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 0 毫秒
1.
Accurate rate constants were calculated from HPLC kinetic measurements of the selective acidolysis of the C-terminal amide bond of eight N-acyl-N-(4-methoxybenzyl)-alpha,alpha-trialkyl glycine amides in TFA at 25.00 degrees C. The results were in all cases consistent with a first order behaviour with respect to the substrate and, apparently, also to the acid, and a clear relationship between reactivity and structure could be observed. The data collected also allowed experimental evidence to be obtained for the first time in support of the previously postulated formation of an intermediate oxazolonium salt. In the case of the more crowded species this intermediate compound undergoes slow hydrolytic ring opening, which takes place in competition with cleavage of the N-alkyl group to give another oxazolonium derivative that hydrolysed still more slowly. The stability of the intermediate cyclic compounds may result either from conjugation of the phenyl group with the oxazolonium ring in the case of N-benzoyl derivatives, or from conformational assistance imparted by the bulky amino acid side chains of the alpha,alpha-dialkyl glycine species, or both. The loss of the N-alkyl group also seems to be assisted by the bulkiness of the amino acid side chains, which thus tends to decrease the selectivity of cleavage.  相似文献   

2.
The C-terminal amide bond of N-acyl-N,alpha,alpha-trialkyl glycine amides is labile to acid and this has been currently assigned to steric crowding within the amino acid residue. However, our previous work has shown that in the acidolysis of some of these compounds steric hindrance seems to play a less important role than what one would expect. Thus, the cleavage of two sets of such compounds bearing different degrees of crowding was investigated at five different temperatures in order to clarify the effect of structure on reactivity in terms of enthalpy and entropy of activation. The compounds exhibited an Arrhenius-type behaviour, and both enthalpies and entropies of activation were calculated by taking advantage of the transition state theory. In addition, the kinetic data were analysed in terms of isokinetic relationships in order to find evidence to support that the compounds react under the same mechanism. The changes in the reaction rate are governed by the changes in both the enthalpy and the entropy of activation, which are related to bond energy and steric hindrance, respectively. In general, the entropies of activation are very negative for all compounds investigated, which reflects large steric constrictions associated with the formation of the transition state. In addition, they are very sensitive to the structure of the substrates.  相似文献   

3.
The synthesis of glyoxylyl peptides by coupling the masked glyoxylic acid derivative (FmocNH)(2)CHCO(2)H, 1, to a peptidyl resin assembled using Fmoc/tert-butyl chemistry has been described recently. Deprotection and cleavage of the peptide from the solid support using TFA was followed by unmasking of the glyoxylyl group in solution in the presence of DBU. [] The glyoxylyl peptide was thus generated using non-oxidizing conditions by comparison with the method based on the periodic oxidation of a seryl-precursor. However, base treatment of the (FmocNH)(2)CHCO(2)-peptide led to the formation of a byproduct besides the desired glyoxylyl peptide. This paper describes an optimized procedure for unmasking the Fmoc-protected alpha,alpha'-diaminoacetic acid moiety in solution which suppressed byproduct formation. Also presented is a series of experiments that permitted a structure and a mechanism of formation for the byproduct to be suggested.  相似文献   

4.
The crystal structures of two diastereomeric alpha,beta-dehydrobutyrine peptides Ac-Pro-(Z)-DeltaAbu-NHMe (I) and Ac-Pro-(E)-DeltaAbu-NHMe (II) have been determined. Both dehydropeptides adopt betaI-turn conformation characterized by the pairs of (phi(i+1), psi(i+1)) and (phi(i+2), psi(i+2)) angles as -66, -19, -97, 11 degrees for I and -59, -27, -119, 29 degrees for II. In each peptide, the betaI turn is stabilized by (i + 3) --> i intramolecular hydrogen bonds with N...O distance of 3.12 A for I and 2.93 A for II. These structures have been compared to the crystal structures of homologous peptides Ac-Pro-DeltaVal-NHMe and Ac-Pro-DeltaAla-NHMe. Theoretical analyses by DFT/B3LYP/6-31 + G** method of conformers formed by these four peptides and by the saturated peptide Ac-Pro-Ala-NHMe revealed that peptides with a (Z) substituent at the C(beta) (i+2) atom of dehydroamino acid, i.e. Ac-Pro-DeltaVal-NHMe and Ac-Pro-(Z)-DeltaAbu-NHMe, predominantly form beta turns, both in vacuo and in polar environment. The tendency to adopt beta-turn conformation is much weaker for the peptides lacking the (Z) substituent, Ac-Pro-(E)-DeltaAbu-NHMe and Ac-Pro-DeltaAla-NHMe. The latter adopts a semi-extended or an extended conformation in every polar environment, including a weakly polar solvent. The saturated peptide Ac-Pro-Ala-NHMe in vacuo prefers a beta-turn conformation, but in polar environment the differences between various conformers are small. The role of pi-electron correlation and intramolecular hydrogen bonds interaction in stabilizing the hairpin structures are discussed.  相似文献   

5.
Peptide foldamers based on alpha,alpha-disubstituted glycyl residues were synthesized and chemically characterized to investigate the effects of the electric field generated by a 3(10)-helix on the rate of intramolecular photoinduced electron-transfer reactions. To this end, two new octapeptides having identical sequences were suitably side-chain functionalized with the same electron-transfer donor-acceptor pair, but inverting the position of the pair along the main chain. The electron-transfer rate constants, measured by time-resolved spectroscopy techniques (nanosecond transient absorption and time-resolved fluorescence), indicated that, in the case of the 3(10)-helix, the electrostatic effect is significant, but smaller than that obtained for alpha-helical peptides. This finding can be likely ascribed to the distortion of the H-bond network with respect to the helical axis taking place in the former secondary structure. Overall, these results could have implications on electron-transfer phenomena in model and biomembranes facilitated by peptaibiotics.  相似文献   

6.
A series of three homologous dimethyldiamides Ac-(Z)-deltaPhe-NMe2, Ac-L-Phe-NMe2 and Ac-DL-Phe-NMe2 have been synthesized and their structures determined from single-crystal X-ray diffraction data. To learn more about the conformational preferences of the compounds studied, the fully relaxed phi, psi conformational energy maps on the free molecules of Ac-deltaAla-NMe2 and Ac-(Z)-deltaPhe-NMe2 were obtained with the HF/3-21G method and the calculated minima re-optimized with the DFT/B3LYP/6-31G** method. The crystal state results have been compared with the literature data. The studied dimethyldiamide Ac-deltaXaa-NMe2 combines the double bond in positions alpha, beta and the C-terminal tertiary amide within one molecule. As the representative probe with deltaXaa = deltaAla, (Z)-deltaLeu and (Z)-deltaPhe shows, in the solid state they adopt the conservative conformation with phi, psi approximately -45 degrees, approximately 130 degrees and with a non-planar tertiary amide bond, whatever the packing forces are. This conformation is located on the Ramachandran map in region H/F, which is of high-energy for common amino acids, but not so readily accessible to them. The free molecule calculations on Ac-deltaAla-NMe2 and Ac-(Z)-deltaPhe-NMe2 reveal that, in spite of dissimilar overall conformational profiles of these molecules, this structure is one of their low-energy conformers and for Ac-(Z)-deltaPhe-NMe2 it constitutes the global minimum. So, the theoretical results corroborate those experimental results proving that this structure is robust enough to avoid conformational distortion due to packing forces. In contrast to Ac-deltaXaa-NMe2, the saturated Ac-L/DL-Xaa-NMe2 shows the constancy of the associative patterns but do not prefer any molecular structure in the solid state.  相似文献   

7.
Retinoid acid receptors (RXR-alpha, -beta, -gamma) and Farnesoid X-activated receptor (FXR) expression in the testis of the marbled newt were investigated with special attention to the changes during the annual testicular cycle, using light microscopy immunohistochemistry and Western blot analysis. The annual testicular cycle of the marbled newt (Triturus marmoratus marmoratus) comprises three periods: (a) proliferative period (germ cell proliferation from primordial germ cells to round spermatids, April-June); (b) spermiogenesis period (July-September); and (c) quiescence period (interstitial and follicular cells form the glandular tissue, October-April). In the proliferative period, primordial germ cells and primary spermatogonia immunostained intensely to the three types of RXRs and also to FXR. In the other periods, immunostaining to these antibodies was weak or absent. Secondary spermatogonia stained weakly to the four antibodies in the proliferative period, and only to FXR, also weakly, in the spermiogenesis period. Immunoreactive primary spermatocytes were weakly labeled with the RXR antibodies in the proliferative period. Spermatids and spermatozoa did not stain to any antibody in any period. Follicular cells only immunostained to RXR-gamma and only in the quiescence period when they are forming the glandular tissue, together with the interstitial cells. As follicular cells, interstitial cells only immunostained in the quiescence period; however, they immunoreacted to the three types of RXRs. These findings suggest that in the newt, RXRs and FXR are involved in spermatogenesis control by regulating the proliferation of primordial germ cells and spermatogonia. In addition, RXR-gamma seems to be also involved in the development of the glandular (steroidogenic) tissue.  相似文献   

8.
Ultraviolet light is the major cause of skin cancers in human, and several effects of ultraviolet light B (UVB) are thought to contribute to skin photocarcinogenesis. 3Beta-methoxy-13alpha,14alpha-epoxyserratan-21beta-ol (PJJ-34; 1) isolated from Picea jezoensis Carr. var. jezoensis showed the strongest antitumor-promoting activity among naturally occurring triterpenoids in the in vivo two-stage mouse skin-carcinogenesis test. To investigate the anti-initiating activity, we further studied mouse models initiated with ultraviolet-B (UVB) and promoted with 12-O-tetradecanoylphorbol-13-acetate (TPA). Oral administration of the PJJ-34 (1) during a period before and after the three times of UVB irradiation led to a remarkable effect: oral administration of a 0.0025% solution of 1 only to the test group, which started one week before and ended one week after the irradiation, showed less than half papillomas, and inhibition of tumor incidence and tumor multiplicity in comparison to the control group. Therefore, it was recognized that PJJ-34 (1) showed strong anti-initiating activity as well as anti-promoting activity. After all, 1 seems to be useful as cancer-chemopreventive agent.  相似文献   

9.
The FTIR spectra were analysed in the region of the nu(s)(N-H), AI(C=O) and nu(s)(Calpha=Cbeta) bands for a series of Ac-DeltaXaa-NMe2, where DeltaXaa = DeltaAla, (Z)-DeltaAbu, (Z)-DeltaLeu, (Z)-DeltaPhe and DeltaVal, to determine a predominant solution conformation of these alpha,beta-dehydropeptide-related molecules. Measurements were taken in CCl4, DCM and MeCN solutions. In the same way, spectra of saturated analogues Ac-Xaa-NMe2, where Xaa = Ala, Abu, Leu, Phe and Val, were investigated. To help interpret the spectroscopic results, conformational maps were calculated by the B3LYP/6-31+G** method. Also, the relative energies of all conformers of the dehydro compounds in vacuo as well as in the studied solvents in addition to the theoretical IR frequencies of these conformers were calculated. For comparison, molecules of two saturated analogues, Ac-L-Ala-NMe2 and Ac-L-Phe-NMe2, were calculated in a similar way. Both unsaturated and saturated compounds, which have an aliphatic side chain, occur in CCl4 and DCM mainly as a mixture of extended conformers with the C5 H-bond and open conformers. As solvent polarity increases, participation of the open conformers also increases, and in MeCN, the model amides are almost exclusively in the open form, except Ac-DeltaAla-NMe2, which shows a small amount of the H-bonded conformer. Ac-DeltaAla-NMe2 and Ac-DeltaAbu-NMe2 have stronger C5 hydrogen bonds than those of their saturated counterparts. As the calculations indicate, the open conformation of the unsaturated amides is conformer H/F with phi, psi -44 +/- 5 degrees, 127 +/- 4 degrees. This is the second lowest in energy conformer in vacuo and in CCl4 and the lowest one in more polar solvents. The open conformation of Ac-L-Ala-NMe2 constitutes conformer C with phi, psi -101.5 degrees, 112.7 degrees. For Ac-DeltaAla-NMe2 and Ac-DeltaAbu-NMe2, FTIR also reveals the presence of a third conformer. Calculations indicate that is the semiextended conformer D with the N1-H1...N2 hydrogen bond/contact. In all solvents, Ac-L-Phe-NMe2 and Ac-(Z)-DeltaPhe-NMe2 show only the extended E and the open H/F, respectively. In both there is an amide/pi(Ph) interaction.  相似文献   

10.
N-Acetyl-(E)-dehydrophenylalanine N'-methylamide [Ac-(E)-DeltaPhe-NHMe], one of a few representative (E)-alpha,beta-dehydroamino acids, was studied by FTIR in dichloromethane and acetonitrile. To support spectroscopic interpretations and to gain some deeper insight into the Ac-(E)-DeltaPhe-NHMe molecule, the Ramachandran potential energy surface was calculated by the B3LYP/6-31G*//HF/3-21G method and the conformers localized were fully optimized at the B3LYP/6-31 + G** level. The spectra and calculations were compared with those of the related molecules Ac-DeltaAla-NHMe and Ac-(Z)-DeltaPhe-NHMe. The title compound assumes two conformational states in equilibrium in dichloromethane solution with a predominance of the extended conformer E. The Ac-(E)-DeltaPhe-NHMe spectrum is like that of Ac-DeltaAla-NHMe, particularly in the region of bands AI and AII, and unlike that of Ac-(Z)-DeltaPhe-NHMe. The positions of bands AI and II together with the nu(s)(N1--H1) band proves that the conformers E of both DeltaAla and (E)-DeltaPhe compounds are stabilized by the quite strong C5 hydrogen bonds N1--H1...O2. The same conclusion is drawn from the Ramachandran diagrams. The conformers E of both compounds are placed in the global minima and the gaps in energy order between them and the second conformer are large. The conformers E of DeltaAla and (E)-DeltaPhe, apart from the N1--H1...O2 hydrogen bond, show the Cbeta--H...O1 interaction, and Ac-(E)-DeltaPhe-NHMe displays the NH/pi interaction with the N2--H2 projecting in the first carbon atom of the phenyl ring. The C5 hydrogen bond is stronger in (E)-DeltaPhe than that in the DeltaAla compound. This is in agreement with interactions found in the calculated structures and can be explained by the influence of the phenyl ring in position (E). In acetonitrile, the molecule of Ac-(E)-DeltaPhe-NHMe loses its C5 hydrogen bond and becomes unfolded, whereas that of Ac-DeltaAla-NHMe does not vary practically. Adopting conformation E in a non-polar solvent seems to be a general feature of the (E)-DeltaXaa residues.  相似文献   

11.
Fifteen years ago it was shown that an alpha-aminoisobutyric acid (Aib) residue is significantly more effective than an L-Pro or a D-amino acid residue in inducing beta-sheet disruption in short model peptides. As this secondary structure element is known to play a crucial role in the neuropathology of Alzheimer's disease, it was decided to check the effect of Aib (and other selected, helix inducer, C(alpha)-tetrasubstituted alpha-amino acids) on the beta-sheet conformation adopted by a protected pentapeptide related to the sequence 17-21 of the beta-amyloid peptide. By use of FT-IR absorption and 1H NMR techniques it was found that the strong self-association characterizing the pentapeptide molecules in weakly polar organic solvents is completely abolished by replacing a single residue with Aib or one of its congeners.  相似文献   

12.
In the present study, the chemopreventive effect of the active metabolite of vitamin D, 1alpha,25-dihydroxyvitamin D(3) (VD(3)), against chemically-induced and diabetes-promoted rat liver carcinogenesis was investigated. Hepatocarcinogenesis was initiated with a single intraperitoneal (i.p.) injection of diethylnitrosamine (DEN) (125 mg kg(-1) body weight) at week 4 followed by promotion with streptozotocin (STZ) (65 mg kg(-1) body weight with a single i.p. injection) at week 7. With this basic experimental regimen, the effect of VD(3) (0.3 microg (0.1 ml)(-1) propylene glycol per os twice a week) was investigated with effect from 4 weeks prior to the exposure of DEN. The results showed that VD(3) supplementation throughout the experimental period reduced the incidence, total number and multiplicity and altered the size of visible persistent nodules (PNs) in DEN- or DEN + STZ-treated rats as compared with their respective controls. In these two groups, it also caused a significant decrease in the number (p < 0.002 and 0.001 respectively) and focal area (p < 0.05) of gamma-glutamyltranspeptidase (GGT)-positive hepatic foci. Moreover, continuous supplementation of VD(3) exhibits a protective effect in maintaining the normal cellular architecture of the hepatocytes in DEN- or DEN + STZ-treated rats. Our results thus strongly suggest that VD(3) is very effective in the inhibition of DEN-initiated and STZ-induced diabetes-promoted rat liver carcinogenesis.  相似文献   

13.
A collection of circularly permuted catalytic chains of aspartate transcarbamoylase (ATCase) has been generated by random circular permutation of the pyrB gene. From the library of ATCases containing permuted polypeptide chains, we have chosen for further investigation nine ATCase variants whose catalytic chains have termini located within or close to an alpha helix. All of the variants fold and assemble into dodecameric holoenzymes with similar sedimentation coefficients and slightly reduced thermal stabilities. Those variants disrupted within three different helical regions in the wild-type structure show no detectable enzyme activity and no apparent binding of the bisubstrate analog N:-phosphonacetyl-L-aspartate. In contrast, two variants whose termini are just within or adjacent to other alpha helices are catalytically active and allosteric. As expected, helical disruptions are more destabilizing than loop disruptions. Nonetheless, some catalytic chains lacking continuity within helical regions can assemble into stable holoenzymes comprising six catalytic and six regulatory chains. For seven of the variants, continuity within the helices in the catalytic chains is important for enzyme activity but not necessary for proper folding, assembly, and stability of the holoenzyme.  相似文献   

14.
The protein kinase CK2 (former name: "casein kinase 2") predominantly occurs as a heterotetrameric holoenzyme composed of two catalytic chains (CK2alpha) and two noncatalytic subunits (CK2beta). The CK2beta subunits form a stable dimer to which the CK2alpha monomers are attached independently. In contrast to the cyclins in the case of the cyclin-dependent kinases CK2beta is no on-switch of CK2alpha; rather the formation of the CK2 holoenzyme is accompanied with an overall change of the enzyme's profile including a modulation of the substrate specificity, an increase of the thermostability, and an allocation of docking sites for membranes and other proteins. In this study we used C-terminal deletion variants of human CK2alpha and CK2beta that were enzymologically fully competent and in particular able to form a heterotetrameric holoenzyme. With differential scanning calorimetry (DSC) we confirmed the strong thermostabilization effect of CK2alpha on CK2beta with an upshift of the CK2alpha melting temperature of more than 9 degrees . Using isothermal titration calorimetry (ITC) we measured a dissociation constant of 12.6 nM. This high affinity between CK2alpha and CK2beta is mainly caused by enthalpic rather than entropic contributions. Finally, we determined a crystal structure of the CK2beta construct to 2.8 A resolution and revealed by structural comparisons with the CK2 holoenzyme structure that the CK2beta conformation is largely conserved upon association with CK2alpha, whereas the latter undergoes significant structural adaptations of its backbone.  相似文献   

15.
The first step of polyamine uptake is the binding of polyamines to the cell membrane. In order to characterize the specificity of the putrescine binding sites at the surface of the glioblastoma cells (U251), we have carried out competition experiments between putrescine bound to latex microspheres and vizualized by scanning electron microscopy and a series of N,N'-tetramethyl-alpha,omega-diaminoalkanes. N,N'-tetramethyl-1,4-butanediamine (N,N'-tetramethylputrescine) and higher homologs inhibit the latex putrescine binding to the cell surface and concomitantly cell proliferation. [14C] putrescine uptake was mainly inhibited by the lower homologs, which were devoid of antiproliferative effects. Our results suggest that putrescine uptake by the human glioblastoma cell line U251, and putrescine binding to the surface of these cells are independent processes. The potential relationship between antitumor effect of N,N'-tetramethyl-alpha,omega-diaminoalkanes and its binding to a specific putrescine acceptor site is discussed.  相似文献   

16.
We present a molecular dynamics simulation at 300 K in explicit solvent environment of chloroform of the stoichiometric complex formed by poly(alpha,L-glutamate) and octyltrimethylammonium ions. We observed that the alpha-helix conformation of the polypeptide chain remains stable during a 2-ns run. The surfactant ions predominantly adopted an extended conformation that is stabilized by favorable interactions with the organic solvent. Analysis of the organization of the surfactant with respect to the polypeptide chain indicated that each octyltrimethylammonium cation was preferentially bound to more than one carboxylate group. It was found that the most populated arrangement was that with the surfactant cations interacting with two carboxylate groups simultaneously.  相似文献   

17.
A series of terminally blocked peptides (to the pentamer level) from l ‐Ala and the cyclic Cα,α‐disubstituted Gly residue Afc and one Gly/Afc dipeptide have been synthesized by solution method and fully characterized. The molecular structure of the amino acid derivative Boc‐Afc‐OMe and the dipeptide Boc‐Afc‐Gly‐OMe were determined in the crystal state by X‐ray diffraction. In addition, the preferred conformation of all of the model peptides was assessed in deuterochloroform solution by FT‐IR absorption and 1H‐NMR. The experimental data favour the conclusion that the Afc residue tends to adopt either the fully‐extended (C5) or a folded/helical structure. In particular, the former conformation is highly populated in solution and is also that found in the crystal state in the two compounds investigated. A comparison with the structural propensities of the strictly related Cα,α‐disubstituted Gly residues Ac5c and Dϕg is made and the implications for the use of the Afc residue in conformationally constrained analogues of bioactive peptides are briefly examined. A spectroscopic (UV absorption, fluorescence, CD) characterization of this novel aromatic Cα,α‐disubstituted Gly residue is also reported. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
We have previously shown that one of the potential mediators of the deleterious effects of high glucose on extracellular matrix protein (ECM) expression in renal mesangial cells is its metabolic flux through the hexosamine biosynthesis pathway (HBP). Here, we investigate further whether the hexosamines induce oxidative stress, cell-cycle arrest and ECM expression using SV-40-transformed rat mesangial (MES) cells and whether the anti-oxidant alpha-lipoic acid will reverse some of these effects. Culturing renal MES cells with high glucose (HG, 25 mM) or glucosamine (GlcN, 1.5 mM) for 48 h stimulates laminin gamma1 subunit expression significantly approximately 1.5 +/- 0.2- and 1.9 +/- 0.3-fold, respectively, when compared to low glucose (LG, 5 mM). Similarly, HG and GlcN increase the level of G0/G1 cell-cycle progression factor cyclin D1 significantly approximately 1.7 +/- 0.2- and 1.4 +/- 0.04-fold, respectively, versus LG (p < 0.01 for both). Azaserine, an inhibitor of glutamine:fruc-6-PO(4) amidotransferase (GFAT) in the HBP, blocks the HG-induced expression of laminin gamma1 and cyclin D1, but not GlcN's effect because it exerts its metabolic function distal to GFAT. HG and GlcN also elevate reactive oxygen species (ROS) generation, pro-apoptotic caspase-3 activity, and lead to mesangial cell death as revealed by TUNEL and Live/Dead assays. FACS analysis of cell-cycle progression shows that the cells are arrested at G1 phase; however, they undergo cell growth and hypertrophy as the RNA/DNA ratio is significantly (p < 0.05) increased in HG or GlcN-treated cells relative to LG. The anti-oxidant alpha-lipoic acid (150 microM) reverses ROS generation and mesangial cell death induced by HG and GlcN. Alpha-lipoic acid also reduces HG and GlcN-induced laminin gamma1 and cyclin D1 expression in MES cells. In addition, induction of diabetes in rats by streptozotocin (STZ) increases both laminin gamma1 and cyclin D1 expression in the renal cortex and treatment of the diabetic rats with alpha-lipoic acid (400 mg kg(-1) body weight) reduces the level of both proteins significantly (p < 0.05) when compared to untreated diabetic rats. These results support the hypothesis that the hexosamine pathway mediates mesangial cell oxidative stress, ECM expression and apoptosis. Anti-oxidant alpha-lipoic acid reverses the effects of high glucose, hexosamine and diabetes on oxidative stress and ECM expression in mesangial cells and rat kidney.  相似文献   

19.
We report the solid‐phase synthesis and some pharmacological properties of 23 new analogs of arginine vasopressin (AVP) which have the Phe3 residue replaced by a broad variety of amino acids. Peptides 1–9 have at position 3: (1) the mixed aromatic/aliphatic amino acid thienylalanine (Thi) and the aliphatic amino acids; (2) cyclohexylalanine (Cha); (3) norleucine (Nle); (4) Leu; (5) norvaline (Nva); (6) Val; (7) alpha‐aminobutyric acid (Abu); (8) Ala; (9) Gly. Peptides 10–23 have at position 3: the aromatic amino acids, (10) homophenylalanine (Hphe); (11) Tyr; (12) Trp; (13) 2‐naphthylalanine (2‐Nal); the conformationally‐restricted amino acids (14) Pro; (15) 2‐aminotetraline‐2‐carboxylic acid (Atc); the polar amino acids (16) Ser; (17) Thr; (18) Gln; and the charged amino acids (19) Asp; (20) Glu; (21) Arg; (22) Lys; (23) Orn. All 23 new peptides were evaluated for agonistic and, where appropriate, antagonistic activities in in vivo antidiuretic (V2‐receptor) and vasopressor (V1a‐receptor) assays and in in vitro (no Mg2+) oxytocic assays. The corresponding potencies (units/mg) in these assays for AVP are: 323±16; 369±6 and 13.9±0.5. Peptides 1–9 exhibit the following potencies (units/mg) in these three assays: (1) 379±14; 360±9; 36.2±1.9; (2) 294±21; 73.4±2.7; 0.33±0.02; (3) 249±28; 84.6±4.3; 4.72±0.16; (4) 229±19; 21.4±0.6; 2.1±0.2; (5) 134±5; 31.2±0.9; 28.4±0.2; (6) 114±9; 45.3±2.3; 11.3±1.6; (7) 86.7±2.5; 4.29±0.13; 0.45±0.03; (8) 15.5±1.5; 0.16±0.01; ∼0.02; (9) 3.76±0.03; <0.02; in vitro oxytocic agonism was not detected. These data show that the aliphatic amino acids Cha, Nle, Leu, Nva and Val are well‐tolerated at position 3 in AVP with retention of surprisingly high levels of antidiuretic activity. Peptides 2–9 exhibit significant gains in both antidiuretic/vasopressor (A/P) and antidiuretic/oxytocic (A/O) selectivities relative to AVP. [Thi3]AVP appears to be a more potent antidiuretic and oxytocic agonist than AVP and is equipotent with AVP as a vasopressor agonist. The antidiuretic potencies of peptides 10–23 exhibit drastic losses relative to AVP. They range from a low of 0.018±0.001 units/mg for the Lys3 analog (peptide 22) to a high of 24.6±4.6 units/mg for the Hphe3 analog (peptide 10). Their vasopressor potencies are also drastically reduced. These range from a low of <0.002 units/mg for peptide 22 to a high of 8.99±0.44 units/mg for the Atc3 analog (peptide 15). Peptides 10–23 exhibit negligible or undetectable in vitro oxytocic agonism. The findings on peptides 10–23 show that position 3 in AVP is highly intolerant of changes with aromatic, conformationally‐restricted, polar and charged amino acids. Furthermore, these findings are in striking contrast to our recent discovery that position 3 in the potent V2/V1a/OT antagonist d(CH2)5d ‐Tyr(Et)2VAVP tolerates a broad latitude of structural change at position 3 with many of the same amino acids, to give excellent retention of antagonistic potencies. The data on peptides 1–4 offer promising clues to the design of more potent and selective AVP V2 agonists. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
The fully extended peptide conformation (2.0(5)-helix) has been investigated for the first time in the solid-state by 13C cross-polarization magic angle spinning NMR. The compounds examined are members of a terminally protected, homo-oligopeptide series (from monomer through hexamer) based on Calpha,beta-didehydroalanine.  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