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 15 毫秒
1.
The [epsilon] subunit of the chloroplast ATP synthase functions in part to prevent wasteful ATP hydrolysis by the enzyme. In addition, [epsilon] together with the remainder of the catalytic portion of the synthase (CF1) is required to block the nonproductive leak of protons through the membrane-embedded component of the synthase (CFO). Mutant [epsilon] subunits of the spinach (Spinacia oleracea) chloroplast ATP synthase that lack 5, 11, or 20 amino acids from their N termini ([epsilon]-[delta]5N, [epsilon]-[delta]11N, and [epsilon]-[delta]20N, respectively), were overexpressed as inclusion bodies. Using a procedure that resulted in the folding of full-length, recombinant [epsilon] in a biologically active form, none of these truncated forms resulted in [epsilon] that inhibited the ATPase activity of CF1 deficient in [epsilon], CF1(-[epsilon]). Yet, the [epsilon]-[delta]5N and [epsilon]-[delta]11N peptides significantly inhibited the ATPase activity of CF1(-[epsilon]) bound to CFO in NaBr-treated thylakoids. Although full-length [epsilon] rapidly inhibited the ATPase activity of CF1(-[epsilon]) in solution or bound to CFO, an extended period was required for the truncated forms to inhibit membrane-bound CF1(-[epsilon]). Despite the fact that [epsilon]-[delta]5N significantly inhibited the ATPase activity of CF1(-[epsilon]) bound to CFO, it did not block the proton conductance through CFO in NaBr-treated thylakoids reconstituted with CF1(-[epsilon]). Based on selective proteolysis and the binding of 8-anilino-1-naphthalene sulfonic acid, each of the truncated peptides gained significant secondary structure after folding. These results strongly suggest (a) that the N terminus of [epsilon] is important in its binding to CF1, (b) that CF0 stabilizes [epsilon] binding to the entire ATP synthase, and (c) that the N terminus may play some role in the regulation of proton flux through CFO.  相似文献   

2.
Raman spectra of six synthetic polydeoxyribonucleotide duplexes with different base sequences have been examined in aqueous solutions with different salt or nucleotide concentrations. Detailed conformational differences have been indicated between B and Z forms of poly[d(G-C)] X poly[d(G-C)], between B forms of poly[d(G-C)] X poly[d(G-C)] and poly[d(G-m5C)] X poly[d(G-m5C)], between A and B forms of poly(dG) X poly(dC), between B and "CsF" forms of poly[d(A-T)] X poly[d(A-T)], between B forms of poly[d(A-U)] X poly[d(A-U)] and poly[d(A-T)] X poly[d(A-T)], and between low- and high-salt (CsF) forms of poly(dA) X poly(dT). The Raman spectrum of calf-thymus DNA in aqueous solution was also observed and was compared with the Raman spectra of its fibers in A, B, and C forms.  相似文献   

3.
Geometry optimization and energy calculations have been performed at the density functional B3LYP/LANL2DZ level on hydrogen sulfide (HS-), dihydrogensulfide (H2S), thiomethanolate (CH3S-), thiomethanol (CH3SH), thiophenolate (C6H5S-), methoxyde (CH3O-), methanol (CH3OH), formiate (HCOO-), acetate (CH3COO-), carbonate (CO3(2-)), hydrogen carbonate (HCO3-), iminomethane (NH=CH2), [ZnS], [ZnS2]2-, [Zn(HS)]+, [Zn(H2S)]2+, [Zn(HS)4]2-, [Zn(CH3S)]+, [Zn(CH3S)2], [Zn(CH3S)3]-, [Zn(CH3S)4]2-, [Zn(CH3SH)]2+, [Zn(CH3SCH3)]2+, [Zn(C6H5S)]+, [Zn(C6H5S)2], [Zn(C6H5S)3]-, [Zn(HS)(NH=CH2)2]+, [Zn(HS)2(NH=CH2)2], [Zn(HS)(H2O)]+, [Zn(HS)(HCOO)], [Zn(HS)2(HCOO)]-, [Zn(CH3O)]+, [Zn(CH3O)2], [Zn(CH3O)3]-, [Zn(CH3O)4]2, [Zn(CH3OH)]2+, [Zn(HCOO)]+, [Zn(CH3COO)]+, [Zn(CH3COO)2], [Zn(CH3COO)3]-, [Zn(CO3)], [Zn(HCO3)]+, and [Zn(HCO3)(Imz)]+ (Imz, 1,3-imidazole). The computed Zn-S bond distances are 2.174A for [ZnS], 2.274 for [Zn(HS)]+, 2.283 for [Zn(CH3S)]+, and 2.271 for [Zn(C6H5S)]+, showing that sulfide anion forms stronger bonds than substituted sulfides. The nature of the substituents on sulfur influences only slightly the Zn-S distance. The optimized tetra-coordinate [Zn(HS)2(NH=CH2)2] molecules has computed Zn-S and Zn-N bond distances of 2.392 and 2.154A which compare well with the experimental values at the solid state obtained via X-ray diffraction for a number of complex molecules. The computed Zn-O bond distances for chelating carboxylate derivatives like [Zn(HOCOO)]+ (1.998A), [Zn(HCOO)]+ (2.021), and [Zn(CH3COO)]+ (2.001) shows that the strength of the bond is not much influenced by the substituent on carboxylic carbon atom and that CH3- and HO- groups have very similar effects. The DFT analysis shows also that the carboxylate Ligand has a preference for the bidentate mode instead of the monodentate one, at least when the coordination number is small.  相似文献   

4.
Six [Pen(6)]oxytocin analogs were synthesized by substituting penicillamine for cysteine in oxytocin, [Mpa(1)]oxytocin, [dPen(1)]oxytocin, [5-t-BuPro(7)]oxytocin, [Mpa(1), 5-t-BuPro(7)]oxytocin and [dPen(1), 5-t-BuPro(7)]oxytocin. When tested in the uterotonic test in vitro [Pen(6)]oxytocin, [Pen(6), 5-t-BuPro(7)]oxytocin, [Mpa(1), Pen(6)]oxytocin and [Mpa(1), Pen(6), 5-t-BuPro(7)]oxytocin, all were found to possess both agonistic and antagonistic properties. Their agonistic potency ranged from negligible (0.08 IU/mg) to low (5.85 IU/mg) and their antagonistic potency (pA2) was estimated to range from 6.6 to 7.9. [dPen(1), Pen(6)]Oxytocin and [dPen(1), Pen(6), 5-t-BuPro(7)]oxytocin were found to be pure antagonists with similarly high pA2 values of approximately 8.2. Replacement of proline by 5-tert-butylproline increased binding affinity by a factor of two in [Pen(6)]oxytocin and had no influence on the binding affinity of [Mpa(1), Pen(6)]oxytocin and [dPen(1), Pen(6)]oxytocin. Assignment of the proton signals for prolyl amide cis- and trans-isomers by NMR experiments in water indicated that the Pen(6)-5-tert-BuPro(7) peptide bond cis-isomer population was augmented relative to the prolyl peptides and measured, respectively, at 20, 35 and 35% in the 5-tert-butylproline(7) analogs of [Pen(6)]oxytocin, [Mpa(1), Pen(6)]oxytocin and [dPen(1), Pen(6)]oxytocin. This augmentation in cis-isomer population was correlated with a 21-fold reduction in the agonistic potency and 2-fold augmentation in antagonistic potency for [Pen(6), 5-t-BuPro(7)]oxytocin relative to [Pen(6)]oxytocin. Augmentation of cis-isomer population was also correlated to reduced agonist potency without effect on antagonism on conversion of [Mpa(1), Pen(6)]oxytocin to [Mpa(1), Pen(6), 5-t-BuPro(7)]oxytocin. In the potent oxytocin antagonist, [dPen(1), Pen(6)]oxytocin, substitution of 5-tert-butylproline for proline augmented the cis-isomer population without affecting antagonistic potency. The synthesis and evaluation of [Pen(6)]oxytocin and [Pen(6), 5-t-BuPro(7)]oxytocin analogs 1-6 indicated that steric interactions influenced agonist and antagonist activity by modifying peptide conformation. Augmentations in the prolyl cis-isomer population caused by 5-tert-butylproline occurred concurrently with enhanced or maintained antagonistic potency and binding affinity and reduced agonistic potency.  相似文献   

5.
Although elevated plasma concentrations of lipoprotein [a] (Lp[a]) are considered to be a risk factor for atherosclerosis, the mechanisms by which Lp[a] mediates its pathogenic effects have not been conclusively determined. The apolipoprotein [a] (apo[a]) component of Lp[a] confers unique structural properties to this lipoprotein, including the ability to bind to lysine residues in biological substrates. It has been shown, however, that only a fraction of plasma Lp[a] (Lp[a]-Lys(+)) binds to lysine-Sepharose in vitro. The nature of the non-lysine-binding Lp[a] fraction in plasma (Lp[a]-Lys(-)) is currently unknown. In the present study, the Lp[a]-Lys(+) fraction was determined in the plasma of six unrelated individuals; the Lp[a]-Lys(+) fraction in these plasma samples ranged from approximately 37 to approximately 48%. Interestingly, purification of the Lp[a] by density gradient ultracentrifugation followed by gel filtration and ion-exchange chromatography resulted in progressive increases in the Lp[a]-Lys(+) fraction. Addition of either purified low density lipoprotein (LDL) or fibronectin to the purified Lp[a] at a 1:1 molar ratio reduced the Lp[a]-Lys(+) fraction (maximal decrease of 34 and 20%, respectively) whereas addition of both fibronectin and LDL to the purified Lp[a] resulted in a further decrease (45% maximally) in this fraction. Similar results were obtained by using a recombinant expression system for apo[a]: addition of a 4-fold molar excess of either LDL or fibronectin to conditioned medium containing metabolically labeled recombinant apo[a] reduced the Lys(+) fraction by 49 and 23%, respectively.Taken together, our data suggest that the lysine-binding heterogeneity of plasma Lp[a] is not primarily an intrinsic property of the lipoprotein, but rather results in large part from its ability to noncovalently associate with abundant plasma components such as LDL and fibronectin. These interactions appear to mask the lysine-binding site in apo[a] kringle IV type 10, which mediates the interaction of Lp[a] with lysine-Sepharose. The contribution of these interactions to the function of Lp[a] in vivo remains to be investigated.  相似文献   

6.
When [3H]inositol-prelabelled N1E-115 cells were stimulated with carbamylcholine (CCh) (100 microM), high K+ (60 mM), and prostaglandin E1 (PGE1) (10 microM), a transient increase in [3H]inositol pentakisphosphate (InsP5) accumulation was observed. The accumulation reached its maximum level at 15 s and had declined to the basal level at 2 min. CCh, high K+, and PGE1 also caused accumulations of [3H]inositol 1,4,5-trisphosphate [Ins(1,4,5)P3], [3H]inositol 1,3,4,6-tetrakisphosphate [Ins(1,3,4,6)P4], and [3H]inositol hexakisphosphate (InsP6). Muscarine and CCh induced accumulations of [3H]Ins(1,4,5)P3, [3H]-Ins(1,3,4,6)P4, [3H]InsP5, and [3H]InsP6 with a similar potency and exerted these maximal effects at 100 microM, whereas nicotine failed to do so at 1 mM. With a slower time course, CCh, high K+, and PGE1 caused accumulations of [3H]-inositol 1,3,4-trisphosphate [Ins(1,3,4)P3] and [3H]inositol 1,3,4,5-tetrakisphosphate [Ins(1,3,4,5)P4]. In an N1E-115 cell homogenate, [3H]Ins(1,4,5)P3, [3H]Ins(1,3,4,5)P4, and [3H]Ins(1,3,4)P3 were converted to [3H]InsP5 through [3H]-Ins(1,3,4,6)P4. The above results indicate that Ins(1,3,4,6)P4, InsP5, and InsP6 are rapidly formed by several kinds of stimulants in N1E-115 cells.  相似文献   

7.
The alpha-form of poly[d(A)].poly[d(T)], observed in fibers at high (greater than 80%) relative humidity, is a 10-fold double-helical structure of pitch 3.2 nm. This new X-ray analysis shows that the two strands of the double helix are of the same kind conformationally and both B-like in containing C-2'-endo-puckered deoxyribose rings. Nevertheless, the two strands are different enough for the overall morphology of the duplex to resemble that of the heteromerous model for the drier (beta) form of poly[d(A)].poly[d(T)] in which one strand has C-2'-endo rings and the other C-3'-endo. Since the orientations of the bases in poly[d(A)].poly[d(T)] are persistently different from those of classical B-DNA it is likely that there will be local bending (about 10 degrees) at the junctions between general sequence tracts and the oligo[d(A)].oligo[d(T)] tracts that occur in some native DNAs. The conclusions about the structure of alpha-poly[d(A)].poly[d(T)] are reinforced by independent analyses of similar X-ray diffraction patterns from poly[d(A)].poly[d(U)] and poly[d(A-I)].poly[d(C-T)].  相似文献   

8.
We describe two sets of plasmid-plasmid interactions in the yeast Saccharomyces cerevisiae. [HOK], [EXL], [NEX], and [KIL-k1] are genetically defined plasmids, and M1 and L are biochemically defined double-stranded RNA plasmids. We show that (i) [HOK], [NEX], and the abundance of L are related, and (ii) under submaximal curing conditions, all colonies retaining M1 also retain L. There are three pieces of evidence that either [NEX] required [HOK] for replication or [NEX] and [HOK] are on the same plasmid. The evidence is as follows. (i) The great majority of strains containing [HOK] also contain [NEX]. However, two [HOK] [NEX-o] strains do exist. (ii) Growth at 39 degrees C or growth at 34 degrees C with 3% ethanol or 2-propanol cures [HOK] and [NEX]. In a [HOK] [NEX] strain, the two plasmids are always co-cured. (iii) [HOK] and [NEX] are both maintained in mak4, mak6, and mak27 strains (mak = maintenance of [KIL-k1]), but not in mak3, mak10, and pet18 strains. Strains containing [HOK] and [NEX] have about fourfold more L double-stranded RNA than their isochromosomal, cured derivatives. In addition, a cytoductant which has acquired [HOK] and [NEX] has fourfold more L than its parent. These results are consistent with either [HOK] being a form of L or [HOK] increasing the copy number of L. Using a K1 killer strain in which L, as well as M1, could be cured by growth at 38 degrees C, we examined the distribution of loss of M1 and L under conditions giving 98% M-o colonies and at least 50% L-o colonies. No M1L-o colonies were observed, supporting the previous suggestion by others that M1 requires L.  相似文献   

9.
Changes in [Mg2+] in a millimolar range have a significant inverse effect on the Ca2+- (or Sr2+)activated tension generation of skeletal muscle fibers. Single frog (Rana pipiens) semitendinosus muscle fibers were "skinned" (sarcolemma removed) and contracted isometrically in bathing solutions of varying [Ca2+] or [Sr2+] and [Mg2+] but a constant pH, [MgATP2-], [K+], [CP2-], [CPK], and ionic strength. Ca2+- (or Sr2+- )activated steady-state tensions were recorded for three [Mg2+]'s: 5 X 10(-5)M, 1 X 10(-3) M, and 2 X 10(-3) M; and these tensions were expressed as the percentages of maximum tension generation of the fibers for the same [Mg2+]. Maximum tension was not affected by [Mg2+] within Ca2+-activating or Sr2+-activating sets of solutions; however, the submaximum Ca2+-(or Sr2+)activated tension is strongly affected in an inverse fashion by increasing [Mg2+]. Mg2+ behaves as a competitive inhibitor of Ca2+ and also affects the degree of cooperativity in the system. At [Mg2+] = 5 X 10(-5)M the shape of tension versus [Ca2+] (or [Sr2+]) curve showed evidence of cooperativity of Ca2+ (or Sr2+) binding or activation of the contractile system. As [Mg2+] increased, the apparent affinity for Ca2+ or Sr2+ and cooperativity of the contractile system declined. The effect on cooperativity suggests that as [Mg2+] decreases a threshold for Ca2+ activation appears.  相似文献   

10.
(1)N-4-Azido-2-nitrophenyl-gamma-[3H]aminobutyryl-AdoPP[NH] P(NAP4-AdoPP[NH]P) a photoactivable derivative of 5-adenylyl imidodiphosphate (AdoPP[NH]P), was synthesized. (2) Binding of [3H]NAP4-AdoPP[NH]P to soluble ATPase from beef heart mitochondria (F1) was studied in the absence of photoirradiation, and compared to that of [3H]AdoPP[NH]P. The photoactivable derivative of AdoPP[NH]P was found to bind to F1 with high affinity, like AdoPP[NH]P. Once [3H]NAP4-AdoPP[NH]P had bound to F1 in the dark, it could be released by AdoPP[NH]P, ADP and ATP, but not at all by NAP4 or AMP. Furthermore, preincubation of F1 with unlabeled AdoPP[NH]P, ADP, or ATP prevented the covalent labeling of the enzyme by [3H]NAP4-AdoPP[NH]P upon photoirradiation. (3) Photoirradiation of F1 by [3H]NAP4-AdoPP[NH]P resulted in covalent photolabeling and concomitant inactivation of the enzyme. Full inactivation corresponded to the binding of about 2 mol [3H]NAP4-AdoPP[NH]P/mol F1. Photolabeling by NAP4-AdoPP[NH]P was much more efficient in the presence than in the absence of MgCl2. (4) Bound [3H]NAP4-AdoPP[NH]P was localized on the alpha- and beta- subunits of F1. At low concentrations (less than 10 microM), bound [3H]NAP4-AdoPP[NH]P was predominantly localized on the alpha-subunit; at concentrations equal to, or greater than 75 microM, both alpha- and beta-subunits were equally labeled. (5) The extent of inactivation was independent of the nature of the photolabeled subunit (alpha or beta), suggesting that each of the two subunits, alpha and beta, is required for the activity of F1. (6) The covalently photolabeled F1 was able to form a complex with aurovertin, as does native F1. The ADP-induced fluorescence enhancement was more severely inhibited than the fluorescence quenching caused by ATP. The precentage of inactivation of F1 was virtually the same as the percentage of inhibition of the ATP-induced fluorescence quenching, suggestion that fluorescence quenching is related to the binding of ATP to the catalytic site of F1.  相似文献   

11.
High-fat diets induce weight gain and fatty liver in wild-type mice. Schistosomiasis mansoni infection also promotes hepatic injury. This study was designed to quantify hepatic alterations in schistosomiasis mansoni-infected mice fed a high fat-rich chow compared to mice fed a standard rodent chow, using stereology. Female SW mice fed each either high-fat diet (29% lipids) or standard chow (12% lipids) over 8 months, and then were infected with Schistosoma mansoni cercariae. Four experimental groups were studied: infected mice fed a high-fat diet (IHFC) or standard chow (ISC), uninfected mice fed a high-fat diet (HFC) or standard chow (SC). Mice were sacrificed during early infection (9 weeks from exposure). The following hepatic biometry and the stereology parameters were determined: volume density (hepatocytes [h], sinusoids [s], steatosis [st] and hepatic fibrosis [hf]); numerical density (hepatocyte nuclei - Nv[h]); absolute number of total hepatocyte N[h], normal hepatocyte N[nh], and binucleated hepatocyte N[bh], percentage of normal hepatocyte P[nh] and binucleated hepatocyte P[bh]. IHFC and HFC groups exhibited TC, HDL-C, LDL-C, and body mass significantly greater (p < 0.05) than control group. No significant differences were found regards liver volume (p = 0.07). Significant differences were observed regards P[nh] (p = 0.0045), P[bh] (p = 0.0045), Nv[h] (p = 0.0006), N[h] (p = 0.0125), N[bh] (p = 0.0164) and N[nh] (p = 0.0078). IHFC mice group presented 29% of binucleated hepatocytes compared to HFC group (19%), ISC group (17%) and SC (6%). Volume density was significantly different between groups: Vv[h] (p = 0.0052), Vv[s] (p = 0.0025), Vv[st] (p = 0.0004), and Vv[hf] (p = 0.0007). In conclusion, schistosomiasis mansoni infection with concurrent high-fat diet promotes intensive quantitative changes in hepatic structure, contributing to an increasing on hepatic regeneration.  相似文献   

12.
4-aminopyridine (4-AP) is a voltage-sensitive K+-channel blocker extensively used in in vitro experiments as a depolarizing agent for the release of glutamate (GLU). This research investigated whether 4-AP could be used in in vivo experiments using microdyalisis. For that, the effects of 4-AP on the extracellular concentrations of glutamate (GLU), glutamine (GLN), taurine (TAU) and citrulline (CIT) in striatum of the freely moving rat were investigated. The effects of 4-AP were compared with those produced by perfusion with a high K+ (100 mM) medium. Intrastriatal perfusion with 4-AP (1, 5 and 10 mM) produced no effects on extracellular [GLU], [TAU] and [CIT], but decreased extracellular [GLN]. Perfusion with a high K+ (100 mM) medium increased extracellular [GLU] and [TAU], decreased extracellular [GLN], and had no effects on [CIT]. To test whether the lack of effects of 4-AP on extracellular [GLU] was due to GLU uptake mechanisms, 4-AP was perfused after a previous inhibition of GLU uptake with L-trans-pyrrolidine-2,4-dicarboxylic acid (PDC). Under the effects of PDC (1 mM), 4-AP (1 mM) had no effects on extracellular [GLU], [TAU] and [CIT], but decreased extracellular [GLN]. These results show that 4-AP decreased extracellular [GLN] but failed to produce a significant release of GLU in striatum of the freely moving rat. Thus, 4-AP can not be used as a depolarizing agent for stimulating the release of GLU in in vivo studies using microdialysis.  相似文献   

13.
The relationship between changes in the intracellular free Ca2+ concentration, [Ca2+]i, and the initiation of proliferation of murine B cells after the addition of mitogens and activators was studied. The effects of lipopolysaccharide (LPS), 12-O-tetradecanoyl phorbol-13-acetate (TPA), rabbit IgG antimouse Fab (IgG RAM Fab), and its F(ab')2 fragment (F(ab')2 anti-Fab) on the [Ca2+]i were measured using the fluorescent calcium indicator Fura-2. In parallel experiments, DNA and/or RNA synthesis were measured by assaying [3H]thymidine and/or [3H]uridine uptake. LPS stimulated a 20-120 X increase in the [3H]thymidine uptake, and a 3-7 X increase in [3H]uridine uptake without inducing any change in the [Ca2+]i. TPA induced a marginal increase in [3H]thymidine and [3H]uridine uptake, without effecting any change in the [Ca2+]i. In contrast, low doses of IgG RAM Fab produced a triphasic change in the [Ca2+]i, but had no effect on the [3H]thymidine or [3H]uridine uptake, even at much higher concentrations. Similarly, low doses of the F(ab')2 fragment induced sizable increases in the [Ca2+]i without affecting the [3H]nucleoside uptake. However, higher concentrations of F(ab')2 anti Fab increased the [3H]thymidine uptake and [3H]uridine uptake, while also increasing the [Ca2+]i. Significantly, pretreating the cells with TPA for 3 min virtually abolished the [Ca2+]i increase induced by IgG RAM Fab while simultaneously potentiating an increase in the IgG RAM Fab-induced [3H]thymidine uptake 85-fold. In the presence of TPA, IgG RAM Fab also induced a 2- to 30-fold increase in [3H]uridine uptake. Similarly, TPA virtually abolished the [Ca2+]i increase induced by the F(ab')2 anti-Fab fragment, yet it stimulated a F(ab')2 anti-Fab-induced uptake of [3H]thymidine and [3H]uridine by 120 and 10 times, respectively.  相似文献   

14.
Solid-state (13)C NMR spectroscopy was used to investigate the three-dimensional structure of melittin as lyophilized powder and in ditetradecylphosphatidylcholine (DTPC) membranes. The distance between specifically labeled carbons in analogs [1-(13)C]Gly3-[2-(13)C]Ala4, [1-(13)C]Gly3-[2-(13)C]Leu6, [1-(13)C]Leu13-[2-(13)C]Ala15, [2-(13)C]Leu13-[1-(13)C]Ala15, and [1-(13)C]Leu13-[2-(13)C]Leu16 was measured by rotational resonance. As expected, the internuclear distances measured in [1-(13)C]Gly3-[2-(13)C]Ala4 and [1-(13)C]Gly3-[2-(13)C]Leu6 were consistent with alpha-helical structure in the N-terminus irrespective of environment. The internuclear distances measured in [1-(13)C]Leu13-[2-(13)C]Ala15, [2-(13)C]Leu13-[1-(13)C]Ala15, and [1-(13)C]Leu13-[2-(13)C]Leu16 revealed, via molecular modeling, some dependence upon environment for conformation in the region of the bend in helical structure induced by Pro14. A slightly larger interhelical angle between the N- and C-terminal helices was indicated for peptide in dry or hydrated gel state DTPC (139 degrees -145 degrees ) than in lyophilized powder (121 degrees -139 degrees ) or crystals (129 degrees ). The angle, however, is not as great as deduced for melittin in aligned bilayers of DTPC in the liquid-crystalline state (approximately 160 degrees ). The study illustrates the utility of rotational resonance in determining local structure within peptide-lipid complexes.  相似文献   

15.
The interaction of three types of chromium(III) complexes, [Cr(salen) (H2O2]+, [Cr(en)3]3+, and [Cr(EDTA) (H2O)]- with AGP has been investigated. [Cr(salen) (H2O2]+, [Cr(en)3]3+ and [Cr(EDTA) (H2O]- bind to Human alpha1-acid glycoprotein with a protein:metal ratio of 1:8, 1:6, and 1:4, respectively. The binding constant, K(b) was estimated to be 1.37 +/- 0.12 x 10(5) M(-1), 1.089 +/- 0.05 x 10(5) M(-1) and 5.3 +/- 0.05 x 10(4) M(-1) for [Cr(salen) (H2O2]+, [Cr(en)3]3+, and [Cr(EDTA) (H2O)]-, respectively. [Cr(en)3]3+ has been found to induce structural transition of AGP from the native twisted beta sheet to a more compact alpha-helix. The complexes, [Cr(salen) (H2O2]+ and [Cr(EDTA) (H2O]-, in the presence of H2O2, have been found to bring about nonspecific cleavage of AGP, whereas [Cr(en)3]3+ does not bring about any protein damage. Treatment of [Cr(salen) (H2O)2]+-protein adduct with iodosyl benzene on the other hand led to site specific cleavage of the protein. These results clearly demonstrate that protein damage brought about by chromium(III) complexes depends on the nature of the coordinated ligand, nature of the metal complex, and the nature of the oxidant.  相似文献   

16.
In this study, we created LDL receptor (LDLr) defective (WHHL) transgenic rabbits expressing human apo[a] to examine whether LDLr mediates the Lp[a] clearance from the plasma. By crossbreeding WHHL rabbits with human apo[a] transgenic rabbits, we obtained two groups of human apo[a] transgenic rabbits with defective LDLr functions: apo[a](1/0) WHHL heterozygous (LDLr(+/-) and apo[a](+/0) WHHL homozygous (LDLr(-/-) rabbits. The lipid and lipoprotein levels of human apo[a] WHHL rabbits were compared to those of human apo[a] transgenic rabbits with normal LDLr functions (LDLr(+/+). The apo[a] production rate was evaluated by analyzing apo[a] mRNA expression in the liver, the major site for apo[a] synthesis in transgenic rabbits. We found that pre-beta lipoproteins were markedly increased accompanied by a 2-fold increase in the plasma Lp[a] in apo[a](+/0)/LDLr(+/-) rabbits and a 4.2-fold increase in apo[a](+/0)/LDLr(-/-) rabbits compared with that in apo[a](+/0) rabbits with normal LDLr function. In apo[a](+/0)/LDLr(-/-) rabbits, there was a marked increase in plasma total cholesterol and triglycerides, as was found in their counterpart non-transgenic WHHL rabbits. Northern blot analysis revealed that hepatic apo[a] expression in WHHL transgenic rabbits was similar to that in LDLr(+/+) transgenic rabbits, suggesting the accumulation of plasma Lp[a] in WHHL transgenic rabbits was not due to increased apo[a] synthesis.In conclusion, absence of a functional LDLr leads to a marked accumulation of plasma Lp[a] in human apo[a] transgenic WHHL rabbits and LDLr may participate in the catabolism of Lp[a] in rabbits.  相似文献   

17.
Factors influencing the binding of tetracationic porphyrin derivatives to DNA have been comprehensively evaluated by equilibrium dialysis, stopped-flow kinetics, etc., for mesotetrakis (4-N-methylpyridiniumyl)porphyrin [TMpyP (4)]. Technical difficulties have previously precluded a comprehensive study of metalloporphyrins. Since electrostatic interactions with the DNA and metal derivatization of the porphyrins have important consequences, we have investigated in greater detail two isomers of TMpyP (4) (meso-tetrakis(3-N-methylpyridiniumyl)porphyrin, [TMpyP (3)] and meso-tetrakis(2-N-methylpyridiniumyl)porphyrin [TMpyP (2)]) in which the position of the charged centers has been varied. A comprehensive study of the Cu(II) derivatives, e.g., CuTMpyP (4), was possible since the difficulties encountered previously with Ni(II) compounds were not a problem with Cu(II) porphyrins [J. A. Strickland, L. G. Marzilli, M. K. Gay, and W. D. Wilson (1988) Biochemistry 27, 8870-8878]. At 25 degrees C, the apparent equilibrium constants [Kobs] decreased with increasing [Na+] for all porphyrins. The Kobs values were comparable for TMpyP (4) and TMpyP (3) binding to either polyd(G-C).polyd(G-C) [poly[d(G-C)2]] or poly[d(A-T)].poly[d(A-T)] [poly[d(A-T)2]]. For the copper(II) porphyrins, the Kobs values were about fivefold greater. The Kobs value for CuTMpyP (2) binding to poly[d(G-C)2] was too small to measure under typical salt conditions; however, Kobs for binding to poly[d(A-T)2] was about two orders of magnitude smaller than those found for CuTMpyP (4) or CuTMpyP (3). Application of the condensation theory for polyelectrolytes suggests about three charge interactions when CuTMpyP (4), CuTMpyP (3), and TMpyP (3) bind to poly[d(G-C)2] or poly[d(A-T)2], a result comparable to that reported for TMpyP (4). At 20 degrees C and 0.115 M [Na+], incorporation of copper decreased the rates of dissociation from poly[d(A-T)2] by a 100-fold compared to those reported for TMpyP (4) but had little effect on the rates of dissociation from poly[d(G-C)2]. Also, movement of the H3CN+ group from the fourth to the third position of the pyridinium ring enhanced the rates of dissociation from poly[d(A-T)2] but decreased the rates of dissociation from poly[d(G-C)2]. From polyelectrolyte theory, the [Na+] dependence of the dissociation rates from poly[d(G-C)2] is consistent with intercalative binding, while that for poly[d(A-T)2] is consistent with an outside binding model. For calf thymus [CT] DNA at 20 degrees C, a greater decrease in the AT than in the GC imino 1H-nmr signal was observed upon addition of CuTMpyP (2), suggesting selective outside binding to the AT regions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
Oota  Y. 《Plant & cell physiology》1973,14(2):307-317
The min-[24(0)] (the IP) and the min-[16(8)] (the min-LD inhr) for Lemna gibba G3 were estimated by extrapolation of thelog. of the flower number vs. the culture time curves at varioustemperatures in the presence or absence of various chelatingagents. With the control culture grown in the M-sucrose medium,both the min- [24(0)] and the min-[16(8)] decreased with therise in temperature until a common plateau of 48 hr was reachedat 26°C. At any given temperature, any metal chelator inany dosage examined equally reduced the min-[16(8)], but themin-[24(0)] was reduced differently depending on the speciesand dose of chelator. In all cultures with chelators, however,both the min-[24(0)] and the min-[16(8)] could not have a lowerlimit of less than 24 hr. Apparently the min-[24(0)] could takeany values, but the min-[16(8)] needed multiples of 24 hr, startingfrom 24 or 48 hr in the presence or absence of chelators. Under[16(8)], but not under [24(0)], the temperaturesensitivity changedrhythmically with a 24-hr period. Presumably the min-[24(0)]is mostly determined by an hourglass-type clock, while the min-[16(8)]is determined by endogenous circadian rhythm. (Received November 2, 1972; )  相似文献   

19.
Accurately predicting plant function and global biogeochemical cycles later in this century will be complicated if stomatal conductance (g(s)) acclimates to growth at elevated [CO(2)], in the sense of a long-term alteration of the response of g(s) to [CO(2)], humidity (h) and/or photosynthetic rate (A). If so, photosynthetic and stomatal models will require parameterization at each growth [CO(2)] of interest. Photosynthetic acclimation to long-term growth at elevated [CO(2)] occurs frequently. Acclimation of g(s) has rarely been examined, even though stomatal density commonly changes with growth [CO(2)]. Soybean was grown under field conditions at ambient [CO(2)] (378 micromol mol(-1)) and elevated [CO(2)] (552 micromol mol(-1)) using free-air [CO(2)] enrichment (FACE). This study tested for stomatal acclimation by parameterizing and validating the widely used Ball et al. model (1987, Progress in Photosynthesis Research, vol IV, 221-224) with measurements of leaf gas exchange. The dependence of g(s) on A, h and [CO(2)] at the leaf surface was unaltered by long-term growth at elevated [CO(2)]. This suggests that the commonly observed decrease in g(s) under elevated [CO(2)] is due entirely to the direct instantaneous effect of [CO(2)] on g(s) and that there is no longer-term acclimation of g(s) independent of photosynthetic acclimation. The model accurately predicted g(s) for soybean growing under ambient and elevated [CO(2)] in the field. Model parameters under ambient and elevated [CO(2)] were indistinguishable, demonstrating that stomatal function under ambient and elevated [CO(2)] could be modelled without the need for parameterization at each growth [CO(2)].  相似文献   

20.
Abstract: Interactions between glutamate (Glu), dopamine (DA), GABA, and taurine (Tau) were investigated in striatum of the freely moving rat by using microdialysis. Intrastriatal infusions of the selective Glu uptake inhibitor l - trans -pyrrolidine-3,4-dicarboxylic acid (PDC) were used to increase the endogenous extracellular [Glu]. Correlations between extracellular [Glu] and extracellular [DA], [GABA], and [Tau], and the effects of a selective blockade of ionotropic Glu receptors, were studied. PDC (1, 2, and 4 m M ) produced a dose-related increase in extracellular [Glu]. At the highest dose of PDC, [Glu] increased from 1.55 ± 0.35 to 6.11 ± 0.88 µ M . PDC also increased extracellular [DA], [GABA], and [Tau]. The increasing [Glu] was correlated significantly with increasing [DA], [GABA], and [Tau]. PDC also decreased extracellular concentrations of DA metabolites 3,4-dihydroxyphenylacetic acid (DOPAC) and 4-hydroxy-3-methoxyphenylacetic acid (HVA). Perfusion with the NMDA-receptor antagonist 3-[( R )-2-carboxypiperazin-4-yl]-propyl-1-phosphonic acid (1 m M ) or the AMPA/kainate-receptor antagonist 6,7-dinitroquinoxaline-2,3-dione (DNQX) (1 m M ) attenuated the increases produced by PDC (4 m M ) on [DA], [GABA], and [Tau], and decreases in [DOPAC] and [HVA]. DNQX also attenuated the increases in [Glu] induced by PDC. These data show that endogenous Glu plays a role in modulating the extracellular concentrations of DA, GABA, and Tau in striatum of the freely moving rat.  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