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 468 毫秒
1.
The temperature dependence of the partition of a neuropeptide, Substance P (SP), and its [Tyr8] analogue in a widely used membrane mimic, dodecylphosphocholine micelles, was studied by using a pulsed field gradient nmr diffusion technique. The partition coefficient was found to decrease when the temperature is increased, indicating a favorable (negative) enthalpy change upon partitioning of the peptides. Thermodynamic functions of the partitioning were determined. The enthalpy of partition ΔHpart, was found to be in the −2.5 to −3.0 kcal/mol range, which is between 2 and 3 times higher than the entropic term −TΔSpart. The free energy of partitioning is consistent with a model in which the SP peptides interact with the micelles mainly through the hydrophobic side chains of the residues Phe7, Phe8 (or Tyr8), Leu10, and Met11, and without the insertion of a major portion of the peptide into the hydrophobic core of the micelles. © 1998 John Wiley & Sons, Inc. Biopoly 45: 395–403, 1998  相似文献   

2.
The vibrational structures of Nociceptin (FQ), its short bioactive fragments, and specifically‐modified [Tyr1]FQ (1‐6), [His1]FQ (1‐6), and [His1,4]FQ (1‐6) fragments were characterized. We showed that in the solid state, all of the aforementioned peptides except FQ adopt mainly turn and disordered secondary structures with a small contribution from an antiparallel β‐sheet conformation. FQ (1‐11), FQ (7‐17) [His1]FQ (1‐6), and [His1,4]FQ (1‐6) have an α‐helical backbone arrangement that could also slightly influence their secondary structure. The adsorption behavior of these peptides on a colloidal silver surface in an aqueous solution (pH = ~8.3) was investigated by means of surface‐enhanced Raman scattering (SERS). All of the peptides, excluding FQ (7‐17), chemisorbed on the colloidal silver surfaces through a Phe4 residue, which for FQ, FQ (1‐11), FQ (1‐6), [Tyr1]FQ (1‐6), and [His1]FQ (1‐6) lies almost flat on this surface, while for FQ (1‐13) and FQ (1‐13)NH2 adopts a slightly tilted orientation with respect to the surface. The Tyr1 residue in [Tyr1]FQ (1‐6) does not interact with the colloidal silver surface, suggesting that the Tyr1 and Phe4 side chains are located on the opposite sides of the peptide backbone, which can be also true for His1 and Phe4 in [His1]FQ (1‐6). The lone pair of electrons on the oxygen atom of the ionized carbonyl group of FQ (1‐13) and FQ (7‐17) appears to be coordinated to the colloidal silver nanoparticles, whereas in the case of the remaining peptides, it only assists in the adsorption process, similar to the ? NH2 group. We also showed that upon adsorption, the secondary structure of these peptides is altered. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1039–1054, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

3.
Abstract

Melanin‐concentrating hormone (MCH) is a neuropeptide occurring in all vertebrates and some invertebrates and is now known to stimulate pigment aggregation in teleost melanophores and food‐intake in mammals. Whereas the two MCH receptor subtypes hitherto cloned, MCH‐R1 and MCH‐R2, are thought to mediate mainly the central effects of MCH, the MCH‐R on pigment cells has not yet been identified, although in some studies MCH‐R1 was reported to be expressed by human melanocytes and melanoma cells. Here we present data of a structure‐activity study in which 12 MCH peptides were tested on rat MCH‐R1 and mouse B16 melanoma cell MCH‐R, by comparing receptor binding affinities and biological activities. For receptor binding analysis with HEK‐293 cells expressing rat MCH‐R1 (SLC‐1), the radioligand was [125I]–[Tyr13]‐MCH with the natural sequence. For B16 cells (F1 and G4F sublines) expressing B16 MCH‐R, the analog [125I]–[D‐Phe13, Tyr19]‐MCH served as radioligand. The bioassay used for MCH‐R1 was intracellular Ca2+ mobilization quantified with the FLIPR instrument, whereas for B16 MCH‐R the signal determined was MAP kinase activation. Our data show that some of the peptides displayed a similar relative increase or decrase of potency in both cell types tested. For example, linear MCH with Ser residues at positions 7 and 16 was almost inactive whereas a slight increase in side‐chain hydrophilicity at residues 4 and 8, or truncation of MCH at the N‐terminus by two residues hardly changed binding affinity or bioactivity. On the other hand, salmonic MCH which also lacks the first two residues of the mammalian sequence but in addition has different residues at positions 4, 5, 9, and 18 exhibited a 5‐ to 10‐fold lower binding activity than MCH in both cell systems. A striking difference in ligand recognition between MCH‐R1 and B16 MCH‐R was however observed with modifications at position 13 of MCH: whereas L‐Phe13 in [Phe13, Tyr19]‐MCH was well tolerated by both MCH‐R1 and B16 MCH‐R, change of configuration to D‐Phe13 in [D‐Phe13, Tyr19]‐MCH or [D‐Phe13]‐MCH led to a complete loss of biological activity and to a 5‐ to 10‐fold lower binding activity with MCH‐R1. By contrast, the D‐Phe13 residue increased the affinity of [D‐Phe13, Tyr19]‐MCH to B16 MCH‐R about 10‐fold and elicited MAP kinase activation as observed with [Phe13, Tyr19]‐MCH or MCH. These data demonstrate that ligand recognition by B16 MCH‐R differs from that of MCH‐R1 in several respects, indicating that the B16 MCH‐R represents an MCH‐R subtype different from MCH‐R1.  相似文献   

4.
The tachykinins, substance P (SP) and neurokinin A (NKA), are agonists for the NK1 and NK2 receptors, respectively. Tachykinins have various respiratory effects, including bronchoconstriction. This study characterizes tachykinin binding sites in the rabbit lung. We hypothesize that (2-[125I]iodohistidyl1)Neurokinin A ([125I]NKA) interacts with NK1 and NK2 binding sites in the rabbit lung. The Kd determined from saturation isotherms was 0.69 X/÷1.14 nM (geometic mean X/÷ SEM) and the Bmax was 4.15±0.22 femtomole/mg protein (arithmetic mean±SEM). Competitive inhibition studies with NKA, SP and various selective tachykinin agonists showed the rank order of potency: [β-Ala8]-Neurokinin A 4–10=SP ≫ NKA ≫ [Sar9,Met(O2)11]-Substance P. [β-Ala8]-Neurokinin A 4–10, a selective NK2 agonist, and SP inhibition of [125I]NKA binding were best described using a two-site model. Competitive inhibition studies using the selective nonpeptide NK2 antagonist (SR 48968) and the selective nonpeptide NK1 antagonist (CP 96,345) revealed Ki's of 5.5 nM and 8.1 nM, respectively. Our data therefore suggest that [125I]NKA binds to both the NK1 and NK2 receptors in the lung. Special issue dedicated to Dr. Kinya Kuriyama.  相似文献   

5.
A photoreactive analogue of human melanin‐concentrating hormone was designed, [d‐ Bpa13,Tyr19]‐MCH, containing the d‐ enantiomer of photolabile p‐benzoylphenylalanine (Bpa) in position 13 and tyrosine for radioiodination in position 19. The linear peptide was synthesized by the continuous‐flow solid‐ phase methodology using Fmoc‐strategy and PEG‐PS resins, purified to homogeneity and cyclized by iodine oxidation. Radioiodination of [d ‐Bpa13,Tyr19]‐MCH at its Tyr19 residue was carried out enzymatically using solid‐ phase bound glucose oxidase/lactoperoxidase, followed by purification on a reversed‐ phase mini‐column and HPLC. Saturation binding analysis of [125I]‐[d‐ Bpa13,Tyr19]‐MCH with G4F‐7 mouse melanoma cells gave a KD of 2.2±0.2×10−10 mol/l and a Bmax of 1047±50 receptors/cell. Competition binding analysis showed that MCH and rANF(1–28) displace [125I]‐[d‐ Bpa13,Tyr19]‐MCH from the MCH binding sites on G4F‐7 cells whereas α‐MSH has no effect. Receptor crosslinking by UV‐irradiation of G4F‐7 cells in the presence of [125I]‐[d‐ Bpa13,Tyr19]‐MCH followed by SDS‐polyacrylamide gel electrophoresis and autoradiography yielded a band of 45–50 kDa. Identical crosslinked bands were also detected in B16‐F1 and G4F mouse melanoma cells, in RE and D10 human melanoma cells as well as in COS‐7 cells. Weak staining was found in rat PC12 phaeochromocytoma and Chinese hamster ovary cells. No crosslinking was detected in human MP fibroblasts. These data demonstrate that [125I]‐[d‐ Bpa13,Tyr19]‐MCH is a versatile photocrosslinking analogue of MCH suitable to identify MCH receptors in different cells and tissues; the MCH receptor in these cells appears to have the size of a G protein‐coupled receptor, most likely with a varying degree of glycosylation. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
Gomesin (Gm) has a broad antimicrobial activity making it of great interest for development of drugs. In this study, we analyzed three Gm analogs, [Trp1]‐Gm, [Trp7]‐Gm, and [Trp9]‐Gm, in an attempt to gain insight into the contributions of different regions of the peptide sequence to its activity. The incorporation of the tryptophan residue in different positions has no effect on the antimicrobial and hemolytic activities of the Gm analogs in relation to Gm. Spectroscopic studies (circular dichroism, fluorescence and absorbance) of Gm and its analogs were performed in the presence of SDS, below and above its critical micelle concentration (CMC) (~8 mM), in order to monitor structural changes induced by the interaction with this anionic surfactant (0–15 mM). Interestingly, we found that the analogs interact more strongly with SDS at low concentrations (0.3‐6.0 mM) than close to or above its CMC. This suggests that SDS monomers are able to cover the whole peptide, forming large detergent‐peptide aggregates. On the other hand, the peptides interact differently with SDS micelles, inserting partially into the micelle core. Among the peptides, Trp in position 1 becomes more motionally‐restricted in the presence of SDS, probably because this residue is located at the N‐terminal region, which presents higher conformational freedom to interact stronger with SDS molecules. Trp residues in positions 7 and 9, close to and in the region of the turn of the molecule, respectively, induced a more constrained structure and the compounds cannot insert deeper into the micelle core or be completely buried by SDS monomers. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
[Tyr6]‐γ2‐MSH(6–12) with a short effecting time of about 20 min is one of the most potent rMrgC receptor agonists. To possibly increase its potency and metabolic stability, a series of analogues were prepared by replacing the Tyr6 residue with the non‐canonical amino acids 3‐(1‐naphtyl)‐L ‐alanine, 4‐fluoro‐L ‐phenylalanine, 4‐methoxy‐L ‐phenylalanine and 3‐nitro‐L ‐tyrosine. Dose‐dependent nociceptive assays performed in conscious rats by intrathecal injection of the MSH peptides showed [Tyr6]‐γ2‐MSH(6–12) hyperalgesic effects at low doses (5–20 nmol) and analgesia at high doses (100–200 nmol). This analgesic activity is fully reversed by the kyotorphin receptor‐specific antagonist Leu–Arg. For the two analogues containing in position 6, 4‐fluoro‐L ‐phenylalanine and 3‐nitro‐L ‐tyrosine, a hyperalgesic activity was not observed, while the 3‐(1‐naphtyl)‐L ‐alanine analogue at 10 nmol dose was found to induce hyperalgesia at a potency very similar to γ2‐MSH(6–12), but with longer duration of the effect. Finally, the 4‐methoxy‐L ‐phenylalanine analogue (0.5 nmol) showed greatly improved hyperalgesic activity and prolonged effects compared to the parent [Tyr6]‐γ2‐MSH(6–12) compound. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
Summary [Aib3,Thr5]OT, [Aib3,Thr(OMe)5]OT, [Aib3,Orn8]OT, [Thr(OMe)5,Orn8]OT and [Phe2,Thr(OMe)5,Orn8]OT were synthesized by solid-phase techniques. From the biological properties of these peptides, it seems that the simultaneous replacement of positions 3 and 5 of oxytocin with Aib and Thr(OMe) results in an analogue devoid of antagonistic activity in comparison with the singly substituted compounds. Simultaneous Orn8 substitution does the same in the case of the Aib3 analogue and even leads to agonistic activity in the case of the Thr(OMe)5 analogue. Replacement of Tyr2 by Phe2, e.g. [Phe2,Thr(OMe)5,Orn8]OT, again favors the appearance of minor antagonistic potency.  相似文献   

9.
    
The synthesis of [Phe(p-CH2SO3Na)52, Nle32,53,56 Nal55]-CCK20–58, [Tyr52, Nle32,53,56, Nal55]-CCK-58 and of [Phe(p-CH2SO3Na)52, Nle32,53,56, Nal55]-CCK-58 using the (9-fluorenylmethyloxy)-carbonyl (Fmoc) strategy on a 2,4-DMBHA resin is described. The crude peptide preparations were extremely complex when analyzed by RP-HPLC, capillary zone electrophoresis (CZE), and ion-exchange chromatography (IE-FPLC). We found that the most effective strategy for purification included cation-exchange chromatography followed by a RP-HPLC desalting step. The highly purified peptides (purity greater than 90%) were characterized by RP-HPLC, size exclusion HPLC (SEC), IE-FPLC, CZE, mass spectrometry, amino acid analysis, and Edman sequence analysis {for [Tyr52, Nle32,53,56, Nal55]-CCK-58}. The results demonstrate the applicability of the 2,4-DMBHA resin for Fmoc solid-phase synthesis of long peptides amides (58 residues in length in this case) as well as the efficacy of an FPLC/RP-HPLC approach for the purification of very long, heterogeneous crude peptides, allowing a true assessment of the biological properties of these analogs to be carried out. [Phe(p-CH2SO3Na)52, Nle32,53,56, Nal55]-CCK20–58 was less than 1% as potent as CCK-8 while [Tyr52, Nle32,53,56, Nal55]-CCK-58 and [Phe(p-CH2SO3Na)52, Nle32,53,56, Nal55]-CCK-58 were inactive at the doses tested (<0.01%).  相似文献   

10.
The synthesis of [Phe(p-CH2SO3Na)52, Nle32,53,56 Nal55]-CCK20–58, [Tyr52, Nle32,53,56, Nal55]-CCK-58 and of [Phe(p-CH2SO3Na)52, Nle32,53,56, Nal55]-CCK-58 using the (9-fluorenylmethyloxy)-carbonyl (Fmoc) strategy on a 2,4-DMBHA resin is described. The crude peptide preparations were extremely complex when analyzed by RP-HPLC, capillary zone electrophoresis (CZE), and ion-exchange chromatography (IE-FPLC). We found that the most effective strategy for purification included cation-exchange chromatography followed by a RP-HPLC desalting step. The highly purified peptides (purity greater than 90%) were characterized by RP-HPLC, size exclusion HPLC (SEC), IE-FPLC, CZE, mass spectrometry, amino acid analysis, and Edman sequence analysis {for [Tyr52, Nle32,53,56, Nal55]-CCK-58}. The results demonstrate the applicability of the 2,4-DMBHA resin for Fmoc solid-phase synthesis of long peptides amides (58 residues in length in this case) as well as the efficacy of an FPLC/RP-HPLC approach for the purification of very long, heterogeneous crude peptides, allowing a true assessment of the biological properties of these analogs to be carried out. [Phe(p-CH2SO3Na)52, Nle32,53,56, Nal55]-CCK20–58 was less than 1% as potent as CCK-8 while [Tyr52, Nle32,53,56, Nal55]-CCK-58 and [Phe(p-CH2SO3Na)52, Nle32,53,56, Nal55]-CCK-58 were inactive at the doses tested (<0.01%).  相似文献   

11.
Liao YY  Lee CW  Ho IK  Chiou LC 《Life sciences》2012,90(7-8):306-312
AimThe nociceptin/orphanin FQ (N/OFQ) peptide (NOP) receptor was reported to be functionally heterogeneous. We investigated if [Tyr10]N/OFQ(1-11), a peptide ligand reported to selectively bind to the high affinity site of 125I-[Tyr14]N/OFQ in rodent brains, can be a tool for revealing the NOP receptor heterogeneity. We have previously founded an NOP receptor subset insensitive to Ro 64-6198 and (+)-5a Compound, two non-peptide NOP agonists, in rat ventrolateral periaqueductal gray (vlPAG) neurons. Here, we examined if [Tyr10]N/OFQ(1-11) differentiated (+)-5a Compound-sensitive and -insensitive vlPAG neurons. Certain mu-opioid (MOP) receptor ligands highly competing with [Tyr10]N/OFQ(1-11) in binding studies also showed high affinity at expressed heteromeric NOP–MOP receptors. We also examined if [Tyr10]N/OFQ(1-11) distinguished heteromeric NOP–MOP receptors from homomeric NOP receptors.Main methodsThe NOP receptor activity was evaluated by G-protein coupled inwardly rectifying potassium (GIRK) currents in rat vlPAG slices, and by inhibition of cAMP accumulation in HEK293 cells expressing NOP receptors or co-expressing NOP and MOP receptors.Key findingsIn vlPAG neurons, [Tyr10]N/OFQ(1-11), like N/OFQ, induced GIRK currents through NOP receptors. It was less potent (EC50: 8.98 μM) but equi-efficacious as N/OFQ. [Tyr10]N/OFQ(1-11) displayed different pharmacological profiles as (+)-5a Compound, and was effective in both (+)-5a Compound-sensitive and -insensitive neurons. In NOP-expressing HEK293 cells and NOP- and MOP-co-expressing cells, [Tyr10]N/OFQ(1-11) displayed similar concentration–response curves in decreasing cAMP accumulation.Significance[Tyr10]N/OFQ(1-11) is an NOP full agonist and less potent than N/OFQ. However, it can neither reveal the functional heterogeneity of NOP receptors in vlPAG neurons nor differentiate heteromeric NOP–MOP and homomeric NOP receptors.  相似文献   

12.
The rate of degradation of 125I-labelled [Tyr11]somatostatin by isolated rat hepatocytes was similar to that of unlabelled somatostatin. Reaction was dependent upon cell concentration and temperature, being rapid at 37°C and negligible at 0°C. The apparent Km for the overall degradation process was approximately the same for degradation by hepatocytes and by partially-purified liver plasma membranes. Extracellular breakdown of somatostatin, by proteases released from cells into the incubation medium, represented less than 10% of the cell-associated degradation. Homogenization of hepatocytes resulted in a 10–20-fold increase in the degrading ability of the cells. After incubation of 125I-labelled [Tyr11]somatostatin and 125I-labelled [Tyr1]somatostatin with hepatocytes, 125I-labelled tyrosine was the major radioactive product identified in the incubation medium. The rate of release of 125I-labelled tyrosine from the labelled [Tyr1] analogue was approximately 11 times greater than from the labelled [Tyr11] analogue. 125I-labelled [Tyr11]somatostatin bound to the cells in a non-saturable manner and approx. 70% of the cell-associated radioactivity could be dissociated by dilute acid. The rate of degradation of somatostatin was unchanged by reagents that inhibit the internalisation and lysosomal degradation of polypeptides by cell suspensions but was reduced by reagents that inhibit sulphydryl-dependent proteases. It is proposed that plasma-membrane associated proteolysis, involving both endo- and exopeptidases may represent the predominant degradative pathway of somatostatin in vivo.  相似文献   

13.
High affinity binding sites for a somatostatin-28 analog in rat brain   总被引:8,自引:0,他引:8  
J C Reubi  M H Perrin  J E Rivier  W Vale 《Life sciences》1981,28(19):2191-2198
Using an iodinated analog of a large (28 residues) and biologically active form of somatostatin, 125I[Leu8,D-Trp22,Tyr25]SS-28, it was possible to demonstrate saturable and high affinity binding sites (dissociation constant = 0.46 ± 0.04 nM) in rat cortical membranes. Somatostatin, somatostatin-28, as well as two potent analogs, [D-Trp8] somatostatin and [D-Trp22] somatostatin-28, could completely displace the radiogland in the nanomolar range whereas the inactive analog Des-Trp8-somatostatin and the unrelated peptide GnRH showed no affinity for these binding sites; octa- and nona-peptide analogs of somatostatin were inactive. High binding was found in hippocampus, amygdala, tuberculum olfactorium, caudate-putamen and cortex; moderate binding in midbrain and hypothalamus, and no binding in the cerebellum. These results suggest that specific somatostatin receptors can be measured within the brain with 125I[Leu8,D-Trp22,Tyr25] SS-28 as radioligand.  相似文献   

14.
Vasopressin and nonmammalian hormone vasotocin are known to increase the water permeability of mammalian collecting ducts, frog skin and the urinary bladder. Neurohypophysial nonapeptides have also been shown to interfere with the regulation of renal ion transport. The subject of this study was a search for vasopressin and vasotocin analogues with selective effects on renal water, sodium and potassium excretion. During this study, we synthesised the following peptides: 13 vasotocin analogues modified at positions 4 (Thr or Arg), 7 (Gly or Leu) and 8 (d ‐Arg, Lys or Glu); 4 vasopressin analogues modified at positions 4 and 8; and 9 peptides shortened or extended at the C‐terminal or with substitutions for Gly‐NH2. Most of these peptides had mercaptopropionic acid (Mpa) instead of Cys in position 1. The effects of these nonapeptides on renal water, sodium and potassium transport were evaluated in in vivo experiments using Wistar rats. Some nonapeptides possessed antidiuretic, natriuretic and kaliuretic activities ([Mpa1]‐arginine vasotocin, [Mpa1, homoArg8]‐vasotocin, [Mpa1, Thr4]‐arginine vasotocin and [Mpa1, Arg4]‐arginine vasopressin). Substitutions at positions 4 and 8 increased the selectivity of peptide actions. The antidiuretic [d ‐Arg8]‐vasotocin analogues had no effects on sodium excretion. [Mpa1, Arg4]‐arginine vasotocin was antidiuretic and kaliuretic but not natriuretic. [Mpa1, Glu8]‐oxytocin had weak natriuretic activity without any effects on water and potassium transport. In accordance with the data obtained, synthesised vasotocin analogues could be good candidates for pharmaceuticals selectively regulating renal sodium and potassium transport, which is of clinical importance. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
Angiotensin II (AII) as well as analog peptides shows antimalarial activity against Plasmodium gallinaceum and Plasmodium falciparum, but the exact mechanism of action is still unknown. This work presents the solid‐phase synthesis and characterization of eight peptides corresponding to the alanine scanning series of AII plus the amide‐capped derivative and the evaluation of the antiplasmodial activity of these peptides against mature P. gallinaceum sporozoites. The Ala screening data indicates that the replacement of either the Ile5 or the His6 residues causes minor effects on the in vitro antiplasmodial activity compared with AII, i.e. AII (88%), [Ala6]‐AII (79%), and [Ala5]‐AII (75%). Analogs [Ala3]‐AII, [Ala1]‐AII, and AII‐NH2 showed antiplasmodial activity around 65%, whereas the activity of the [Ala8]‐AII, [Ala7]‐AII, [Ala4]‐AII, and [Ala2]‐AII analogs is lower than 45%. Circular dichroism data suggest that AII and the most active analogs adopt a β‐fold conformation in different solutions. All AII analogs, except [Ala4]‐AII and [Ala8]‐AII, show contractile responses and interact with the AT1 receptor, [Ala5]‐AII and [Ala6]‐AII. In conclusion, this approach is helpful to understand the contribution of each amino acid residue to the bioactivity of AII, opening new perspectives toward the design of new sporozoiticidal compounds. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
Abstract: Stimulation of rat parotid acinar cells by the tachykinin neurokinin (NK) 1 receptor agonist substance P (SP) resulted in a significant reduction in the initial accumulation of cytosolic myo-[3H]inositol. This effect was rapid, because a reduction of ~15% could be seen already at 30 s, with the maximal effect (~45%) being observed at 15 min. The response to SP stimulation Was temperature dependent, because at 4°C no reduction was found, jln addition, at 4°C, cytosolic myo-[3H]inositol represented only 10% of the labeled inositol accumulated at 37°C. The SP-induced reduct on in cytosolic ravo[3H]inositol accumulation was concentration dependent; the EC50 obtained for SP was 5.8 ± 2.5 nM. Spantide [N Arg1, D-Trp79, Leu]SP), a SP antagonist, used at a concentration oif 105 A/, gave a competitive shift of the dose-response curve to SP. Various tachykinins and their analogs were evaluated for their ability to reduce cytosolic mvo-[3H]inositol. [L-Pro9]SP and SP methyl ester, two highly selective agonists of NK1 receptors, reduced the initial accumulation of myo-H]inositol with EQo values of 2.3 and 67.0 nM, respectively. Long SP C-terminal fragments were more potent than shorter ones. SP N-terminal fragments and SP free acid were -without effect. [Pro7]NKB, a selective NKB analog, had no effect. The rank order of potency of mammalian tachykinins was SP > NKA > NKB. These findings and the close correlation between EC50 values and IC50 values obtained in binding studies implicate the NK 1 receptor. In addition, stimulation of muscarinic receptors by carbachol alscp resulted in a reduction in level of cytosolic mjw-[3H]inositol, with this effect being reversed by atropine. Moreover, atropine was unable tjo alter the SP-induced reduction in cytosolic myo-[3H]inositol accumulation. Other neurotransmitters, such as glutamic acid, serotonin, chplecystokinin, neurotensin, bradykinin, and neuropeptide Y, were without effect on initial cytosolic myo-[3H]inositol accumulation. In conclusion, NK1 and muscarinic receptors seem to regulate the membrane transport of inositol in acinar cells of the rat parotid gland. Measurement of the initial accumulation of cytosolic myo-[3H]inositol in this tissue could profitably be adopted as a very simple, rapid, [sensitive, and specific biochemical procedure for screening the activity of potential agonists and antagonists at NK1 receptors.  相似文献   

17.
The binding and biological activities of neurotensin and two analogues, [Trp11]-neurotensin and xenopsin, in which a tryptophan replaces the neurotensin residue Tyr11, were compared in rat and guinea-pig. The binding activity of the three peptides was measured as their ability to inhibit the binding of [3H]neurotensin to rat and guinea-pig brain synaptic membranes. Their biological activities were measured as their effects on the contractility of rat and guinea-pig ileal smooth muscle preparations. In binding as well as biological assays, it was found that [Trp11]-neurotensin and xenopsin were as potent as neurotensin in the rat. In contrast, the two analogues were about 10 times less potent than neurotensin in the guinea-pig. These findings reveal differences between rat and guinea-pig neurotensin receptors. Such species-related differences in neurotensin receptors should be considered when comparing the activity of neurotensin analogues in assays using tissue preparations from various animal species.  相似文献   

18.
Two somatostatin analogues, [99mTc]Demotide and [99mTc]Demotate 4, were compared with [99mTc]Demotate 1, a previously reported somatostatin receptor subtype 2 (sst2) targeting tracer. Conjugates were prepared by coupling an open‐chain tetraamine chelator to D ‐Phe1 of [Tyr3]‐octreotide or [Tyr3]‐octreotate, respectively, via a p‐benzylaminodiglycolic acid spacer adopting solid‐phase peptide synthesis techniques. Peptide conjugates were collected in a highly pure form after chromatographic purification. Eventually, [99mTc]Demotide and [99mTc]Demotate 4 were obtained in ~1 Ci/µmol specific activity and >96% purity after labeling under alkaline conditions. Demotide and Demotate 4 exhibited similar high binding affinities for the sst2 expressed in AR4‐2J cells with IC50 values 0.16 and 0.10 nM, respectively. The (radio)metallated analogues [99mTc]Demotide and [99mTc]Demotate 4 showed equally high affinities to the sst2 during saturation binding assays in AR4‐2J cell membranes (Kds 0.08 and 0.07 nM, respectively). During incubation at 37 °C with AR4‐2J cells, the radiopeptides internalized effectively via a receptor‐mediated process, with [99mTc]Demotate 4 exhibiting a faster internalization rate than [99mTc]Demotide. After injection in athymic mice bearing sst2‐expressing AR4‐2J tumors, the radiotracers showed high and specific uptake in the tumor (>25%ID/g at 1 h) and in the sst2–positive organs. However, both [99mTc]Demotide and [99mTc]Demotate 4 showed unfavorably higher background activity, especially in the abdomen, in comparison to [99mTc]Demotate 1 and are, therefore, less suited than [99mTc]Demotate 1 for sst2‐targeted tumor imaging in man. Copyright © 2005 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Antimicrobial peptides are recognized candidates with pharmaceutical potential against epidemic emerging multi‐drug resistant bacteria. In this study, we use nuclear magnetic resonance spectroscopy and molecular dynamics simulations to determine the unknown structure and evaluate the interaction with dodecylphosphatidylcholine (DPC) and sodium dodecylsulphate (SDS) micelles with three W6‐Hylin‐a1 analogs antimicrobial peptides (HyAc, HyK, and HyD). The HyAc, HyK, and HyD bound to DPC micelles are all formed by a unique α‐helix structure. Moreover, all peptides reach the DPC micelles' core, which thus suggests that the N‐terminal modifications do not influence the interaction with zwiterionic surfaces. On the other hand, only HyAc and HyK peptides are able to penetrate the SDS micelle core while HyD remains always at its surface. The stability of the α‐helical structure, after peptide‐membrane interaction, can also be important to the second step of peptide insertion into the membrane hydrophobic core during permeabilization. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
The subject of our studies was the synthesis, biological evaluation, and conformational studies of insect tridecapeptide alloferon (H‐His‐Gly‐Val‐Ser‐Gly‐His‐Gly‐Gln‐His‐Gly‐Val‐His‐Gly‐OH) and its analogues such as: [des‐His1]‐, [Lys1]‐, [Arg1]‐, and [Ala1]‐alloferon. These peptides were synthesized to check the influence of the His residue at position 1 of the alloferon chain on its antiviral activity. Two aspects of the biological effects of these peptides were determined: (i) the cytotoxicity in vitro in the Vero, LLC‐MK2, and HEp‐2 cell lines, and (ii) the antiviral activity in vitro in respect to DNA and RNA viruses. We found that alloferon inhibited the herpes virus multiplication and failed to affect the coxsackie virus replication, whereas [Lys1]‐alloferon exhibited a high inhibitory action towards both viruses. Moreover, the peptides did not show any cytotoxic activity against the Vero, LLC‐MK2, and HEp‐2 cells. The preliminary circular dichroism conformational studies showed that the peptides investigated seem to prefer an unordered conformation. Copyright © 2011 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