde ω‐chain synthon (–)‐(S)‐ 16a . Subsequent hydrolysis of protecting groups and final amidation of the diol 26a yielded bimatoprost ( 10a ). The main advantage of the current strategy is the preparation of high‐purity bimatoprost ( 10a ). The novel convergent strategy allows the synthesis of a whole series of 13,14‐en‐15‐ol prostamideF analogues with the desired C‐15 asymmetric center configuration from a common and structurally advanced prostaglandin intermediate (+)‐(5Z)‐ 15 . The preparation and identification of two synthetic impurities, 15‐epi isomer ( 10b ) of bimatoprost and a new prostaglandin related amide (+)‐(5Z)‐ 18 , are also described. Chirality 25:170–179, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

16.
This review gives a broad overview of the state of play with respect to the synthesis, conformational properties, and biological activity of α‐fluorinated β‐amino acids and derivatives. General methods are described for the preparation of monosubstituted α‐fluoro‐β‐amino acids (Scheme 1). Nucleophilic methods for the introduction of fluorine predominantly involve the reaction of DAST with alcohols derived from α‐amino acids, whereas electrophilic sources of fluorine such as NFSI have been used in conjunction with Arndt? Eistert homologation, conjugate addition or organocatalyzed Mannich reactions. α,α‐Difluoro‐β‐amino acids have also been prepared using DAST; however, this area of synthesis is largely dominated by the use of difluorinated Reformatsky reagents to introduce the difluoro ester functionality (Scheme 9). α‐Fluoro‐β‐amino acids and derivatives analyzed by X‐ray crystal and NMR solution techniques are found to adopt preferred conformations which are thought to result from stereoelectronic effects associated with F located close to amines, amides, and esters (Figs. 26). α‐Fluoro amide and β‐fluoro ethylamide/amine effects can influence the secondary structure of α‐fluoro‐β‐amino acid‐containing derivatives including peptides and peptidomimetics (Figs. 79). α‐Fluoro‐β‐amino acids are also components of a diverse range of bioactive anticancer (e.g., 5‐fluorouracil), antifungal, and antiinsomnia agents as well as protease inhibitors where such fluorinated analogs have shown increased potency and spectrum of activity.  相似文献   

17.
Several stereoisomeric, monocyclic analogs of (−)‐β‐vetivone (1), one of the main constituents of vetiver oil, were studied to examine if the olfactory properties of (−)‐β‐vetivone (1) could be reproduced from these structurally simpler, synthetically accessible compounds. The effects of diastereomeric and enantiomeric structural differences on the odor of the partial vetivone structure were studied. A chiral phenyl sulfoximine was used for separation of the racemic mixtures. Detailed nuclear magnetic resonance (NMR)‐spectroscopic studies (1H, 13C) were used to determine relative configurations whereas absolute configurations were determined by circular dichroism (CD) methods. Chirality 11:133–138, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

18.
Type 3 17β-hydroxysteroid dehydrogenase (17β-HSD), a key steroidogenic enzyme, transforms 4-androstene-3,17-dione (Δ4-dione) into testosterone. In order to produce potential inhibitors, we performed solid-phase synthesis of model libraries of 3β-peptido-3α-hydroxy-5α-androstan-17-ones with 1, 2, or 3 levels of molecular diversity, obtaining good overall yields (23–58%) and a high average purity (86%, without any purification steps) using the Leznoff's acetal linker. The libraries were rapidly synthesized in a parallel format and the generated compounds were tested as inhibitors of type 3 17β-HSD. Potent inhibitors were identified from these model libraries, especially six members of the level 3 library having at least one phenyl group. One of them, the 3β-(N-heptanoyl- -phenylalanine- -leucine-aminomethyl)-3α-hydroxy-5α-androstan-17-one (42) inhibited the enzyme with an IC50 value of 227 nM, which is twice as potent as the natural substrate Δ4-dione when used itself as an inhibitor. Using the proliferation of androgen-sensitive (AR+) Shionogi cells as model of androgenicity, the compound 42 induced only a slight proliferation at 1 μM (less than previously reported type 3 17β-HSD inhibitors) and, interestingly, no proliferation at 0.1 μM.  相似文献   

19.
Glycogen synthase kinase‐3 (GSK‐3) plays an important regulatory role in various signaling pathways; such as PI3 K/AKT, which is closely related to the occurrence and development of tumors. At present, the most reported active GSK‐3 inhibitors have the same structure: lactam ring or amide structure. To find out the GSK‐3β small molecule inhibitor with novel, safe, efficient and more uncomplicated synthesis method, we analyzed in‐depth reported crystal‐binding patterns of GSK‐3β small molecule inhibitor with GSK‐3β protein, and designed and synthesized 17 non‐reported 3,5‐diamino‐N‐substituted benzamide compounds. Their structures were confirmed by 1H‐NMR, 13C‐NMR, and HR‐MS. The preliminary screening of tumor cytotoxicity of compounds in vitro was detected by MTT, and their structure–activity relationships were illustrated. The results have shown that 3,5‐diamino‐N‐[3‐(trifluoromethyl)phenyl]benzamide ( 4d ) exhibited significant tumor cytotoxicity against human colon cancer cells (HCT‐116) with IC50 of 8.3 μm and showed commendable selectivity to GSK‐3β. In addition, Compound 4d induced apoptosis to some extent and possessed modest PK properties.  相似文献   

20.
β‐Amino acids containing α,β‐hybrid peptides show great potential as peptidomimetics. In this paper, we describe the synthesis and affinity to μ‐opioid and δ‐opioid receptors of α,β‐hybrids, analogs of the tetrapeptide Tyr‐ d ‐Ala‐Phe‐Phe‐NH2 (TAPP). Each amino acid was replaced with an l ‐ or d ‐β3h‐amino acid. All α,β‐hybrids of TAPP analogs were synthesized in solution and tested for affinity to μ‐opioid and δ‐opioid receptors. The analog Tyr‐β3h‐ d ‐Ala‐Phe‐PheNH2 was found to be as active as the native tetrapeptide. 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