首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 534 毫秒
1.
In earlier reports we identified specific point substitutions (DPhe12,Nle21,38), cyclization strategies [in particular, introduction of lactam rings such as that of cyclo(Glu30,Lys33)], and deletions (residues 1-7) in the CRF molecule that led to agonists. We also noted that further deletions (residues 8-14) produced antagonists such as astressin ?cyclo(30-33)[DPhe12,Nle21,38, Glu30, Lys33]hCRF(12-41)? (1). We hypothesized that the lactam ring promoted conformational stability to yield analogues with increased potency both in vitro and in vivo as compared to that of their linear counterparts. Additionally, we reported that cyclo(30-33)[DPhe12,Nle21,38, Glu30,DHis32,Lys33]hCRF(12-41) (3) and dicyclo(26-36,30-33)[Ac-Asp9,DPhe12,Nle21,38, Cys26, Glu30,Lys33, Cys36]hCRF(9-41) were ca. twice and 1/100 as potent as astressin, respectively, suggesting a putative turn that encompasses residues 30-33 (previous paper: Koerber et al. J. Med. Chem. 1998, 41). To increase the potency of 1 and/or 3 in vivo, we extended their chain length by one (5-8), two (9, 10), and three (11, 12) residues at the N-terminus and acetylated (6, 8, 10, 12). Of the compounds tested for duration of action (1, 3-6, 8), we found 6 and 8 to be slightly longer-acting than astressin or [DHis32]astressin, while their potencies in vitro were not significantly different from that of 3. Additionally, we introduced CalphaMe-leucine residues in lieu of leucine at positions 14, 15, 19, 27, and 37 in [DHis32]astressin. The analogue [CalphaMe-Leu27,DHis32]astressin (16) was more potent (although not statistically in all cases) than the other four analogues in vitro. While acetylation of the N-terminus of 16 (i.e., 18) or of [CalphaMe-Leu27]astressin (i.e., 19) did not have a significant effect on in vitro potency, elongation of the N-terminus by one or three residues in addition to acetylation resulted in cyclo(30-33)[DPhe12,Nle21,CalphaMe-Leu27,Glu3 0,DHis32,Lys33, Nle38]Ac-hCRF(11-41) (21), cyclo(30-33)[DPhe12,Nle21,CalphaMe-Leu27, Glu30,Lys33,Nle38]Ac-hCRF(9-41) (22), and cyclo(30-33)[DPhe12, Nle21, CalphaMe-Leu27,Glu30,DHis32,Lys33,Nle38 ]Ac-hCRF(9-41) (23) that were longer-acting than 6 and 8 (ca. 2 h inhibition of ACTH secretion at 25 micrograms/adrenalectomized rat). Analogues 22 and 23 were also more potent than astressin at reversing intracisternal CRF- and abdominal surgery-induced delay of gastric emptying in conscious rats.  相似文献   

2.
In the present study, human growth hormone-releasing factor (hGRF) and analogs were successfully pegylated at the carboxy-terminus using a novel solid- and solution-phase strategy. Following synthesis, these pegylated hGRF analogs were evaluated for in vitro and in vivo biological activity. Specifically, hGRF (1-29)-NH2, [Ala15]-hGRF (1-29)-NH2, [desNH2Tyr1, D-Ala2, Ala15]-hGRF(1-29)-NH2 and [His1, Val2, Gln8, Ala15, Leu27]-hGRF(1-32)-OH were each C-terminally extended using a Gly-Gly-Cys-NH2 spacer (previously demonstrated not to alter intrinsic biological activity), and then monopegylated via coupling to an activated dithiopyridyl-PEG reagent. PEG moieties of 750, 2000, 5000 or 10,000 molecular weight (MW) were examined to determine the effect of polymer weight on activity. Initial biological evaluations in vitro revealed that all C-terminally pegylated hGRF analogs retained high growth hormone (GH)-releasing potencies, regardless of the MW of PEG polymer employed. Two of these pegylated hGRF analogs, [desNH2Tyr1, D-Ala2, Ala15]-hGRF (1-29)-Gly-Gly-Cys(NH2)-S-Nle-PEG5000 and [His1, Val2, Gln8, Ala15, Leu27]-hGRF(1-32)-Gly-Cys(NH2)-S-Nle-PEG5000, were subsequently evaluated in both pig and mouse models and found to be highly potent (in vivo potency range = 12-55-fold that of native hGRF). Relative to their non-pegylated counterparts, these two pegylated hGRF analogs exhibited enhanced duration of activity.  相似文献   

3.
Conditions have been developed for the site-specific pegylation (NH2-terminus, side-chain and carboxy-terminus) of a potent analog of growth hormone-releasing factor, [Ala15]-hGRF(1-29)-NH2. These pegylated peptides were prepared by solid-phase peptide synthesis using the Fmoc/tBu strategy, and were fully characterized by analytical HPLC, amino-acid analysis, 1H-NMR spectroscopy and laser desorption mass spectrometry. Biological activities of hGRF analogs were determined in vitro utilizing stimulation of growth hormone release by cultured rat pituitary cells as an index. GH-releasing potencies of the pegylated hGRF analogs were compared to a series of model analogs of [Ala15]-hGRF(1-29)-NH2 that were acetylated or protected as the ethylamides at the pegylation sites. It was found that acetylation at the NH2-terminus resulted in reduced potency, which was not further affected when the NH2-terminus was pegylated, regardless of the size of poly(ethyleneglycol) (PEG) employed (e.g. PEG2000 or PEG5000). Pegylation at Asp8 or Lys12 decreased biological potency, a situation which was exacerbated by increasing the molecular weight of PEG. Pegylation at Lys21 or Asp25 did not significantly affect biological activity. The C-terminal model peptide, [Ala15,Orn(Ac)30]-hGRF(1-29)-NH2, was the most potent analog identified in this series (ca. 4-5-fold that of hGRF(1-44)-NH2. The COOH-terminal pegylated analogs retained this increased level of biological activity independent of PEG molecular weight. These studies demonstrate that a biologically active peptide can be pegylated and retain the full in vitro potency of the peptide. However, the biological activity is highly dependent on the site of pegylation and, in some cases, the molecular weight of PEG (degree of pegylation) moiety used.  相似文献   

4.
Conformational searching, computer simulations, synthesis and NMR are used on a variety of alpha melanocyte-stimulating hormone (alpha-MSH) analogues to understand the physical characteristics required for biological potency. Peptides I (Ac-[Nle4,Asp5,D-Phe7,Lys10]alpha-MSH(4-10)-NH2), II (Ac-c[Nle4,Asp5,D-Phe7,Lys10]alpha-MSH(4-10)-NH2) and III (Ac-[Nle4,Asp5,D-Phe7,Dap10]alpha-MSH(4-10)-NH2 all show very similar conformational properties (backbone and side-chain torsional angles), and all display high biological potencies. The modeling results for these compounds are supported by the NMR data. Peptide IV (Ac-c[Nle4,Asp5,D-Phe7,Dap10]alpha-MSH(4-10)-NH2) appears to have a markedly different conformation and has decreased biological potency.  相似文献   

5.
The octadecaneuropeptide ODN (QATVGDVNTDRPGLLDLK), originally characterized as an endogenous ligand for central-type benzodiazepine receptors, increases intracellular calcium concentration ([Ca2+]i) in rat astroglial cells. A series of ODN analogues was synthesized, and each compound was studied for its ability to induce Ca2+ mobilization in cultured rat astrocytes. Replacement of each amino acid by an L-alanine residue (AlaScan) showed that the N-terminal region of the molecule was relatively tolerant to alanine substitution (2-8, 10), except for the Ala9-substituted analogue (9) which was totally devoid of activity. Pyroglutamization (21) and acetylation (22) of the Gln1 residue reduced the Ca2+ response suggesting that a free N-terminal amine function is required for full activity of ODN. Alanine substitution of the residues in the C-terminal region of the molecule (11-14, 16-18) significantly reduced the biological activity of ODN. In particular, modifications of the Leu15 residue (15, 20) abolished the Ca2+-mobilizing activity. The analogues [Ala9]ODN (9), [Ala15]ODN (15), [D-Thr9]ODN (19), and [D-Leu15]ODN (20) partially antagonized the Ca2+ response evoked by ODN. Most importantly, the octapeptide ODN11-18 (OP, 24) produced a dose-response curve that was superimposable to that obtained with ODN, indicating that the C-terminal region of the molecule possesses full biological activity. Finally, the AlaScan of OP revealed that replacement of the Leu5 residue by Ala (29) or D-Leu (33) totally suppressed the calcium response, confirming the crucial contribution of the Leu15 residue of ODN to the biological activity of the neuropeptide.  相似文献   

6.
Two series of CRF antagonists with N alpha- and C alpha-methylated alanine and leucines were evaluated for their biological activities in vitro and in vivo in several systems. The poly-N-methylated analogue of alpha-helical-CRF9-41, [N alpha MeLeu10,15,27,37,N alpha MeAla22,32,41]-alpha-Hel-CRF9-41, was found to be considerably less potent than the parent non-N-methylated analogue. This result was expected on the basis that alpha-helicity was thought to be required for biological activity and the prediction that backbone substitutions on the nitrogen have a tendency to break alpha-helices (a hypothesis that was confirmed by circular dichroism). Next, a series of constrained analogues of the potent CRF antagonist, [DPhe12,Nle21,38]h/rCRF12-41, was synthesized that contained C alpha-methylleucine and/or C alpha-methylalanine (Aib) residues at selected positions. Because C alpha-methylation is recognized to increase alpha-helicity, and because there is now strong NMR data suggesting that residues 6-36 assume a well-defined alpha-helix, it was expected that these analogues would be more potent. Although usual solid-phase peptide synthesis procedures were followed, success in coupling the C alpha-methyl amino acids was obtained only with a 1:1 mixture of BOP/HOBt. In vitro potencies of the synthesized compounds were measured in a collagenase-dispersed anterior pituitary cell culture bioassay. Monosubstituted analogues were shown to be twice to one fourth as potent as the parent compound; while the pluri-substituted peptides were slightly less potent. This decrease in potency might be correlated to an unexpected lower helical content of the pluri-substituted compounds (as determined by CD spectroscopy), as it was suggested that the bioactive conformation of the CRF was predominantly alpha-helical. Interestingly, one analogue, [DPhe12,Nle21,38,C alpha-MeLeu37]h/rCRF12-41, was found to be more potent and longer acting than the parent compound in two in vivo assays measuring ACTH release after intravenous administration to adrenalectomized rats and reversal of stress-induced delay in gastric emptying in the rat after intracisternal administration. The molecular basis for this increased duration of action and potency is being investigated.  相似文献   

7.
Hybrid peptides were constructed from endothelin B receptor (ET(B)) selective antagonist RES-701-1 (1) and endothelin (ET-1). They have N-terminal 10 amino acids derived from 1 and C-terminal 10 amino acids derived from ET-1. RES-701-1(1-10)-[Ala15]ET-1(12-21) and its analogues substituted or truncated at the residues derived from RES-701-1 had proved to possess high receptor binding activity selective for ETB as well as 1. Substitutions at the residues derived from ET-1 had produced some analogues that possessed high affinity not only for ETB but for ETA. Although all analogues had antagonistic effects on ETA, some analogues had proved to function as agonist on ETB confirmed by the changes in intracellular calcium concentrations of ET receptor-transfected COS-7 cells. We have found four types of ET receptor-binding peptides: (1) ETB-selective agonist with weak ETA antagonism (3, KT7421); (2) ETB-selective antagonist with weak ETA antagonism (29, KT7539); (3) ETB agonist with potent ETA antagonism (27, KT7538); and (4) non-selective ETA/ETB antagonist (26, KT7540).  相似文献   

8.
The uterotonic potencies of the naturally occurring mammalian tachykinins and the synthetic subtype-selective agonist analogues of these agents [Lys5,MeLeu9,Nlel0]neurokinin A-(4-10) and [Nle10]neurokinin A-(4-10) (tachykinin NK2 receptor-selective), [Sar9,Met(O2)11]substance P (tachykinin NK1 receptor-selective) and senktide (tachykinin NK3 receptor-selective) were determined using preparations from oestradiol-treated rats. The endopeptidase 24.11 inhibitor, N-[N-[1-(S)-carboxyl-3-phenylpropyl]-(S)-phenyl-alanyl-(S)-isoserine+ ++ (SCH 39370), potentiated responses to neurokinin A, neurokinin B and substance P, but not to [Lys5,MeLeu9,Nle10)]neurokinin A-(4-10) or senktide. [Nle10]neurokinin A-(4-10) effects were potentiated by SCH 39370 with amastatin and those to [Sar9,Met(O2)11]substance P were potentiated by SCH 39370 and captopril in combination. In the presence of optimal concentrations of peptidase inhibitors the relative order of agonist potency was: neurokinin A > substance P > neurokinin B for the naturally occurring mammalian tachykinins and [Lys5,MeLeu9,Nle10]neurokinin A-(4-10) > [Nle10]neurokinin A-(4-10) > [Sar9,Met(O2)11]substance P > senktide for the synthetic tachykinin analogues. Thus, while a tachykinin NK2 receptor predominates in the oestrogen-primed uterus, a tachykinin NK1 receptor may also be present. The non-peptide tachykinin NK3 receptor antagonist, SR 142801, did not antagonise the effects of senktide suggesting that tachykinin NK3 receptors do not mediate its relatively minor effect on the uterus of the oestrogen-primed rat.  相似文献   

9.
The present study describes the labelling of the nociceptin (NC) receptor, ORL1, in mouse forebrain membranes with a new ligand partially protected from metabolic degradation at the C-terminal; the ligand, [3H]-NC-NH2, has a specific activity of 24.5 Ci mmol(-1). Saturation experiments revealed a single class of binding sites with a KD value of 0.55 nM and Bmax of 94 fmol mg(-1) of protein. Non specific binding was 30% of total binding. Kinetic binding studies yielded the following rate constants: Kobs = 0.104 min(-1); K1 =0.034 min(-1): T1/2=20 min; K(+1)=0.07 min nM(-1). Thermodynamic analyses indicated that [3H]-NC-NH2 binding to the mouse ORL1 is totally entropy driven, similar to what has been observed for the labelled agonists to the opioid receptors OP1(delta), OP2(kappa) and OP3(mu). Receptor affinities of several NC fragments and analogues, including the newly discovered ORL-1 receptor antagonist [Phe1psi(CH2-NH)Gly2]NC(1-13)-NH2([F/G]NC(1-13)-NH2), were also evaluated in displacement experiments. The competition curves for these compounds were found to be parallel to that of NC and the following order of potency was determined for NC fragments: NC-OH = NC-NH2-NC(1-13)-NH2 > > NC(1-12)-NH2 > NC(1-13)-OH > > NC(1-11)-NH2, and for NC and NC(1-13)-NH2 analogues: [Tyr1]NC-NH2 > or = [Leu1]NC(1-13)-NH2 > or = [Tyr1]NC(1-13)-NH2 > or = [F/G]NC(1-13)-NH2 > > [Phe3]NC(1-13)-NH2 > [DF/G]NC(1-13)-NH2. Standard opioid receptor ligands (either agonists or antagonists) were unable to displace [3H]-NC-NH2 binding when applied at concentrations up to 10 microM indicating that this new radioligand interacts with a non opioid site, probably the ORL1 receptor.  相似文献   

10.
In the present study we show that the synthetic peptides [4-Cl-D-Phe6,Leu17]VIP and the growth hormone releasing factor (GRF) analog [Ac-Tyr1,D-Phe2]GRF-(1-29)-NH2 inhibit in a competitive manner the specific [125I]VIP binding to both rat and mouse peritoneal macrophages. In rat peritoneal macrophages, the order of potency of the different peptides, as expressed by the IC50 values was: VIP (IC50 = 1.90 +/- 0.16 nM) > [4-Cl-D-Phe6,Leu17]VIP (IC50 = 125.8 +/- 13.2 nM) > [Ac-Tyr1,D-Phe2]GRF-(1-29)-NH2 (IC50 = 354.8 +/- 21.2 nM). In mouse peritoneal macrophages a similar pattern of potency was observed: VIP (IC50 = 1.58 +/- 0.12 nM) > [4-Cl-D-Phe6,Leu17]VIP (IC50 = 110.8 +/- 10.7 nM) > [Ac-Tyr1,D-Phe2]GRF-(1-29)-NH2 (IC50 = 251 +/- 19.2 nM). The behavior as VIP receptor antagonists of both [4-Cl-D-Phe6,Leu17]VIP and [Ac-Tyr1,D-Phe2]GRF-(1-29)-NH2 in rat and mouse peritoneal macrophages was confirmed by: (a) the shift to the right of VIP dose-stimulated cyclic AMP production curves in the presence of the two antagonists; (b) the agreement between the order of efficacy of the two peptides in competition experiments with the corresponding inhibition of cyclic AMP production; (c) the inefficiency of the two antagonists on the stimulation of cyclic AMP production by the beta-adrenoceptor agonist isoproterenol, which indicates the specificity of the interaction; (d) the synergic effect of VIP on isoproterenol-stimulated cyclic AMP production was completely abolished by [4-Cl-D-Phe6,Leu17]VIP or [Ac-Tyr1,D-Phe2]GRF-(1-29)-NH2, suggesting that both antagonists acted via specific VIP receptors. Moreover, propranolol, a beta-adrenoceptor antagonist, did not affect the VIP-stimulated cyclic AMP production and the antagonist role of [4-Cl-D-Phe6,Leu17]VIP or [Ac-Tyr1,D-Phe2]GRF-(1-29)-NH2; (e) in cross-linking experiments, the intensity of the labeling of the [125I]VIP/receptor complexes was significantly lower with the antagonists than in the control experimental situation in both mouse and rat peritoneal macrophage membranes.  相似文献   

11.
1. Potency orders were determined for a series of agonists and antagonists on the calcitonin gene-related peptide (CGRP) receptor of rat L6 myocytes. The agents tested were all shown to have been active against CGRP, amylin or adrenomedullin receptors. 2. AC187 had a pIC50 of 6.8 +/- 0.10, making it 14 fold less potent as an antagonist than CGRP8-37 (pIC50, 7.95 +/- 0.14). Amyline8-37 was equipotent to AC187 (pIC50, 6.6 +/- 0.16) and CGRP19-32 was 3 fold less potent than either (pIC50, 6.1 +/- 0.24). 3. [Ala11]-CGRP8-37 was 6 fold less potent than CGRP8-37, (pIC50, 7.13 +/- 0.14), whereas [Ala18]-CGRP8-37 was approximately equipotent to CGRP8-37 (pIC50, 7.52 +/- 0.15). However, [Ala11,Ala18]-CGRP8-37 was over 300 fold less potent than CGRP8-37 (pIC50, 5.30 +/- 0.04). 4. [Tyr0]-CGRP28-37, amylin19-37 and adrenomedullin22-52 were inactive as antagonists at concentrations of up to 1 microM. 5. Biotinyl-human alpha-CGRP was 150 fold less potent than human alpha-CGRP itself (EC50 values of 48 +/- 17 nM and 0.31 +/- 0.13 nM, respectively). At 1 microM, [Cys(acetomethoxy)2,7]-CGRP was inactive as an agonist. 6. These results confirm a role for Arg11 in maintaining the high affinity binding of CGRP8-37. Arg18 is of less direct significance for high affinity binding, but it may be important in maintaining the amphipathic nature of CGRP and its analogues.  相似文献   

12.
The synthesis and biological activity are reported for 21 6-substituted 2,4-diaminopyrido[3,2-d]pyrimidine analogues (4-24) of piritrexim (PTX) as inhibitors of dihydrofolate reductase (DHFR) and as antitumor agents. Recombinant DHFR from Pneumocystis carinii (pc) and native DHFR from Toxoplasma gondii (tg) were the target enzymes tested; these organisms are responsible for fatal opportunistic infections in AIDS patients. Rat liver (rl) DHFR served as the mammalian reference enzyme to determine selectivity for the pathogenic DHFR. The synthesis of S9-bridged compounds 4-6 was achieved by aryl displacement of 2,4-diamino-6-chloropyrido[3, 2-d]pyrimidine (27) with thiol nucleophiles. Oxidation of 4-6 with hydrogen peroxide in glacial acetic acid afforded the corresponding sulfone analogues 7-9. The N9-bridged compounds 10-24 were synthesized from their precursor 3-amino-6-(arylamino)-2-pyridinecarbonitriles via a thermal cyclization with chloroformamidine hydrochloride. Unlike the S9-bridged compounds, the arylamino side chains of the N9-bridged analogues were introduced prior to the formation of the 2, 4-diaminopyrido[3,2-d]pyrimidine nucleus. A reversed two-atom-bridged analogue (25) was also synthesized using a synthetic strategy similar to that utilized for compounds 10-24. The IC50 values of these compounds against pcDHFR ranged from 0.0023 x 10(-6) M for 2,4-diamino-6-(N-methyl-3',4'-dimethoxyanilino)pyrido[3, 2-d]pyrimidine (21), which was the most potent, to 90.4 x 10(-6) M for 2,4-diamino-6-(4'-methoxyanilino)pyrido[3,2-d]pyrimidine (12), which was the least potent. The three S9-bridged compounds tested were more potent than the corresponding sulfone-bridged compounds for all three DHFRs. N9-Methylation increased the potency by as much as 17 000-fold (compounds 15 and 21). None of the analogues were selective for pcDHFR. Against tgDHFR the most potent analogue was again 21 with an IC50 value of 0.00088 x 10(-6) M and the least potent was 12 with an IC50 of 2.8 x 10(-6) M. N9-Methylation afforded an increase in potency of up to 770-fold (compound 15 NH vs 21 N-CH3) compared to the corresponding N9-H analogue. In contrast to pcDHFR, several analogues had a greater selectivity ratio for tgDHFR compared to trimetrexate (TMQ) or PTX, most notably 2, 4-diamino-6-[(3',4'- dimethoxyphenyl)thio]pyrido[3,2-d]pyrimidine (4), 2,4-diamino-6-[(2'-methoxyphenyl)sulfonyl]pyrido[3, 2-d]pyrimidine (7), and 2,4-diamino-6-(2', 5'-dimethoxyanilino)pyrido[3,2-d]pyrimidine (14) which combined relatively high potency at 10(-7)-10(-8) M along with selectivity ratios of 3.97, 6.67, and 4.93, respectively. Several analogues synthesized had better selectivity ratios than TMQ or PTX for both pcDHFR and tgDHFR, and the potencies of the N9-methylated compounds were comparable to or greater than that of TMQ or PTX. Selected compounds were evaluated as inhibitors of the growth of a variety of tumor cells in culture. The N9-CH3 analogues were, in general, highly potent with GI50 values in the nanomolar range. The N9-H and S9 analogues were less potent with GI50 values in the millimolar to micromolar range.  相似文献   

13.
We determined the effects of two classical angiotensin II (ANG II) antagonists, [Sar1, Ala8]-ANG II and [Sar1, Thr8]-ANG II, and losartan (a nonpeptide and selective antagonist for the AT1 angiotensin receptors) on diuresis, natriuresis, kaliuresis and arterial blood pressure induced by ANG II administration into the median preoptic nucleus (MnPO) of male Holtzman rats weighing 250-300 g. Urine was collected in rats submitted to a water load (5% body weight) 1 h later. The volume of the drug solutions injected was 0.5 microliters over 10-15 s. Pre-treatment with [Sar1, Ala8]-ANG II (12 rats) and [Sar1, Thr8]-ANG II (9 rats), at the dose of 60 ng reduced (13.7 +/- 1.0 vs 11.0 +/0 1.0 and 10.7 +/0 1.2, respectively), whereas losartan (14 rats) at the dose of 160 ng totally blocked (13.7 +/- 1.0 vs 7.6 +/- 1.5) the urine excretion induced by injection o 12 ng of ANG II (14 rats). [Sar1, Ala8]-ANG II impaired Na+ excretion (193 +/- 16 vs 120 +/- 19), whereas [Sar1, Thr8]-ANG II and losartan block Na+ excretion (193 +/- 16 vs 77 +/- 15 and 100 +/- 12, respectively) induced by ANG II. Similar effects induced by ANG II on K+ excretion were observed with [Sar1, Ala8]-ANG II, [Sar1, Thr8]- ANG II, and losartan pretreatment (133 +/- 18 vs 108 +/- 11, 80 +/- 12, and 82 +/- 15, respectively). The same doses as above of [Sar1, Ala8]-ANG II (8 rats), [Sar1, Thr8]-ANG II (8 rats), and losartan (9 rats) blocked the increase in the arterial blood pressure induced by 12 ng of ANG II (12 rats) (32 +/- 4 vs 4 +/- 2, 3.5 +/- 1, and 2 +/- 1, respectively. The results indicate that the AT1 receptor subtype participates in the increases of diuresis, natriuresis, kaliuresis and arterial blood pressure induced by the administration of ANG II into the MnPO.  相似文献   

14.
In the linear gramicidins, the four aromatic residues at positions 9, 11, 13, and 15 are well-known to be important for the structure and function of membrane-spanning gramicidin channels. To investigate whether the "spacer" residues between the tryptophans in gramicidin A (gA) are important for channel structure and function, D-Leu-10, -12. and -14 of gA were replaced by Ala, Val, or Ile. (For practical reasons, the Ile substitutions were introduced into the enantiomeric gramicidin A-, gA-.) Circular dichroism spectra of [D-Ala10,12,14]gA, [D-Val10,12,14]gA, or [Ile10,12,14]gA- incorporated into sodium dodecyl sulfate micelles or 1, 2-dimyristoyl-sn-glycero-3-phosphocholine vesicles differ from the spectrum of the native [D-Leu10,12,14]gA. All the analogue spectra display reduced ellipticity at both 218 and 235 nm, indicating the presence of double-stranded conformers with the Ala analogue spectra showing the largest departure from the native gA spectra. Size-exclusion chromatograms of the Val and Ile analogues show both monomer and dimer peaks, accompanied by peak broadening; the chromatograms for the Ala analogue show broad, overlapping peaks and suggest the presence of higher oligomers and/or (rapidly) interconverting conformations. All three analogues form membrane-spanning channels, with the channel-forming potency of the Ala analogue being much less than that of gA or the other analogues. In 1.0 M CsCl, the conductance of each analogue channel is approximately 25% less than that of [D-Leu10,12,14]gA channels. The lifetimes of the analogue channels also are less than of [D-Leu10,12, 14]gA channels, with the largest (8-fold) reduction being for [D-Ala10,12,14]gA channels. Hybrid channel experiments show that the beta6.3-helical backbone folding pattern is retained in the channel-forming subunits and that the substitutions primarily influence ion entry. Both the bulk and the stereochemistry of the aliphatic residues between the tryptophans of gA are important for channel structure and function.  相似文献   

15.
The neurotoxic fragment corresponding to residues 25-35 of the beta-amyloid (A beta) peptide [A beta-(25-35)] has been shown to exert effects on (+)-[3H]5-methyl-10, 11-dihydro-5H-dibenzo[a,d]-cyclohepten-5,10-imine maleate ([3H]MK-801) binding to the cation channel of the N-methyl-D-aspartate (NMDA) receptor. In the present study, we investigated whether the amidated and carboxylic acid C-terminated forms of A beta-(25-35) [A beta-(25-35-NH2) and A beta-(25-35-COOH), respectively] exert effects on other excitatory amino acid receptor and cation channel types in rat cortical membranes. Both A beta-(25-35-NH2) and A beta-(25-35-COOH) gave statistically significant dose-dependent inhibitions of [3H]glutamate and [3H]glycine binding to the agonist recognition sites of the NMDA receptor. Ten microM A beta-(25-35-NH2) and A beta-(25-35-COOH) gave 25% and 20% inhibitions of [3H]glutamate binding and 75% and 70% inhibitions of [3H]glycine binding, respectively. A beta-(25-35-NH2), but not A beta-(25-35-COOH), gave a small (ca. 17% at 10 microM) statistically significant increase of [3H]amino-3-hydroxy-5-methylisoxazole-4-propionate ([3H]AMPA) binding. [3H]kainate binding was not significantly affected by either peptide. Similarly, neither peptide affected either the maximal level or EC50 value for calcium stimulation of [3H]nitrendipine binding. It is concluded that A beta-(25-35) shows slight affinity for the agonist recognition sites of the NMDA receptor, but not for other excitatory amino acid receptor types or for L-type voltage-dependent calcium channels.  相似文献   

16.
Receptors for regulatory peptides such as somatostatin or vasoactive intestinal polypeptide are expressed by a number of human neoplasms and can be visualized in vivo with peptide receptor scintigraphy. Recently, the CCK-B receptor, which binds both gastrin and cholecystokinin with high affinity, was shown using in vitro methods to be overexpressed in a number of human tumor tissues, including medullary thyroid carcinomas, small cell lung cancers, astrocytomas, gastrointestinal tumors, and stromal ovarian cancers. In the present study, we have designed novel, unsulfated CCK octapeptide analogs linked to the metal chelating DTPA and DOTA, and have tested them for their binding affinity to CCK-B receptor-positive tissue from human tumors: The most potent compounds assayed were DTPA-[Nle28, 31]-CCK(26-33) (MP2286) and DTPA-[d-Asp26,Nle28,31]-CCK(26-33) (MP2288) with an IC50 of 1.5 nM. For comparison, analogs with C-terminal DTPA, such as [Nle28,31,Aphe33(p-NH-DTPA)]-CCK(26-33) and CCK-(26-33)-NH(CH2)2 NH-DTPA, had an IC50 of >100 nM. DOTA-[D-Asp26, Nle28,31]-CCK(26-33) had an IC50 of 3.9 nM. The compounds were selective for CCK-B receptors as they did not bind with high affinity to CCK-A receptors expressed in human tumors (meningiomas or gastroenteropancreatic tumors). In vivo rat biodistribution studies with indium-111 labeled MP2286 and MP2288 showed that the primary mode of clearance was renal, and the primary sites of uptake (% ID/g 24 h p.i.) were kidneys (0.270 and 0.262, respectively) and the gastrointestinal tract. The CCK-B receptor-expressing gastric mucosa showed specific in vivo accumulation of 111In-labeled MP2288 which could be blocked in the presence of excess unlabeled MP2288. 111In-labeled MP2286 and MP2288 were also found to be stable in human plasma whereas both compounds were degraded in urine (>40% after 3 h at 37 degrees C). The affinity, specificity, biodistribution, and stability of these two DTPA-CCK analogs indicate that these compounds have substantial promise for use in the in vivo visualization of CCK-B receptor-expressing tumors.  相似文献   

17.
Examination of conformationally constrained melanotropin peptide (Ac-Nle4-c[Asp5-His-Phe7-Arg-Trp9-Ala-Lys]-NH2) on four human melanotropin receptors (hMC1R, hMC3R, hMC4R, and hMC5R) resulted in identifying the importance of ligand stereochemistry at positions 5, 7, and 9 for agonist binding affinity and receptor selectivity. A trend in ligand structure-activity relationships emerged for these peptides, with the hMC1R and hMC4R possessing similar tendencies, as did the hMC3R and hMC5R. alpha-MSH (Ac-Ser-Tyr-Ser-Met4-Glu-His-Phe7-Arg-Trp-Gly-Lys-Pro-Val-NH2), NDP-MSH (Ac-Ser-Tyr-Ser-Nle4-Glu-His-D-Phe7-Arg-Trp-Gly-Lys-Pro-Val-NH2), and MTII (Ac-Nle4-c[Asp5,D-Phe7,Lys10]-alpha-MSH(4-10)-NH2) were also examined at each of these melanocortin receptors. Interestingly, the linear NDP-MSH possessed greater binding affinity for the hMC3R and hMC5R than did the cyclic analogue MTII. The peptide Ac-Nle-c[Asp-His-Phe-Arg-D-Trp9-Ala-Lys]-NH2 demonstrated the greatest differentiation in binding affinity between the hMC1R and hMC4R (78-fold). Analogue Ac-Nle-c[Asp-His-Phe7-Arg-Trp-Ala-Lys]-NH2 resulted in micromolar binding affinity (or greater) at the hMC3R and hMC5R, demonstrating the importance of D-Phe7 for ligand binding potency at these receptors. Ac-c[Asp-His-Phe-Arg-Trp-Ala-Lys]-NH2 resulted in loss of binding affinity at the hMC5R, implicating the importance of Nle4 (or a hydrophobic residue in this position) for binding to this receptor. Ac-Nle-c[D-Asp5-His-Phe-Arg-Trp-Ala-Lys]-NH2 was unable to competitively displace [125I]NDP-MSH binding at micromolar concentrations on the hMC3R and hMC5R, suggesting the importance of chirality of Asp5 either for ligand-receptor interactions or for orientation of the side chain lactam bridge and the structural integrity of the peptide conformation. Energy calculations performed for these peptides resulted in the identification of a low-energy ligand conformer family that is common to all the ligands. The differences in ligand binding affinities observed in this study are postulated to be a result of different ligand-receptor complexed interactions and not solely to the ligand structure.  相似文献   

18.
Cyclic ADP-ribose (cADPR) is a potentially important intracellular Ca2+ releasing messenger [1-5]. In pancreatic acinar cells where intracellular infusion of both inositol trisphosphate (IP3) and cADPR evoke repetitive Ca2+ spiking [6], the cADPR antagonist 8-NH2-cADPR [7], which blocks cADPR-evoked but not IP3-evoked Ca2+ spiking, can abolish Ca2+ spiking induced by physiological levels of the peptide hormone cholecystokinin (CCK) [8]. We have tested the effect of intracellular glucose on the ability of IP3, cADPR and CCK to induce cytosolic Ca2+ spikes in pancreatic acinar cells. In order to gain access to the intracellular cytosol, we used the whole-cell configuration of the patch-clamp technique [9] and monitored cytosolic Ca2+ concentration changes by measuring the Ca(2+)-dependent ionic current [10-13]. Glucose (300 microM to 10 mM) in the patch pipette/intracellular solution prevented cADPR from evoking Ca2+ spiking. The same effect was observed with 2-deoxy-glucose, but not L-glucose. In contrast, glucose potentiated IP3-evoked Ca2+ spiking. CCK evoked Ca2+ spiking irrespective of the presence or absence of intracellular glucose, but the cADPR antagonist 8-NH2-cADPR blocked CCK-evoked Ca2+ spiking only in the absence of intracellular glucose. This suggests that the hormone can evoke Ca2+ spiking via either the IP3 or the cADPR pathway. The intracellular glucose level may control a switch between these two pathways.  相似文献   

19.
The synthesis and biological activity of 42 6-substituted-2,4-diaminopyrido[3,2-d]pyrimidines (2,4-diamino-8-deazafolate analogues) are reported. The compounds were synthesized in improved yields compared to previous classical analogues using modifications of procedures reported previously by us. Specifically, the S-phenyl-; mono-, di-, and trimethoxyphenyl-; and mono-, di-, and trichlorophenyl-substituted analogues with H or CH3 at the N10 position and methyl and trifluoromethyl phenyl ketone analogues with H, CH3, and CH2C identical to CH at the N10 position were synthesized. The S10 and N10 alpha- and beta-naphthyl analogues along with the N10 CH3 analogues were also synthesized. These compounds were evaluated as inhibitors of dihydrofolate reductases (DHFR) from Pneumocystis carinii (pc) and Toxoplasma gondii (tg); selectivity ratios were determined against rat liver (rl) DHFR as the mammalian reference enzyme. Against pcDHFR the IC50 values ranged from 0.038 x 10-6 M for 2,4-diamino-6-[(N-methyl-2'-naphthylamino)methyl]pyrido[3,2-d]pyrimidine (28) to 5.5 x 10(-6) M for 2,4-diamino-6[(2',4'-dimethoxyanilino)methyl]pyrido[3,2-d]pyrim idi ne (15). N10 methylation in all instances increased potency. None of the analogues were selective for pcDHFR. Against tgDHFR the most potent analogue was 2,4-diamino-6-[(N-methylanilino)methyl]pyrido[3,2-d]pyrimidine (5) (IC50 0.0084 x 10(-6) M) and the least potent was 2,4-diamino-6[(2'-naphthylamino)methyl]-pyrido[3,2-d]pyrimidine (37) (IC50 0.16 x 10-6 M). N10 methylation afforded an increase in potency up to 10-fold. In contrast to pcDHFR, several of the 8-deaza analogues were significantly selective for tgDHFR, most notably 2,4-diamino-6-[(2'-chloro-N-methylanilino)-methyl]pyrido[3,2-d] pyrimidine (13), 2,4-diamino-6-[(3',4',5'-trimethoxyanilino)methyl]pyrido[3,2-d]pyr pyrimidine (29), and 2,4-diamino-6-[(2',4',6'-trichloroanilino)methyl]pyrido[3,2-d] pyrimidine (32) which combined high potency at 10-8 M along with selectivities of 8.0, 5.0, and 12.4, respectively. The potency of these three analogues are comparable to the clinically used agent trimetrexate while their selectivities for tgDHFR are 17-43-fold better than trimetrexate.  相似文献   

20.
A chimeric protein comprised of melanocortin-4 receptor (MC4R) and the green fluorescent protein (GFP) was created for studying receptor/ligand localization and trafficking. The ligand binding affinities and second messenger stimulation induced by MC4R-GFP closely resembled those of the wild-type receptor, suggesting functional integrity of the chimeric protein. As observed with a confocal microscope, in human embryonic kidney (HEK)-293 cells MC4R/GFP was distributed evenly along the cell membrane. Addition of [Nle4-d-Phe7]-alpha-melanocyte-stimulating hormone (NDP-MSH), a peptide MC4R agonist, induced receptor translocation into intracellular compartments in a time- and concentration-dependent manner. [Ac-Nle-c[Asp-His-d-Nal(2')-Arg-Trp-Lys]-NH2] (SHU9119), a potent MC4R antagonist, completely inhibited NDP-MSH-mediated internalization. MC4R-GFP internalization was unaffected by a protein kinase A inhibitor N-[2-(p-bromocinnamylamino)ethyl]-5-isoquinolinesulfonamide (H89), but was impaired by pretreatment with inhibitors of endocytosis through clathrin-coated pits, hypertonic sucrose, or concanavalin A. Time-dependent colocalization of MC4R-GFP with rhodamine-transferrin, an early endosome marker, and with LysoTraker, a lysosome marker, was observed after short-term (45 min) and prolonged (20 h) agonist exposure, respectively. Rhodamine-[AcNle-c[Asp-His-d-Phe-Arg-Trp-Lys]-NH2] (MTII), a fluorescent derivative of an MC4R agonist, was found to cointernalize with MC4R-GFP into intracellular vesicles. No significant receptor recycling or segregation from the ligand was observed 60 min after removal of the agonist. In contrast, an antagonist rhodamine-Ac-Cys-Glu-His-(d-Nal)-Arg-Trp-Gly-Cys-Pro-Pro-Lys-Asp-NH2 (HS014) bound to and colocalized with MC4R-GFP on the cell surface and did not stimulate receptor internalization. In sum, these results suggest that MC4R is subject to agonist-dependent endocytosis via clathrin-coated pits. Prolonged agonist exposure directs MC4R into lysosomes, possibly for degradation. Receptor and ligand recycling is not efficient for MC4R in HEK-293 cells.  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-23

京公网安备 11010802026262号