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 156 毫秒
1.
Synthetic cyclic octapeptides of general structure cyclo[Glu(γOBzl)-Sar-Gly-(N-R)Gly]2 (R = n-hexyl and cyclohexyl) transport calcium ions selectively across organic phases and phospholipid membranes. We have now used proton nmr spectroscopy (360 MHz) to study the solution conformation(s) of their calcium complexes. When Ca(ClO4)2 was added to solutions of these peptides in CDCl3, nmr spectra of the resulting calcium complexes were characteristic of a single C2-symmetric conformer. From a Karplus-Bystrov analysis of vicinal coupling constants in both the peptide backbone and Glu side chain (treated as an ABCCMX spin system), in conjuction with model-building studies, a structure was proposed in which the calcium ion is bound in an octahedral-type complex by the four (coplanar) carbonyl groups of the (all-trans) Glu-Sar and Gly-(N-R)Gly peptide bonds. Occurrence of preferred rotamers about Glu side chain Cα–Cβ bonds indicated that restricted rotation in peptide side chains arises upon calcium binding.  相似文献   

2.
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   

3.
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   

4.
Substances which can perturb the transmembrane cation balance in a predictable manner have wide-ranging uses in the study of cellular processes. We have undertaken to examine transmembrane calcium transport on the molecular level through the design and synthesis of a series of ionophoric peptides as models for protein-mediated calcium transport. General mechanisms for carrier-mediated membrane transport are discussed. Cation transport profiles are presented for transport by synthetic peptides of structure cyclo(Glu(OR)-Sar-Gly-(N-R1)-Gly)2, where R = benzyl ester or H; R1 = n-decyl or cyclohexyl. Transport of physiologically abundant cations across "liquid membranes" in Pressman cells mediated by cyclo(Glu-Sar-Gly-(N-decyl)Gly)2 was observed to be essentially calcium specific, as long as calcium ions were present in the system. Multilamellar and unilamellar phosphatidylcholine vesicles were each found to be emptied of internal 45Ca2+ ions upon addition of cyclo(Glu(OBz)-Sar-Gly-(N-cyclohexyl)Gly)2 to the vesicle suspension. The results are compared with the naturally occurring calcium ionophore A23187.  相似文献   

5.
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   

6.
Cyclic di- and tetrapeptides bearing carboxyl or carboxylate groups, cyclo[Glu(OBzl)-Glu(OMe)], cyclo[Glu-Glu(OMe)], cyclo(Glu-Glu), cyclo[Glu(OMe)-Pro)2, and cyclo(Glu-Pro)2, were synthesized and investigated on the intramolecular interaction of carboxyl side chains in the complexation with metal ions in relation with the conformation. The three kinds of cyclic dipeptides were found to take a flagpole boat conformation. Folded conformation of side chains was predominant for cyclo[Glu(OBzl)-Glu(OMe)] and cyclo[Glu-Glu(OMe)]. However, cyclo(Glu-Glu) took an unfolded conformation. Intramolecular interaction of carboxyl groups was observed neither in free state nor in complexation with metal ions. The intramolecular interaction of carboxyl groups was observed in the case of cyclo(Glu-Pro)2 in the absence of metal ions added. Cyclo[Glu(OMe)-Pro]2 and cyclo(Glu-Pro)2 formed a complex with Ca2+ and Ba2+ without participation of side chains.  相似文献   

7.
Bacillus sp. No. 106, which was isolated from soil, secreted an enzyme that hydrolyzed cyclo(Gly-Gly). The enzyme was purified to the ultracentrifugally homogeneous state and an activity more than 450-fold that of culture broth. The enzyme was activated by Na+, Mg2+, Ca2+, and Sr2+, and strongly inhibited by Ni2+, Cu2+, p-chloromercuribenzoate, and monoiodoacetic acid. The Km value for cyclo(Gly-Gly) was estimated to be 11.1 mm. The enzyme hydrolyzed only cyclo(Gly-Gly) among various diketopiperazines tested. Aslo, the enzyme was inert toward Gly-Gly, milk casein, and hemoglobin.  相似文献   

8.
The hydrodynamic properties of α-helical poly(L -glutamic acid), (Glu)n in aqueous solutions and in mixtures of water with organic solvents have been interpreted in terms of the persistence length of the macromolecule. A modification of the method of Vitovskaya and Tsvetkov has been proposed in order to allow a more accurate determination of this parameter. The addition of an organic solvent increases strongly the rigidity of the helical conformation of (Glu)n. A comparison is made with some data of the literature of poly[N5-(3-hydroxy propyl)L -glutamine], [Gln(CH2)3OH]n, and poly(γ-benzyl-L -glutamate), [Glu(OBzl)]n.  相似文献   

9.
The electric field dependence (up to 21 kV/cm) of the steady-state and decay signals has been examined on the four well-fractionated samples of poly(γ-benzyl-L -glutamate), [Glu(OBzl)]n, in N,N-dimethylformamide at 535 nm and 20°C. Together with the data previously obtained from the reversing-pulse electric birefringence [Ueda, K., Nomura, M. & Yamaoka, K. (1983) Biopolymers 22 , 2077–2090], the steady-state birefringence and field-free relaxation time were analyzed by a method that takes into account the polydispersity of the chain length. The weight-average chain length, (lw), permanent dipole moment, (μw), electric polarizability anisotropy, (Δαw), and the length-independent optical anisotropy factor were evaluated. The axial translation per residue was calculated for the [Glu(OBzl)]n helix, but the uncertainly involved in the weight-average molecular weights, determined from light scattering by different investigators, makes the determination of the exact conformation of [Glu(OBzl)]n difficult. The contribution of Δαw to electric field orientation was found to be significant, since Δαw was approximately proportional to lw. A linear relationship also exists between μw and lw, when the [Glu(OBzl)]n helix is shorter than about 1200 Å.  相似文献   

10.
Poly(Glu(OBzl)-Gly)n, poly(Glu-Gly)n, poly(Gly)-(Glu(OBzl)-Gly), and poly(Gly-Glu-Gly) were synthesized from the pentachlorophenyl esters of the sequential monomer. Both of the polymers containing free glumatic-acid residues are soluble in water, as is the lower molecular weight fraction of the polytripeptides with the benzyl ester in place. Circular dichroism studies and infrared dichroism studies suggest that the 21 helix is favored for the polydipeptide with removal of the benzyl ester reducing the conformational integrity. The polytripeptide showed evidence of 31 helix in addition to the 21 form, depending on solvent. A rationale for the conformations observed is developed based on the bulkiness of the side-chain residues and conformational stabilization, in certain cases, by hydrophobic interactions between the benzyl ester groups.  相似文献   

11.
Reversing-pulse electric birefringence (RPEB) was measured for the first time for four fractionated poly(γ-benzyl-L -glutamate), [Glu(OBzl)]n, samples in N,N-dimethylformamide (DMF) at 20°C and at 535 nm. The RPEB signal showed a deep minimum for each sample on reversal of an applied electric field. The profiles of the reverse-transient signal were analyzed by taking into account the polydispersity for the continuous distribution of molecular lengths. The best set of three quantities (lw, lw/ln, (βw)2/2γw), which determine a signal profile, was evaluated for each sample. By combining the experimental data of intrinsic viscosity and RPEB, the diameter of a cylinder, which is assumed for the [Glu(OBzl)]n helix, was found to be 17 Å. The value (βw)2/2γw, which is related to the ratio of weight-average permanent dipole moment μw, and electric polarizability anisotropy Δαw, was found to be in the range of 20–45. This indicates that the former contributes predominantly to electric field orientation, but the latter also should not be ignored. With the three parameters from the reverse portion, the rise and decay curves were regenerated theoretically in excellent agreement with experimental signals.  相似文献   

12.
N.m.r. and c.d. spectroscopy have been used to study the interactions of cyclic hexapeptide cyclo(Pro-Sar-Sar)2 with metal ions and ammonium ions. Cyclo(Pro-Sar-Sar)2 was found to form complexes with Li+, K?, Ba2+ and Cu2+, accompanying the conformational change into a single conformer, and the conformation of cyclo(Pro-Sar-Sar)2 in the Li+-complex was different from that in the Cu2+-complex. These findings indicate conformational flexibility of cyclo(Pro-Sar-Sar)2. The equilibrium constant for the complexation with Li+ was 2.3 × 102l mol?1, and cyclo(Pro-Sar-Sar)2 adopted an asymmetric conformation in the complex. The addition of α-amino acid ester hydrochloride also caused the conformational change of cyclo(Pro-Sar-Sar)2), but in this case it did not converge into a single conformation. This type of interaction was strengthened with aromatic α-amino acid ester hydrochloride due to the aromatic-amide interactions. Finally, the rates of exchange between unbound α-amino acid ester hydrochlorides and those complexed with cyclo(Pro-Sar-Sar)2 were found to be different, according to the nature of α-amino acid.  相似文献   

13.
The molybdenum and tungsten dinitrogen-organonitrile complexes trans-[M(N2)(NCR)(dppe)2] (2, M=Mo; 4, M=W; R=Ph, C6H4Me-p, C6H4OMe-p, Me; dppe=Ph2PCH2CH2PPh2) underwent double protonation at the nitrile carbon atom with loss of N2 and a change in oxidation state to +4 on treatment with hydrochloric acid to afford the cationic imido complexes trans-[MCl(NCH2R)(dppe)2]+. The solid-state structure of trans-[WCl(NCH2CH3)(dppe)2][PF6]·CH2Cl2 was determined by single-crystal X-ray analysis. Protonation of complexes 2 by fluoroboric acid or hydrobromic acid also formed the similar imido complexes trans-[MoX(NCH2R)(dppe)2]+ (X=F, Br). In contrast, the dinitrogen complex trans-[Mo(N2)2(dppe)2] reacted with two equiv. of benzoylacetonitrile, a nitrile with acidic CH hydrogen atoms, to give the nitrido complex trans-[Mo(N)(NKCCHCOPh)(dppe)2] (12), which was accompanied by evolution of dinitrogen and the formation of 1-phenyl-2-propen-1-one in high yields. For complex 12, the zwitterionic structure, where the anionic enolate ligand PhC(O+)=CHCN coordinates to the cationic Mo(IV) center through its nitrogen atom, was confirmed by spectroscopic measurements and single-crystal X-ray analysis. A unique intermolecular aromatic C---HO hydrogen bonding was observed in that crystal structure. Complex 12 is considered to be formed via the cleavage of the CN triple bond of benzoylacetonitrile on the metal. A reaction mechanism is proposed, which includes the double protonation of the nitrile carbon atom of the ligating benzoylacetonitrile on a low-valent molybdenum center.  相似文献   

14.
In liver homogenate the biosynthesis ofN-acetylneuraminic acid usingN-acetylglucosamine as precursor can be followed stepwise by applying different chromatographic procedures. In this cell-free system 16 metal ions (Zn2+, Mn2+, La3+, Co2+, Cu2+, Hg2+, VO 3 , Pb2+, Ce3+, Cd2+, Fe2+, Fe3+, Al3+, Sn2+, Cs+ and Li+) and the selenium compounds, selenium(IV) oxide and sodium selenite, have been checked with respect to their ability to influence a single or possible several steps of the biosynthesis ofN-acetylneuraminic acid. It could be shown that the following enzymes are sensitive to these metal ions (usually applied at a concentration of 1 mmoll–1):N-acetylglucosamine kinase (inhibited by Zn2+ and vandate), UDP-N-acetylglucosamine-2-epimerase (inhibited by zn2+, Co2+, Cu2+, Hg2+, VO 3 , Pb2+, Cd2+, Fe3+, Cs+, Li+, selenium(IV) oxide and selenite), andN-acetylmannosamine kinase (inhibited by Zn2+, Cu2+, Cd2+, and Co2+). Dose dependent measurements have shown that Zn2+, Cu2+ and selenite are more efficient inhibitors of UDP-N-acetylglucosamine-2-epimerase than vanadate. As for theN-acetylmannosamine kinase inhibition, a decreasing inhibitory effect exists in the following order Zn2+, Cd2+, Co2+ and Cu2+. In contrast, La3+, Al3+ and Mn2+ (1 mmoll–1) did not interfere with the biosynthesis ofN-acetylneuraminic acid. Thus, the conclusion that the inhibitory effect of the metal ions investigated cannot be regarded as simply unspecific is justified.Dedicated to Professor Theodor Günther on the occasion of his 60th birthday  相似文献   

15.
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   

16.
Synthetic peptides of structure cyclo(Glu(OBz)-Sar-Gly-(N-R)Gly)2 (I), electrogenic Ca2+-selective carriers in phospholipid vesicle membranes, are shown to mediate the uptake of Ca2+ ions into the cytoplasm of dog and human lymphocytes. Ca2+ transport by DECYL-2E (I, R = n-decyl) - monitored by measurements of the fluorescence of an intracellular dye, quin-2 - occurred at a rate comparable to that produced by electroneutral Ca2+ ionophores ionomycin and Br-A23187. Fluorescence quenching experiments using Mn2+ suggested a greater selectivity by DECYL-2E for Ca2+/Mn2+ vs. the other two ionophores. The result that Ca2+ ions can traverse biological membranes bound in a neutral cavity consisting exclusively of peptide carbonyl ligands may imply the functional significance of binding sites of similar structures in membrane transport proteins.  相似文献   

17.
Poly(γ-benzyl-L -glutamate) having a terminal dimethylaminoanilide group as an electron donor (D) and a terminal 3,5-dinitrobenzoyl group as an electron acceptor (A) (A-[Glu(OBzl)]n-D) was synthesized by the N-carboxyanhydride method. Polymer samples were fractionated by gel chromatography and their number-average degrees of polymerization n were determined by the absorbances of the terminal chromophores. These polymers in chloroform and dimethylformamide solutions showed a charge-transfer (CT) absorption band around 455 nm, and the fraction of the polymer forming the CT complex was evaluated as a function of the chain length. CT absorption for short chains (n = 5 ~ 20) was attributed to intramolecular CT complex in which the A-[Glu(OBzl)]n-D chain takes cyclic conformations. An optimum chain length for the intramolecular CT was found to be n ? 10, where the [Glu(OBzl)]n chain may most easily bend back to form cyclic conformations. Stronger CT absorption observed for longer chains than n = 20 was shown to be intermolecular, and an intermolecular head-to-tail aggregation was found to be a cause of the strong CT interaction. All helical A-[Glu(OBzl)]n-D chains were found to form the head-to-tail dimers in chloroform solution.  相似文献   

18.
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   

19.
Zn2+‐responsive repressor ZiaR and Co2+‐responsive activator CoaR modulate production of P1‐type Zn2+‐ (ZiaA) and Co2+‐ (CoaT) ATPases respectively. What dictates metal selectivity? We show that Δ ziaΔcoa double mutants had similar Zn2+ resistance to Δzia single mutants and similar Co2+ resistance to Δcoa single mutants. Controlling either ziaA or coaT with opposing regulators restored no resistance to metals sensed by the regulators, but coincident replacement of the deduced cytosolic amino‐terminal domain CoaTN with ZiaAN (in ziaRp ziaAziaANcoaT) conferred Zn2+ resistance to ΔziaΔcoa, Zn2+ content was lowered and residual Co2+ resistance lost. Metal‐dependent molar absorptivity under anaerobic conditions revealed that purified ZiaAN binds Co2+ in a pseudotetrahedral two‐thiol site, and Co2+ was displaced by Zn2+. Thus, the amino‐terminal domain of ZiaA inverts the metals exported by zinc‐regulated CoaT from Co2+ to Zn2+, and this correlates simplistically with metal‐binding preferences; KZiaAN Zn2+ tighter than Co2+. However, Zn2+ did not bleach Cu+‐ZiaAN, and only Cu+ co‐migrated with ZiaAN after competitive binding versus Zn2+. Bacterial two‐hybrid assays that detected interaction between the Cu+‐metallochaperone Atx1 and the amino‐terminal domain of Cu+‐transporter PacSN detected no interaction with the analogous, deduced, ferredoxin‐fold subdomain of ZiaAN. Provided that there is no freely exchangeable cytosolic Cu+, restricted contact with the Cu+‐metallochaperone can impose a barrier impairing the formation of otherwise favoured Cu+–ZiaAN complexes.  相似文献   

20.
Summary 1. While intracellular calcium concentrations are closely regulated, two types of ion channels in neurons allow calcium influx: both voltage-activated and NMDA-activated channels are significantly permeable to calcium. In this study we compare the effects of lead (Pb2+) on currents carried through voltage-activated calcium channels and NMDA-activated channels.2. Pb2+ reduces voltage-activated calcium channel currents elicited by a voltage jump from –80 to 0 mV at 0.1 to 1 µM, with an IC50 of 0.64 µM and a Hill slope of 1.22. This effect was partially reversible and not voltage dependent. Sodium and potassium currents were relatively unaffected at Pb2+ concentrations sufficient to block calcium channel currents by more than 80%. Pb2+ is, thus, a potent, reversible and selective blocker of voltage-dependent calcium channel currents.3. A fast reversible and slow irreversible blocking action of Pb2+ was found on NMDA-activated currents. When Pb2+ was applied simultaneously with aspartate and glycine (Asp/Gly), the inward currents were rapidly and reversibly reduced in a dose-dependent manner with a minimum effective concentration below 2 µM and a total blockade (>80%) with 100 µM Pb2+. The IC50 was 45 µM and the Hill coefficient 1.1. Preincubation with 50 µM Pb2+ resulted in a greater reduction in the response to Asp/Gly/Pb2+. This effect was reversed within 2 to 5 sec of wash. The lack of voltage dependence suggests that Pb2+ does not block the channel but rather alters the binding of agonists. Prolonged superfusion of a cell with the Asp/Gly/Pb2+-containing external solution resulted in a slow and irreversible decrease in the Asp/Gly activated current. No clear threshold concentration is found for this slow and irreversible effect of Pb2+. This slow action might be more important for neurotoxic effects of Pb2+.  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