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 812 毫秒
1.
The chemically synthesized signal peptide (native-sequence signal peptide) of preproparathyroid hormone exhibits signal sequence-like activity by inhibiting the translocation/processing of precursor proteins to their mature forms in an in vitro translation system. In order to prepare a biologically functional radiolabeled form of this peptide, we undertook structure-function studies of the native-sequence signal peptide. Since conventional iodination of peptides is performed under oxidizing conditions, chemical design efforts were focused on the oxidation-labile residues, methionine and cysteine, present in the native sequence. Substitution of the three methionines with norleucine and the single cysteine with alanine yielded a surfur-free analog, [Nle-(-25), Nle-(-21),Nle-(-18),Ala-(-14),D-Tyr-(+1)]pre-proPTH-(-29-+1)amide, which is resistant to oxidation and active in the inhibition of processing assay. An interaction between the signal region and one of the components of the intracellular secretory apparatus, signal recognition particle (SRP), was demonstrated: iodinated sulfur-free analog was cross-linked (using the homo-bifunctional reagent disuccinimidyl suberate) to the 54 kilodalton (kDa) subunit of SRP. The 68 kDa and 72 kDa subunits of SRP were also labeled, but to a lesser extent, by the iodinated peptide.  相似文献   

2.
T Karashima  D Olsen  A V Schally 《Life sciences》1987,40(25):2437-2444
The effect of the repeated or continuous administration of an analog of GH releasing factor (GH-RF), D-Tyr-1, D-Ala-2, Nle-27, GH-RF(1-29)-NH2 (DBO-29), on the subsequent response to this peptide was investigated in pentobarbital-anesthetized male rats. A sc administration of this analog induced a greater and more prolonged GH release than doses 10 times larger of GH-RF(1-29). The GH increase after sc injection of 10 micrograms/kg bw of the analog was greater than that induced by iv administration of 2 micrograms/kg bw of GH-RF(1-44). Pretreatment with 10 micrograms/kg bw of the analog did not affect the pituitary response to a strong stimulus (20 micrograms/kg bw) of GH-RF(1-44), 24 h later. Pretreatment with the analog in doses of 10 micrograms/kg bw, sc twice a day, 5 days per week for 4 weeks, significantly diminished the GH release in response to a sc injection of the analog (10 micrograms/kg bw), as compared to vehicle-pretreated controls (P less than 0.01). On the other hand, a continuous sc administration of 0.4 micrograms/h of the analog to intact rats for 7 days did not result in a decrease in GH response to a sc injection of the analog (10 micrograms/kg bw). Since the rats injected repeatedly with the analog for 4 weeks still showed a marked, although somewhat reduced response, analogs of this type may be useful clinically.  相似文献   

3.
Human pancreatic growth hormone releasing factor (1-29)-amide [hpGRF (1-29)-NH2] and the following analogs: [D-Tyr-1]-hpGRF(1-29)-NH2, [D-Ala-2]-hpGRF(1-29)-NH2, [D-Asp-3]-hpGRF(1-29)-NH2, and [N-Ac-Tyr-1]-hpGRF (1-29)-NH2 were synthesized using solid phase methodology and tested for their ability to stimulate growth hormone (GH) secretion in the rat and the pig in vivo. [D-Ala-2]-hpGRF (1-29)-NH2 was approximately 50 times more potent than the parent molecule in eliciting GH secretion in the rat. The other analogs were less active, but all were more potent than the 1-29 amide in the rat. [D-Tyr-1]-hpGRF(1-29)-NH2 was 10 times more potent, [D-Asp-3]-hpGRF(1-29)-NH2 7 times more potent, and the acetylated molecule approximately 12 times more potent than hpGRF(1-29)-NH2.  相似文献   

4.
Effects of bovine beta-casein (1-28) having a phosphoserine-rich region (Glu14-SerP-Leu-SerP-SerP-SerP-Glu-Glu21) and its chemically synthesized partial fragments on proliferation of lymphocytes and immunoglobulin production were investigated in mouse spleen cell cultures. The parent fragment 1-28 and all fragments containing SerP-Leu-SerP and/or SerP-SerP-SerP had a significant mitogenic effect, stimulated proliferation of lymphocytes induced by lipopolysaccharide, phytohemagglutinin, or concanavalin A, and increased immunoglobulin (IgG + IgM + IgA) or IgA levels in the cell cultures. In contrast, dephosphorylated beta-casein (14-21) and SerP-SerP amide had hardly any immunoregulatory activity. On the other hand, SerP-Leu-SerP amide reacted little with antibodies specific to bovine beta-casein (1-28), but beta-casein (14-21), and SerP-SerP-SerP amide obviously reacted with the antibody. These results confirm that the immunoregulatory activity of casein phosphopeptides is attributable to SerP-X-SerP, which may well be available as a non-allergic food ingredient having an adjuvant activity for mucosal IgA responses.  相似文献   

5.
The gene coding for a Trichosanthes trypsin inhibitor analog (Ala-6-TTI) in which methionine at position 6 was replaced by alanine was synthesized chemically. The synthetic gene was cloned into plasmid pWR590-1 and expressed in Escherichia coli as a fusion protein composed of beta-galactosidase fragment of 590 amino acid residues and (Ala-6)-TTI, with methionine as a connecting residue. After cyanogen bromide cleavage and reduction of the fusion protein, followed by refolding with trypsin-Sepharose 4B as a matrix and affinity chromatography on the immobilized enzyme, the fully active (Ala-6)-TTI was obtained. The trypsin inhibitory activity and amino acid composition of the recombinant (Ala-6)-TTI were consistent with those of the natural one. The (Ala-6)-TTI gene was also cloned into the secretion expression vector, pVT102U/alpha, in Saccharomyces cerevisiae. In order to make the reading frame of the gene compatible with the vector, a nucleotide was inserted into the (Ala-6)-TTI gene via site-directed mutagenesis. The secreted (Ala-6)-TTI was purified and found to be correctly processed at the junction between the alpha-factor leader peptide and (Ala-6)-TTI downstream. Of the two expression systems, the latter is more advantageous in the high yield (greater than 2 mg/liter), easy purification and needlessness of disulfide refolding.  相似文献   

6.
The carboxyl-terminal portions of parathyroid hormone (PTH)-(1--34) and PTH-related peptide (PTHrP)-(1-36) are critical for high affinity binding to the PTH/PTHrP receptor (P1R), but the mechanism of receptor interaction for this domain is largely unknown. To identify interaction sites between the carboxyl-terminal region of PTHrP-(1--36) and the P1R, we prepared analogs of [I(5),W(23),Y(36)]PTHrP-(1--36)-amide with individual p-benzoyl-l-phenylalanine (Bpa) substitutions at positions 22--35. When tested with LLC-PK(1) cells stably transfected with human P1R (hP1R), the apparent binding affinity and the EC(50) of agonist-stimulated cAMP accumulation for each analog was, with the exception of the Bpa(24)-substituted analog, similar to that of the parent compound. The radiolabeled Bpa(23)-, Bpa(27)-, Bpa(28)-, and Bpa(33)-substituted compounds affinity-labeled the hP1R sufficiently well to permit subsequent mapping of the cross-linked receptor region. Each of these peptides cross-linked to the amino-terminal extracellular domain of the P1R: [I(5),Bpa(23),Y(36)]PTHrP-(1-36)-amide cross-linked to the extreme end of this domain (residues 33-63); [I(5),W(23),Bpa(27),Y(36)]PTHrP-(1--36)-amide cross-linked to residues 96--102; [I(5),W(23),Bpa(28),Y(36)]PTHrP-(1--36)- amide cross-linked to residues 64--95; and [I(5),W(23), Bpa(33),Y(36)]PTHrP-(1--36)-amide cross-linked to residues 151-172. These data thus predict that residues 23, 27, 28, and 33 of native PTHrP are each near to different regions of the amino-terminal extracellular receptor domain of the P1R. This information helps define sites of proximity between several ligand residues and this large receptor domain, which so far has been largely excluded from models of the hormone-receptor complex.  相似文献   

7.
8.
The synthesis and biological evaluation of penicillamine(6)-5-tert-butylproline(7)-oxytocin analogs and comparison with their proline(7)-oxytocin counterparts has led to the discovery of two potent oxytocin (OT) antagonists: [dPen(1),Pen(6)]-oxytocin (1, pA(2) = 8.22, EC(50) = 6.0 nM) and [dPen(1),Pen(6),5-tBuPro(7)]-oxytocin (2, pA(2) = 8.19, EC(50) = 6.5 nM). In an attempt to understand the conformational requirements for their biological activity, spectroscopic analyses of 1 and 2 were performed using (1)H NMR, laser Raman and CD techniques. In H(2)O, oxytocin analogs 1 and 2 exhibited cis-isomer populations of 7% and 35%, respectively. Measurement of the amide proton temperature coefficients revealed solvent shielded hydrogens for Gln(4) and Pen(6) in the major trans-conformer of 1 as well as for Gln(4) in the minor cis-conformer of 2. Few long-distance NOEs were observed, suggesting conformational averaging for analogs 1 and 2 in water; moreover, a lower barrier (16.6 +/- 0.2 kcal/mol) for isomerization of the amide N-terminal to 5-tBuPro(7) relative to OT was calculated from measuring the coalescence temperature of the Gly(9) backbone NH signals in the NMR spectra of 2. Observed bands in the Raman spectra of 1 and 2 correspond to C(beta)-S-S-C(beta) dihedral angles of +110-115 degrees and +/-90 degrees , respectively. In water, acetonitrile and methanol, the CD spectra for 1 exhibited a positive maximum around 236-239 nm; in trifluoroethanol, the spectra shifted and a negative maximum was observed at 240 nm. The CD spectra of 2 were unaffected by solvent changes and exhibited a negative maximum at 236-239 nm. The CD and Raman data both suggested that a conformation having a right-handed screw sense about the disulfide and a chi(CS-SC) dihedral angle value close to 115 degrees was favored for analog 1 in water, methanol and acetonitrile, but not trifluoroethanol, where a +/-90 degrees angle was favored. Analog 2 was more resilient to conformational change about the disulfide, and adopted a preferred disulfide geometry corresponding to a +/-90 degrees chi(CS-SC) dihedral angle. Monte Carlo conformational analysis of analogs 1 and 2 using distance restraints derived from NMR spectroscopy revealed two prominent conformational minima for analog 1 with disulfide geometries around +114 degrees and +116 degrees . Similar analysis of analog 2 revealed one conformational minimum with a disulfide geometry around +104 degrees . In sum, the conformation about the disulfide in [dPen(1),Pen(6)]-OT (1) was shown to be contingent on environment and in TFE, adopted a geometry similar to that of [dPen(1),Pen(6),5-tBuPro(7)]-OT (2) which appeared to be stabilized by hydrophobic interactions between the 5-tBuPro(7) (5R)-tert-butyl group, the Leu(8) isopropyl sidechain and the Pen(6)beta-methyl substituents. In light of the conformational rigidity of 2 about the disulfide bond, and the similar geometry adopted by 1 in TFE, a S-S dihedral angle close to +110 degrees may be a prerequisite for their binding at the receptor.  相似文献   

9.
Several analogues of the biologically active fragment of bovine parathyroid hormone (bPTH), based on the sequence of the NH2-terminal 34 amino acids, were prepared by solid phase synthesis and bioassayed in the in vitro adenylyl cyclase assay to provide further information concerning structure-activity relations in parathyroid hormone. In two analogues both methionines of the natural hormone were replaced with the sulfur-free and closely isosteric amino acid norleucine (Nle). The synthetic analogue [Nle-8, Nle-18]bPTH-(1-34) was highly active in the in vitro rat adenylyl cyclase bioassay, thus demonstrating that neither of the methionines, found in the native sequence, is indispensable for biological activity. Tyrosine was substituted for phenylalanine at position 34 in the synthesis of two other hormone analogues, [Try-34]bPTH-(1-34) and [Nle-8,Nle-18,Tyr-34]bPTH-(1-34). Both derivatives were exposed to conventional iodination procedures involving use of the oxidant chloramine T. Although iodination of [Try-34]bPTH-(1-34) resulted in virtually complete loss of biological activity, [Nle-8,Nle-18,Tyr-34]-bPTH-(1-34), which lacks methionine, could be exposed to oxidants and labeled efficiently with iodine with retention of nearly complete biological activity. These findings confirm that the loss of biological activity after oxidation of bPTH, as previously observed with the native hormone, is indeed attributable to the oxidation lability of methionine rather than to any other modifications. This sulfur-free, radioiodinated, biologically active analogue of parathyroid hormone may prove useful in studies of interaction of the hormone with the membrane receptors of target tissues and in studies of the metabolism of parathyroid hormone.  相似文献   

10.
Genomic and cDNA clones encoding portions of a putative catfish parathyroid hormone (PTH) 2 receptor (PTH2R) led to the isolation of a cDNA encoding a full-length zebrafish PTH2R (zPTH2R). The zPTH2R shared 63 and 60% amino acid sequence identity with human and rat PTH2Rs, respectively, 47-52% identity with mammalian and frog PTH/PTHrP receptors (PTH1R), and less than 37% with other members of this family of G protein-coupled receptors. COS-7 cells expressing zPTH2R(43), a 5' splice variant that lacked 17 amino acids in the amino-terminal extracellular domain, showed cAMP accumulation when challenged with [Tyr(34)]hPTH(1-34)-amide (hPTH) (EC(50), 1.64 +/- 0. 95 nM) and [Ile(5),Trp(23),Tyr(36)]hPTHrP-(1-36)-amide ([Ile(5), Trp(23)]hPTHrP) (EC(50), 46.8 +/- 12.1 nM) but not when stimulated with [Tyr(36)]hPTHrP-(1-36)-amide (hPTHrP), [Trp(23), Tyr(36)]hPTHrP-(1-36)-amide ([Trp(23)]hPTHrP), or [Ala(29),Glu(30), Ala(34),Glu(35),Tyr(36)]fugufish PTHrP-(1-36)amide (fuguPTHrP). FuguPTHrP also failed to activate the human PTH2R but had similar efficiency and efficacy as hPTH and hPTHrP when tested with cells expressing the human PTH1R. Agonist-dependent activation of zPTH2R was less efficient than that of zPTH2R(43), and both receptor variants showed no cAMP accumulation when stimulated with either secretin, growth hormone-releasing hormone, or calcitonin. The zPTH2R thus has ligand specificity similar to that of the human homolog, which raises the possibility that a PTH-like molecule exists in zebrafish, species which lack parathyroid glands.  相似文献   

11.
B N Rao  C A Bush 《Biopolymers》1987,26(8):1227-1244
The antifreeze glycopeptide (AFGP-8) from polar cod, B. saida, is a 14-amino acid polypeptide having alternating glycotripeptide sequences of Ala-[Gal(β1 → 3)GalNAc(β1 → O)]-Thr-Pro and Ala-[Gal(β1 → 3)GalNAc(β1 → O)]-Thr-Ala, with alanyl residues at amino and carboxy terminals. Conformational studies of AFGP-8 have been carried out by 1H-nmr and empirical energy calculations to investigate the difference in its antifreeze behavior from that of the more active high-molecular weight AFGP 1-4 of P. borchgrevinki. The 1H-nmr spectra, including the resonances of the exchangeable amide protons, were assigned by two-dimensional correlated spectroscopy (COSY), one-dimensional difference decoupling, and nuclear Overhauser effect (NOE) measurements. For the four threonyl residues, the amide proton coupling constants and the small coupling constants between Hα and Hβ indicate similar conformations, despite significant chemical shift differences. The strong NOE between the α protons and the amide protons of the residue following together with large temperature coefficients of chemical shifts, indicate an extended conformation not consisting of α-helix, turns or bends. Energy computations indicate several low-energy conformations consistent with the observed coupling constants for ?. Among these, a left-handed helical conformation with three repeating residues per turn has been proposed, which is in accordance with the observed NOE between the methyl group of the α-GalNAc and Ala Hβs. While the observed Overhauser effects in the threonyl side chain suggest a certain amount of conformational averaging, the effect involving the acetmido methyl of α-GalNAc and Hβs of Ala indicate that it as is a major conformer. In view of the close similarity between the conformations of AFGP-8 and the more active antifreeze polymer, AFGP 1-4, we propose that the difference in their activities is due to the length of the regular repeating structure with glycosylation at every third amino acid residue, and not due to any fundamental difference in their conformations.  相似文献   

12.
13.
Thermogenic endurance and development of metabolic cold adaptation in birds may critically depend on their ability to synthesize and use fatty acids (FA) as fuel substrates. Hepatic lipogenesis and the capacity to oxidize FA in thermogenic tissues were measured in cold-acclimated (CA) ducklings (Cairina moschata) showing original mechanisms of metabolic cold adaptation in the absence of brown adipose tissue, the specialized thermogenic tissue of rodents. The rate of FA synthesis from [U-(14)C]glucose and from [1-(14)C]acetate, measured in incubated hepatocytes isolated from 5-wk-old thermoneutral (TN; 25 degrees C) or CA (4 degrees C) fed ducklings, was higher than in other species. Hepatic de novo lipogenesis was further increased by cold acclimation with both glucose (+194%) and acetate (+111%) as precursor. Insulin slightly increased (+11-14%) hepatic lipogenesis from both precursors in CA ducklings, whereas glucagon was clearly inhibitory (-29 to -51%). Enhanced de novo lipogenesis was associated with higher (+171%) hepatocyte activity of glucose oxidation and larger capacity (+50 to +100%) of key lipogenic enzymes. The potential for FA oxidation was higher in liver (+61%) and skeletal muscle (+29 to +81%) homogenates from CA than from TN ducklings, suggesting that the higher hepatic lipogenesis may fuel oxidation in thermogenic tissues. Present data underline the high capacity to synthesize lipids from glucose in species like muscovy ducks susceptible to hepatic steatosis. Lipogenic capacity can be further increased in the cold and may represent an important step in the metabolic adaptation to cold of growing ducklings.  相似文献   

14.
The activation volumes for kcat of the carboxypeptidase-Y-catalyzed hydrolysis of ester substrates were slightly negative (-1 to -4 ml/mol), while those for peptide and depsi-peptide analog were highly positive (+10 to +27 ml/mol). These values and the contrasting pH dependences of these two groups of the substrates are explained by a mechanism involving three ionic states of the enzyme and the second stable intermediate (acyl-enzyme). Esters are mostly rate-controlled by the deacylation step and peptides are controlled by both the acylation and the deacylation steps. Pressure increase induced a partial shift of the rate-determining step. Reaction volumes for Km-1 of peptide and depsi-peptide analog showed large and positive values (+16 to +29 ml/mol) which reflects the electrostatic interaction in the substrate recognition by this enzyme.  相似文献   

15.
Anderson WG  Leprince J  Conlon JM 《Peptides》2008,29(8):1280-1286
A bradykinin (BK)-related peptide was isolated from heat-denaturated plasma from an elasmobranch fish, the little skate, Leucoraja erinacea after incubation with porcine pancreatic kallikrein. The primary structure of the peptide (H-Gly-Ile-Thr-Ser-Trp-Leu-Pro-Phe-OH; skate BK) shows limited structural similarity to the mammalian B1 receptor agonist, des-Arg(9)-BK. The myotropic activities of synthetic skate BK, and the analog skate [Arg(9)]BK, were examined in isolated skate vascular and intestinal smooth muscle preparations. Skate BK produced a concentration-dependent constriction of the mesenteric artery (EC(50)=4.37x10(-8)M; maximum response=103.4+/-10.23% of the response to 60mM KCl) but the response to skate [Arg(9)]BK was appreciably weaker (response to 10(-6)M=73.0+/-23.4% of the response to 60mM KCl). Neither the first branchial gill arch nor the ventral aorta responded to either purified peptide. Skate BK also produced a concentration-dependent constriction of intestinal smooth muscle preparations (EC(50)=2.74x10(-7)M; maximum response 31.0+/-12.2% of the response to 10(-5)M acetylcholine). Skate [Arg(9)]BK was without effect on the intestinal preparation. The data provide evidence for the existence of the kallikrein-kinin system in a phylogenetically ancient vertebrate group and the greater potency of skate BK compared with the analog skate [Arg(9)]BK suggests that the receptor mediating vascular responses resembles the mammalian B1 receptor more closely than the B2 receptor.  相似文献   

16.
The major substrate for Ca2+/calmodulin-dependent protein kinase III in mammalian cells is a species of Mr 100,000 that has a primarily cytoplasmic localization. This substrate has now been identified as elongation factor-2 (EF-2), a protein that catalyzes the translocation of peptidyl-tRNA on the ribosome. The amino acid sequence of 18 residues from the N-terminal of the Mr 100,000 CaM-dependent protein kinase III substrate purified from rat pancreas was found to be identical to the N-terminal sequence of authentic rat EF-2 as previously deduced from nucleic acid sequencing of a cDNA (Kohno, K., Uchida, T., Ohkubo, H., Nakanishi, S., Nakanishi, T., Fukui, T., Ohtsuka, E., Ikehara, M., and Okada, Y. (1986) Proc. Natl. Acad. Sci. U.S.A. 83, 4978-4982). CaM-dependent protein kinase III phosphorylated EF-2 in vitro with a stoichiometry of approximately 1 mol/mol on a threonine residue. Amino acid sequencing of the purified tryptic phosphopeptide revealed that this threonine residue lies within the sequence: Ala-Gly-Glu-Thr-Arg-Phe-Thr-Asp-Thr-Arg (residues 51-60 of EF-2). The Mr 100,000 protein was stoichiometrically ADP-ribosylated in vitro by the addition of diphtheria toxin and NAD. The Mr 100,000 protein was photoaffinity labeled with a GTP analog and the protein had an endogenous GTPase activity that could be stimulated by the addition of salt-washed ribosomes. These properties are all characteristic of EF-2. Dephospho-EF-2 could support poly(U)-directed polyphenylalanine synthesis in a reconstituted elongation system when combined with EF-1. In the same system, phospho-EF-2 was virtually inactive in supporting polypeptide synthesis; this effect could be reversed by dephosphorylation of phospho-EF-2. These results suggest that intracellular Ca2+ inhibits protein synthesis in mammalian cells via CaM-dependent protein kinase III-catalyzed phosphorylation of EF-2.  相似文献   

17.
18.
A cDNA clone for the alpha subunit of mouse brain Ca2+/CaM-dependent protein kinase II (CaM-kinase II) was transcribed in vitro and translated in a rabbit reticulocyte lysate system. Inclusion of [35S]methionine in the translation system yielded a single 35S-polypeptide of about 50 kDa. When the translation system was assayed for CaM-kinase II activity, there was a 5-10-fold enrichment of kinase activity which was totally dependent on Ca2+/calmodulin (CaM). Both the 50-kDa 35S-polypeptide and the Ca2+/CaM-dependent protein kinase activity were quantitatively immunoprecipitated by rat brain CaM-kinase II antibody. When the translated wild-type kinase was subjected to autophosphorylation conditions in the presence of Ca2+, CaM, Mg2+, and ATP, the Ca2+-independent activity (assayed in the presence of [ethylenebis(oxyethylenenitrilo)]tetraacetic acid) increased from 5.8 +/- 0.7 to 26.5 +/- 2.1% of total activity (assayed in the presence of Ca2+/CaM). These properties confirm the identity of the kinase translated in vitro as CaM-kinase II. The role of Thr-286 autophosphorylation in formation of the Ca2+-independent activity was investigated by site-directed mutation of Thr-286 to Ala (Ala-286 kinase) and to Asp (Asp-286 kinase). The Ala-286 kinase was completely dependent on Ca2+/CaM for activity prior and subsequent to autophosphorylation. The Asp-286 kinase exhibited 21.9 +/- 0.8% Ca2+-independent activity, and this was not increased by autophosphorylation. These results establish that introduction of negative charge(s) at residue 286, either by autophosphorylation of Thr or by mutation to Asp, is sufficient and necessary to generate the partially Ca2+-independent form of CaM-kinase II.  相似文献   

19.
The level of cis-unsaturated fatty acids in phosphatidylglycerol (PG) from rice leaves was genetically altered from 19.3% in the wild-type to 29.4 and 32.0% in T1 plants segregated with cDNAs for glycerol-3-phosphate acyltransferase of chloroplasts (GPAT; EC 2.3.1.15) from Arabidopsis (+AGPAT plant) and spinach (+SGPAT plant), respectively; and to 21.4% in a non-transformant segregated from +SGPAT plants (-SGPAT plant). In all these plants, O2 evolution from leaves was similar at 25 degrees C and was impaired to a similar extent at 5 and 11 degrees C. However, in parallel with the levels of cis-unsaturated fatty acids in PG, +AGPAT and +SGPAT plants showed less impaired rates of O(2) evolution from leaves than the wild-type and -SGPAT plants at 14 and 17 degrees C. In agreement with this, the fresh weight of 14-day-old seedlings increased to 571 + or - 18, 591 + or - 23, 687 + or - 32 and 705 + or - 31 mg in the wild-type, -SGPAT, +AGPAT and +SGPAT plants, respectively, after 6 weeks at 17/14 degrees C (day/night). These results demonstrate the practical importance of the present technology with GPAT in improvement of the chilling sensitivity of crops.  相似文献   

20.
The amyloid beta-peptide (AbetaP) is the major protein component of brain senile plaques in Alzheimer's disease. The redox state of methionine-35 residue plays a critical role in peptide neurotoxic actions. We used the fragment 31-35 of AbetaP [AbetaP(31-35)], containing a single methionine-35 residue (Met-35), to investigate the relationship between the oxidative state of Met-35 and neurotoxic and pro-apoptotic actions induced by the peptide; in rat cerebellar granule cells (CGC), we compared the effects of AbetaP(31-35), in which the Met-35 is present in the reduced state, with those of a modified peptide with oxidized Met-35 [AbetaP(31-35)Met-35(OX)](,) as well as an AbetaP-derivative with Met-35 substituted by norleucine [AbetaP(31-35)Nle-35]. AbetaP(31-35) induced a time-dependent decrease in cell viability. AbetaP(31-35)Met-35(OX) was significantly less potent, but still induced a significant decrease in cell viability compared to control. No toxic effects were observed after treatment with AbetaP(31-35)Nle-35. AbetaP(31-35) induced a 2-fold increase in bax mRNA levels after 4h, whereas AbetaP(31-35)Met-35(OX) raised bax mRNA levels by 41% and AbetaP(31-35)Nle-35 had no effect. Finally, AbetaP(31-35) caused a 43% increase in caspase-3 activity after 24h; AbetaP(31-35)Met-35(OX) caused only a 18% increase, and AbetaP(31-35)Nle-35 had no effect. These findings suggest that AbetaP(31-35)-induced neurodegeneration in CGC is mediated by a selective early increase in bax mRNA levels followed by delayed caspase-3 activation; the redox state of the single Met-35 residue is crucial in the occurrence and extent of the above phenomena.  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