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.
It has been reported previously that a cyclic dipeptide, cyclo(D -Leu-L -His), showed a high hydrolytic activity toward a hydrophobic ester, p-nitrophenyl laurate. In order to determine the reason for the high catalytic activity, the conformation of cyclo(D -Leu-L -His) in aqueous solution was investigated by nuclear magnetic resonance and circular dichroism spectroscopy and compared with the conformation of cyclo(L -Leu-L -His), which was nearly inactive in otherwise the same conditions for the hydrolysis. It was demonstrated that the spatial arrangement of the hydrophobic isobutyl group of the D -leucyl residue and of the nucleophilic imidazolyl group of the L -histidyl residue in cyclo(D -Leu-L -His) matches very well with the long acyl chain and the active ester function of p-nitrophenyl laurate. On the other hand, in cyclo(L -Leu-L -His) the hydrophobic and the nucleophilic pendant groups are too close with each other to cooperate intramolecularly for the hydrolysis. It was concluded that the different steric structures of the diastereomers can explain the large difference of the catalytic activities.  相似文献   

2.
Cyclic dipeptide cyclo(l- or d-Glu-l-His) carrying an anionic site and a nucleophilic site has been synthesized and used as a catalyst for the solvolysis of cationic esters in aqueous alcohols. In the solvolysis of 3-acyloxy-N-trimethylanilinium iodide (S+n, n = 2 and 10) and Cl?H3N+(CH2)11COOPh(NO2), no efficient nucleophilic catalysis was observed. On the other hand, in the solvolysis of Gly-OPh(NO2)·HCl, Val-OPh(NO2)·HCl and Leu-OPh(NO2)·HCl a very efficient general base-type catalysis by cyclo(l-Glu-l-His) was observed. In particular, with the latter two substrates the catalysis by cyclo(l-Glul-His) was more efficient than that by imidazole, although the catalysis was not enantiomer-selective. The diastereomeric cyclic dipeptide cyclo(d-Glu-l-His) was almost inactive under the same conditions. Confomation of cyclo(l- or d-Glu-l-His) in aqueous solution was investigated and the structure/catalysis relationship is discussed.  相似文献   

3.
Tripeptides with cyclic dipeptide backbones, cyclo[l-Glu(l-Leu-O Bzl)-t-His] and cyclo[l-Glu(l-Leu-OH)-Ir His, and the corresponding tripeptides with linear backbones, Me3COCO-l-Glu(l-Leu-OBzl)-l-His-OMe and Me3COCO-l-Glu(l-Leu-OH)-l-His-OMe, were synthesized and used as catalysts for the hydrolysis of carboxylic acid active esters of various types. The experimental results are summarized as follows. (I) In the hydrolysis of a neutral and hydrophobic substate, p-nitrophenyl laurate, in 20% dioxane/H2O mixture of pH 7.8, a hydrophobic and flexible peptide, Me3COCO-l-Glu(l-Leu-OH)-l-His-OMe, was more reactive than imidazole. On the other hand, cyclo[l-Glu(l-Leu-OBzl)-l-His] and cyclo[l-Leu-OH)-l-His], which have rigid backbone chain and fixed sidechain conformation, were not particularly reactive. (2) in the solcolysis of a positively charged substrate, p-nitrophenyl glycinate hydrochloride, in 42% i-PrOH/H2O mixture at pH 6.95, a positively charged substrate, p-nitrophenyl glycinate hydrochloride, in 42% i-PrOH/H2O mixture at pH 6.95, a negatively charged and flexible peptide, Me3COCO-l-Glu(l-Leu-OH)-l-His-OMe, was more reactive than imidazole. However, cyclo [l-Glu(l-Leu-OH)-l-His] was not particularly reactive in the same reaction. In the hydrolysis of p-nitrophenyl glycinate hydrochloride in aqueous solution at pH 7.8 a hydrophobic and rigid peptide, cyclo[(l-Glu(l-Leu-OBzl)-l-His], was more reactive than imidazole. However, in the hydrolysis of p-nitrophenyl CO-AMINODODECANOATE hydrochloride, which has a positive charge and a rective site separated by a long hydrophobic chain, peptide catalysts did not show efficient catalysis. (3) In the hydrolysis of a positively charged, hydrophobic and chiral substrate, p-nitrophenyl leucinate hydrochloride, in aqueous solution at pH 6.95, the d-enantiomer was hydrolysed more quickly that the t-enantiomer with cyclo[l-Glu(l-Leu-OBzl)l-His] or cyclo[t-Glu(l-Leu-OH)-l-His] as catalyst. On the other hand, the tripeptides with linear backbone did not effect an enantiomer-selective catalysis. The solvolytic reaction catalysed by the tripeptides with cyclic dipeptide backbone in 42% i-PrOH/water mixture was also enantiomer-selective.  相似文献   

4.
Conformation in aqueous solution at pH 6.95 of tripeptides having cyclic dipeptide backbones, cyclo[l-Glu(l-Leu-OBzl)-l-His] and cyclo[l-Glu(l-Leu-OH)-l-His], was investigated by u.v., c.d. and n.m.r. spectroscopy and by the lanthanide probe method. In the major conformation of cyclo[l-Glu(l-Leu-OBzl)-l-His], the cyclic dipeptide backbone takes a flagpole-boat conformation in which the sidechain of the l-His residue is nearly parallel with the backbone plane and the sidechain of the l-Glu residue protrudes outside the backbone plane. In the major conformation of cyclo[l-Glu(l-Leu-OH)-l-His], the cyclic dipeptide backbone takes a flagpole-boat conformation in which the sidechains of the l-His and l-Glu residues are accommodated in the same side of the backbone plane so that the imidazolyl sidechain of l-His residue is twisted slightly. Tripeptides were not found to change the conformation when metal salts or ammonium salts such as Cl?H3N?(CH2)11 COOEt, Gly-OEt-HCl, dl-Val-OEt-HCl and l-Leu-OEt-HCl were added, but a significant conformation change occurred upon adding d-Leu-OEt·HCl. If the same situation holds with the addition of α-amino acid p-nitrophenyl ester hydrochlorides, the previously reported enantiomer-selective catalysis by the tripeptides which hydrolysed d-Leu-OPh(NO2·HCl faster than l-Leu-OPh(NO2)·HCl can be explained; that is, the tripeptides change the conformation only when d-Leu-OPh(NO2)·HCl is bound and consequently the intramolecular reaction is facilitated. This phenomenon may be compared with that of ‘induced fit’ in enzyme catalysis.  相似文献   

5.
Cyclo(D -Leu-L -Leu) and cyclo(L -Leu-L -Leu) were synthesized, and their carbon-13 nmr spectra at 65 MHz were examined in dimethylsulfoxide and trifluoroacetic acid solutions. The chemical shift data are consistent with a boat or “twisted” boat conformation of the diketopiperazine ring in both solvents. There was no indication of protonation of the cyclic dipeptides by trifluoroacetic acid. Attempts at polymerizing the cyclic dipeptides were unsuccessful.  相似文献   

6.
Cyclo(His-Pro), or histidyl-proline diketopiperazine, is an endogenous cyclic dipeptide that is ubiquitously distributed in tissues and body fluids of both man and animals. This cyclic dipeptide is not only structurally related to thyrotropin-releasing hormone (TRH, pGlu-His-ProNH2), but it can also arise from TRH by the action of the enzyme pyroglutamate amino-peptidase (pGlu-peptidase). The data on the distribution of TRH, cyclo(His-Pro), and pGlu-peptidase under normal and abnormal conditions are summarized and potential relationships analyzed. We conclude that all of the cyclo(His-Pro) cannot be derived from TRH. Two additional sources of cyclo(His-Pro) are suggested. It is proposed that 29,247 molecular weight TRH prohormone, prepro TRH, which contains 5 copies of TRH sequence, can be processed to yield cyclo(His-Pro). Thus, both TRH and cyclo(His-Pro) share a common precursor, prepro[TRH/Cyclo(His-Pro)].  相似文献   

7.
Cyclo(L -Pro-Sar)n (n = 2–4) with moderate flexibility and hydrophobicity of molecular structure was synthesized, and the characteristics of these cyclic peptides and their metal complexes in acetonitrile were investigated in connection with the residual properties using 13C-nmr measurements. The cyclic tetrapeptide cyclo(L -Pro-Sar)2 showed a sterically hindered phenomenon in acetonitrile in which the amide backbone adopted a cis-trans-cis-trans sequence. The cyclic hexapeptide cyclo(L -Pro-Sar)3 existed as a mixture of several conformers whose interconversion is slow on the nmr time scale, including cis-cis-trans and/or cis-trans-trans arrangement of the Sar-Pro bond. Finally, it was demonstrated that the cyclic octapeptide cyclo(L -Pro-Sar)4 behaved as a mixture of multiple conformers which allowed for cis-trans isomerism about the Pro-Sar peptide bond, of which 20–30% had the all-cis Sar-Pro bond isomer and the remaining 70–80% had one (or more) cis Sar-Pro bond isomer. 13C-nmr spectra also demonstrated that cyclo(L -Pro-Sar)n (n = 3,4) formed a 1:1 ion complex whose conformation was characterized by an all-trans peptide bond in the presence of excess metal salt. Cation binding studies, using CD measurements, established that the ion selectivity of cyclo(L -Pro-Sar)4 in acetonitrile decreased in the order, Ba2+ > Ca2+ > Na+ > Mg2+ > Li+.  相似文献   

8.
Cation transport through a chloroform liquid membrane by cyclic octapeptides—cyclo(Leu-Pro)4, cyclo(Phe-Pro)4, and cyclo[Lys(Z)-Pro]4—was investigated. All of these cyclic octapeptides transported K+ and Ba2+, and the rate of cation transport was correlated with the ability to extract cations from the aqueous phase to the chloroform phase. Among them, cyclo (Leu-Pro)4 was the most efficient and transported K+ and Ba2+ selectively from other alkali and alkaline earth cations, respectively. The rate of K+ transport by cyclo(Leu-Pro)4 was about one-third as fast as that by dicyclohexyl 18-crown-6. Picrate anion transport against its concentration gradient was observed by cyclo(Leu-Pro)4, which is conjugated with the selective transport of K+. Complex formation in a liposome between cyclo(Leu-Pro)4 and Ba2+ was observed, but the binding constant was low.  相似文献   

9.
Cyclic octapeptides, cyclo(X-Pro)4, where X represents Phe, Leu, or Lys(Z), were synthesized and their conformations investigated. A C2-symmetric conformer containing two cis peptide bonds was found in all of these cyclic octapeptides. The numbers of available conformations due to the cistrans isomerization of Pro peptide bonds depended on the nature of the solvent and X residue: they decreased in the following order: cyclo[Lys(Z)-Pro]4 > cyclo(Leu-Pro)4 > cyclo(Phe-Pro)4 in CDCl3. 13C spin-lattice relaxation times (T1) of these cyclic octapeptides were measured, and the contribution of segmental mobility to T1 was found to vary with the nature of the X residue.  相似文献   

10.
A complete series of configurationally isomers (L -L , L -D , D -L AND D -D ) of a dipeptide Leu-Phe benzyl ester have been synthesized and assayed for chymotrypsin. In the conformational analysis by 400 MMz 1H NMR, the L -D and D -L isomers, but not hte L -L and D -D isomers, showed fairly large up field shifts (0.2–0.4 ppm) of Leu-βCH2 and γCH proton signals, indicating the presence of shielding effects from the benzene ring. In addition to distinct signal splitting of Phe-βCH2, the NOE enhancement observed between Leu-δCH3 and Phe-phenyl groups revealed that these groups are in close proximity. These data indicated that L -D and D -L isomers from a hydrophobic core between side chains of adjacent Leu and Phe residues. When the dipeptides were examined for inhibition of chymotrypsin using Ac-Try-OEt as a substrate, the L -L isomer showed no inhibition, itself becoming a substrate. However, the other three isomers inhibited chymotrypsin in a competitive manner, and the D -L isomer was strongest with Ki of 2.2 × 10?5 M . It was found that the D -L isomer was only slowly hydrolysed but the L (or D )-D isomer was not. H-D -Phe-L -Leu-OBzl with the inverse sequence of H-D -Leu-L -Pre-OBzl inhibited chymotrypsin more strongly (Ki = 6.3 × 10?6 M ). Since the free acid analogue of the D -L isomer exhibited no inhibition, the benzyl ester moiety itself was thought to be involved in the enzyme inhibition. It is assumed that in the inhibitory conformation the ester-benzyl group fits the S1 site of chymotrypsin, while the side chain-side chain complexing hydrophobic core fits the S2 site.  相似文献   

11.
C M Deber  P D Adawadkar 《Biopolymers》1979,18(10):2375-2396
We have synthesized and characterized a series of cation-binding cyclic octapeptides which may function as potential ionophoric substances. The materials contain varying degrees of hydrophobic character, which was controlled systematically through the incorporation of N-alkylglycine residues where N-alkyl = methyl, n-hexyl, cyclohexyl, or n-decyl. The peptides reported include cyclo(Phe-Sar-Gly-Sar)2, cyclo(Glu(OBzl)-Sar-Gly-Sar-Glu(OBzl)-Sar-Gly-(N-decyl)Gly), cyclo(Glu(OBzl)-Sar-Gly-(N-decyl)Gly)2, cyclo(Glu(OBzl)-Sar-Gly-(N-hexyl)Gly)2, cyclo(Glu(OBzl)-Sar-Gly-(N-cyclohexyl)Gly)2, and the corresponding free diacid forms of the Glu-containing compounds. Using 13C- and 1H-nmr spectra, we demonstrated that the mixture of cis/trans peptide bond-isomer conformers, characteristic of the free-peptide benzyl esters in solution, was converted to unique C2-symmetric, presumably all-trans conformers on complexation with calcium ions. Cation-transport experiments, using the thick-liquid model of transport in a Pressman cell, established that these compounds transport a variety of cations and that one peptide examined in detail, cyclo(Glu(OBzl)-Sar-Gly-(N-decyl)Gly)2 (selectivity Ca2+ > Na+ > K+ > Mn2+ > Cu2+ > Mg2+ > Co2+ > Zn2+), transports calcium about an order of magnitude more efficiently than magnesium.  相似文献   

12.
Cyclo(His–Phe) was effectively converted to its dehydro derivatives by the enzyme of Streptomyces albulus KO-23, an albonoursin-producing actinomycete. Two types of dehydro derivatives were isolated from the reaction mixture and identified as cyclo(ΔHis–ΔPhe) and cyclo(His–ΔPhe). This is the first report on cyclo(His–ΔPhe) and the enzymatic preparation of both compounds. Cyclo(ΔHis–ΔPhe), a tetradehydro cyclic dipeptide, exhibited a minimum inhibitory concentration of 0.78 μmol/ml inhibitory activity toward the first cleavage of sea urchin embryos, in contrast to cyclo(His–ΔPhe) that had no activity. The finding that the isoprenylated derivative of cyclo(ΔHis–ΔPhe), dehydrophyenylahistin, had 2,000 times higher activity than cyclo(ΔHis–ΔPhe) indicates that an isoprenyl group attached to an imidazole ring of the compound was essential for the inhibitory activity.  相似文献   

13.
The side-chain conformations of D - orL - Thr, D - or L -Ser, L -Asp, and L - His residues in cyclic and linear dipeptides in D2O or in DMSO-d6 are deduced from vicinal (1H,1H) and (13C, 1H) coupling constants. Vicinal (13C, 13C) coupling constants strongly depend on substituents and cannot be used without a more sound analysis. In cyclic dipeptides, the Thr and Ser side chains are folded above the DKP ring, with χ1 near 60°. The L -Asp side chain interacts more specifically with peptide bonds (χ1 near 300°). The L - His side chain is more flexible and its conformation depends on the proximity of a second side chain and on solute-solvent interactions. In all cases, this side chain is not completely folded. In linear dipeptides, the conformation of a C-terminal L -His residue is mainly influenced by the end carboxylic group. On the other hand, a N-terminal L -His residue interacts more easily with a neighboring L -Asp residue. In aqueous solution, the imidazole pKa depends on the proximity of terminal and lateral charged groups but does not reveal any specific interaction in cyclic dipeptides. A comparison between the conformations of cyclic peptides observed in solution, in the crystalline state and calculated by empirical methods, allows one to point out the discriminating role of the packing in crystals, and of solute-solvent interactions in solution.  相似文献   

14.
The possibility of selectively reducing the number of β-helical structures theoretically possible for a D ,L -alternating peptide by using a N-methyl group as conformational constraint is considered. Some 1H-nmr data regarding Boc(L -Nle-D -Nle)3-L -Nle-D -MeNle -L -Nle-D -Nle-L -Nle-OMe (I), its formyl analogue (II), and the pentadecapeptide Boc(D -Leu-L -Leu)5-D -MeLeu -(L -Leu-D -Leu)2-OMe (III) are presented. It is shown that these alternating stereocooligopeptides with a N-methyl group in the (n ? 3) (I and II) or (n ? 4) position (III) differ drastically in their behavior from the corresponding nonmethylated compounds. In chloroform, I and II form predominantly ↑↓ β7.2-helices and III forms almost exclusively ↑↓ β5.6 or ↑↓ β7.2-helices. The helices are in every case those having the maximum possible number of interchain H bonds.  相似文献   

15.
Summary Mixtures of cyclic peptides, formed by head-to-tail cyclizations of side-chain resin-bound linear sequences, have been prepared using solid-phase synthesis. Fast atom bombardment mass spectrometry of cyclic peptides with various metal ions can reveal preferred modes of host-guest patterns, albeit in a nonquantitative manner. This approach could prove useful for more rapid screening of potential peptide ionophores. A cyclic heptapeptide with a dipeptide tail proved to be a particularly effective host for a Ca2+ ion; in a small three-component mixture, cyclo[Gly-Asp-d-Pro-Xxx-Asp-d-Pro-Asp(Aca-Phe-NH2)], binding to Ca2+ varied from Xxx=N-MeAla>GlySar. In a 15-component mixture, cyclo[Pro-Xxx-Asn-Pro-Xxx-Asn] where Xxx=Ala, Glu, Leu, Lys or Phe, there were no significant differences with respect to binding to metal ions. We believe this to be the first reported use of cyclic peptide libraries for screening metal ions to discern host-guest relationships.Abbreviations Aca aminocaproic acid - Boc tert-butyloxycarbonyl - BOP benzotriazolyloxy-tris(dimethylamino)phosphonium hexafluorophosphate - DCM dichloromethane - DIEA diisopropylethylamine - DMF N,N-dimethylformamide - ESI electrospray ionization - FABMS fast atom bombardment mass spectrometry - pMBHA 4-methylbenzhydrylamine - TFA trifluoroacetic acid This paper is based on a presentation given at the Symposium on Peptide Structure and Design as part of the 31st Annual ACS Western Regional Meeting held in San Diego, CA, USA, October 18–21, 1995.  相似文献   

16.
Complex formation with alkali and alkaline earth metal ions of cyclic octapeptides, cyclo(Phe-Pro)4, cyclo(Leu-Pro)4, and cyclo[Lys(Z)-Pro]4 was investigated in relation to conformation. In an alcohol solution, cyclo(Phe-Pro)4 did not form complexes. However, cyclo(Leu-Pro)4 and cyclo[Lys(Z)-Pro]4 formed complexes selectively with Ba2+ and Ca2+ ions. Changing the solvent from alcohol to acetonitrile, the complexation behavior was very different. In acetonitrile, cyclo(Phe-Pro)4 was found to form a complex with Ba2+, and CD spectra of cyclo(Leu-Pro)4 and cyclo[Lys(Z)-Pro]4 changed sharply on complexation with K+. Rate constants of the complex formation between the cyclic octapeptides and metal salts were in the range of 0.7–12 L mol?1 min?1 in an alcohol solution. One of the two types of complex formation in acetonitrile was much faster than that in an alcohol solution.  相似文献   

17.
Two series of peptides with hydrophobic side chains, Nps-(L -Leu-L -Leu-L -Ala)n-OEt and Nps-(L -Met-L -Met-L Leu)n-OEt (n = 1–6), were synthesized by the fragment condensation method using dicyclohexylcarbodiimide in the presence of N-hydroxysuccinimide. The tripeptide fragments were prepared stepwise by dicyclohexylcarbodiimide-mediated reaction of Nps-amino acids, which were synthesized by an improved rapid procedure.  相似文献   

18.
The 1H-nmr spectra (270 MHz) of cyclic di- and tri-L -azatidine-2-carboxylic acid [cy-clo(L -Aze)2 and cyclo(L -Aze)3] were determined in CDCl3 and D2O and computer simulated. The spectral results were compared with those obtained with cyclo (L -Pro)2 and cyclo (L, -Pro)3. In CDCl3 and D2O solution, the four membered ring of cyclo (L -Aze)2 is puckered with the α-proton in a pseudo-axial position, and the ? angle is smaller in absolute value than ?60°, as found for cyclo (L -Pro)2,. The puckering of the four-membered ring of cyclo(L, -Aze)3 in CDCl3 has the α-proton in a pseudo–equatorial position and ? angle larger in absolute value than ?60°, in agreement with cyclo(L -Pro)3. In D2O, cyclo(L -Aze)3 was found to interconvert rapidly between different conformers. In the azetidine cyclic peptides studied, the range of values found for the ? angles was smaller than in the related proline cyclic peptides, indicating greater rigidity in the four-membered ring.  相似文献   

19.
The solid state conformations of cyclo[Gly–Proψ[CH2S]Gly–D –Phe–Pro] and cyclo[Gly–Proψ[CH2–(S)–SO]Gly–D –Phe–Pro] have been characterized by X-ray diffraction analysis. Crystals of the sulfide trihydrate are orthorhombic, P212121, with a = 10.156(3) Å, b = 11.704(3) Å, c = 21.913(4) Å, and Z = 4. Crystals of the sulfoxide are monoclinic, P21, with a = 10.662(1) Å, b = 8.552(3) Å, c = 12.947(2) Å, β = 94.28(2), and Z = 2. Unlike their all-amide parent, which adopts an all-trans backbone conformation and a type II β-turn encompassing Gly-Pro-Gly-D -Phe, both of these peptides contain a cis Gly1-Pro2 bond and form a novel turn structure, i.e., a type II′ β-turn consisting of Gly–D –Phe–Pro–Gly. The turn structure in each of these peptides is stabilized by an intramolecular H bond between the carbonyl oxygen of Gly1 and the amide proton of D -Phe4. In the cyclic sulfoxide, the sulfinyl group is not involved in H bonding despite its strong potential as a hydrogen-bond acceptor. The crystal structure made it possible to establish the absolute configuration of the sulfinyl group in this peptide. The two crystal structures also helped identify a type II′ β-turn in the DMSO-d6 solution conformers of these peptides. © 1993 John Wiley & Sons, Inc.  相似文献   

20.
A new microbial cyclic dipeptide (diketopiperazine), cyclo(d ‐Tyr‐d ‐Phe) was isolated for the first time from the ethyl acetate extract of fermented modified nutrient broth of Bacillus sp. N strain associated with rhabditid Entomopathogenic nematode. Antibacterial activity of the compound was determined by minimum inhibitory concentration and agar disc diffusion method against medically important bacteria and the compound recorded significant antibacterial against test bacteria. Highest activity was recorded against Staphylococcus epidermis (1 µg/ml) followed by Proteus mirabilis (2 µg/ml). The activity of cyclo(d ‐Tyr‐d ‐Phe) against S. epidermis is better than chloramphenicol, the standard antibiotics. Cyclo(d ‐Tyr‐d ‐Phe) recorded significant antitumor activity against A549 cells (IC50 value: 10 μM) and this compound recorded no cytotoxicity against factor signaling normal fibroblast cells up to 100 μM. Cyclo(d ‐Tyr‐d ‐Phe) induced significant morphological changes and DNA fragmentation associated with apoptosis in A549 cells. Acridine orange/ethidium bromide stained cells indicated apoptosis induction by cyclo(d ‐Tyr‐d ‐Phe). Flow cytometry analysis showed that the cyclo(d ‐Tyr‐d ‐Phe) did not induce cell cycle arrest. Effector molecule of apoptosis such as caspase‐3 was found activated in treated cells, suggesting apoptosis as the main mode of cell death. Antioxidant activity was evaluated by free radical scavenging and reducing power activity, and the compound recorded significant antioxidant activity. The free radical scavenging activity of cyclo(d ‐Tyr‐d ‐Phe) is almost equal to that of butylated hydroxyanisole, the standard antioxidant agent. We also compared the biological activity of natural cyclo(d ‐Tyr‐d ‐Phe) with synthetic cyclo(d ‐Tyr‐d ‐Phe) and cyclo(l ‐Tyr‐l ‐Phe). Natural and synthetic cyclo(d ‐Tyr‐d ‐Phe) recorded similar pattern of activity. Although synthetic cyclo(l ‐Tyr‐l ‐Phe) recorded lower activity. But in the case of reducing power activity, synthetic cyclo(l ‐Tyr‐l ‐Phe) recorded significant activity than natural and synthetic cyclo(d ‐Tyr‐d ‐Phe). The results of the present study reveals that cyclo(d ‐Tyr‐d ‐Phe) is more bioactive than cyclo(l ‐Tyr‐l ‐Phe). To the best of our knowledge, this is the first time that cyclo(d ‐Tyr‐d ‐Phe) has been isolated from microbial natural source and also the antibacterial, anticancer, and antioxidant activity of cyclo(d ‐Tyr‐d ‐Phe) is also reported for the first time. Copyright © 2013 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