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 46 毫秒
1.
The promoting activity of polyamine analogs (IV approximately XV) on staphylococcal nuclease with DNA as the substrate was compared with that of natural polyamines (I APPROXIMATELY III): I. NH2(CH2)3NH(CH2)4NH(CH2)3NH2(spermine); II. NH2(CH2)3NH(CH2)3NH(CH2)3NH2(thermine); III. NH2(CH2)4NH2 (putrescine); IV. CN(CH2)2NH(CH2)4NH(CH2)2CN; V. HOOC(CH2)2NH(CH2)4NH(CH2)2COOH; VI. C2H5OOC(CH2)2NH(CH2)4NH(CH2)2COOC2H5; VII. HO(CH2)3NH(CH2)4HH(CH2)3OH; VIII. CH3COHH(CH2)3NH(CH2)4NH(CH2)3NHCOCH3; IX. C2H5NH(CH2)3NH(CH2)4NH(CH2)3NHC2H5; X. NH2(CH2)3S(CH2)4S(CH2)3NH2; XI. NH2(CH2)3NH(CH2)2O(CH2)2NH(CH2)3NH2; XII. NH2(CH2)3NCH3(CH2)4HCH3(CH2)3NH2; XIII. CN(CH2)2NCH3(CH2)4NCH3(CH2)2CN; XIV. (CH3)2N(CH2)3NCH3(CH2)4NCH3(CH2)3N(CH3)2; XV. NH2(CH2)2O(CH2)2NH2 Replacement of the terminal groups by CN, COOH, COOEt, NHAc, NHEt, or N(CH3)2 remarkably decreased the activity. The compound VII with terminal hydroxyl groups had a lower promoting activity at low concentrations, but revealed higher activity at higher concentrations and, in contrast to spermine, no inhibition at all even at very high concentrations. Replacement of both internal amino groups by sulfur or NCH3 decreased the activity. The introduction of an ether bond into the internal methylene groups (compound XI) highly decreased the activity. Based upon these findings the possible relationship between structure and activity is discussed.  相似文献   

2.
The organometallic precursor (NEt(4))(2)[ReBr(3)(CO)(3)] was reacted with bidendate dithioethers (L) of the general formula H(3)C-S-CH(2)CH(2)-S-R (R = -CH(2)CH(2)COOH, CH(2)-C&tbd1;CH) and R'-S-CH(2)CH(2)-S-R' (R' = CH(3)CH(2)-, CH(3)CH(2)-OH, and CH(2)COOH) in methanol to form stable rhenium(I) tricarbonyl complexes of the general composition [ReBr(CO)(3)L]. Under these conditions, the functional groups do not participate in the coordination. As a prototypic representative of this type of Re compounds, the propargylic group bearing complex [ReBr(CO(3))(H(3)C-S-CH(2)CH(2)-S-CH(2)C&tbd1;CH)] Re2 was studied by X-ray diffraction analysis. Its molecular structure exhibits a slightly distorted octahedron with facial coordination of the carbonyl ligands. The potentially tetradentate ligand HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH was reacted with the trinitrato precursor [Re(NO(3))(3)(CO)(3)](2-) to yield a cationic complex [Re(CO)(3)(HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH)]NO(3) Re8 which shows the coordination of one hydroxy group. Re8 has been characterized by correct elemental analysis, infrared spectroscopy, capillary electrophoresis, and X-ray diffraction analysis. Ligand exchange reaction of the carboxylic group bearing ligands H(3)C-S-CH(2)CH(2)-S-CH(2)CH(2)-COOH and HOOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH with (NEt(4))(2)[ReBr(3)(CO)(3)] in water and with equimolar amounts of NaOH led to complexes in which the bromide is replaced by the carboxylic group. The X-ray structure analysis of the complex [Re(CO)(3)(OOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH)] Re6 shows the second carboxylic group noncoordinated offering an ideal site for functionalization or coupling a biomolecule. The no-carrier-added preparation of the analogous (99m)Tc(I) carbonyl thioether complexes could be performed using the precursor fac-[(99m)Tc(H(2)O)(3)(CO)(3)](+), with yields up to 90%. The behavior of the chlorine containing (99m)Tc complex [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 in aqueous solution at physiological pH value was investigated. In saline, the chromatographically separated compound was stable for at least 120 min. However, in chloride-free aqueous solution, a water-coordinated cationic species Tc1a of the proposed composition [(99m)Tc(H(2)O)(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))](+) occurred. The cationic charge of the conversion product was confirmed by capillary electrophoresis. By the introduction of a carboxylic group into the thioether ligand as a third donor group, the conversion could be suppressed and thus the neutrality of the complex preserved. Biodistribution studies in the rat demonstrated for the neutral complexes [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 and [(99m)TcCl(CO)(3)(CH(2)-S-CH(2)CH(2)-S-CH(2)-C&tbd1;CH)] Tc2 a significant initial brain uptake (1.03 +/- 0.25% and 0.78 +/- 0.08% ID/organ at 5 min. p.i.). Challenge experiments with glutathione clearly indicated that no transchelation reaction occurs in vivo.  相似文献   

3.
Vanadium pentoxide (V(2)O(5)) is a transition metal derived from the burning of petrochemicals that causes airway fibrosis and remodeling. Vanadium compounds activate many intracellular signaling pathways via the generation of hydrogen peroxide (H(2)O(2)) or other reactive oxygen species. In this study, we investigated the regulation of heparin-binding epidermal growth factor-like growth factor (HB-EGF) in human lung fibroblasts after V(2)O(5) treatment. V(2)O(5)-induced HB-EGF mRNA expression was abolished by N-acetyl-l-cysteine, suggesting an oxidant-mediated effect. Exogenous H(2)O(2) (>10 microM) mimicked the effect of V(2)O(5) in upregulating HB-EGF expression. Fibroblasts spontaneously released low levels of H(2)O(2) (1-2 microM), and the addition of V(2)O(5) depleted the endogenous H(2)O(2) pool within minutes. V(2)O(5) caused a subsequent increase of H(2)O(2) into the culture medium at 12 h. However, the burst of V(2)O(5)-induced H(2)O(2) occurred after V(2)O(5)-induced HB-EGF mRNA expression at 3 h, indicating that the V(2)O(5)-stimulated H(2)O(2) burst did not mediate HB-EGF expression. Either V(2)O(5) or H(2)O(2) activated ERK-1/2 and p38 MAP kinase. Inhibitors of the ERK-1/2 pathway (PD-98059) or p38 MAP kinase (SB-203580) significantly reduced either V(2)O(5)- or H(2)O(2)-induced HB-EGF expression. These data indicate that vanadium upregulates HB-EGF via ERK and p38 MAP kinases. The induction of HB-EGF is not related to a burst of H(2)O(2) in V(2)O(5) treated cells, yet the action of V(2)O(5) in upregulating HB-EGF is oxidant dependent and could be due to the reaction of V(2)O(5) with endogenous H(2)O(2).  相似文献   

4.
The binding of carbon dioxide by horse haemoglobin   总被引:15,自引:7,他引:8  
1. Three modified horse haemoglobins have been prepared: (i) alpha(c) (2)beta(c) (2), in which both the alpha-amino groups of the alpha- and beta-chains have reacted with cyanate, (ii) alpha(c) (2)beta(2), in which the alpha-amino groups of the alpha-chains have reacted with cyanate, and (iii) alpha(2)beta(c) (2), in which the two alpha-amino groups of the beta-chain have reacted with cyanate. 2. The values of n (the Hill constant) for alpha(c) (2)beta(c) (2), alpha(2)beta(c) (2) and alpha(c) (2)beta(2) were (respectively) 2.5, 2.0 and 2.6, indicating the presence of co-operative interactions between the haem groups for all derivatives. 3. In the alkaline pH range (about pH8.0) all the derivatives show the same charge as normal haemoglobin whereas in the acid pH range (about pH6.0) alpha(c) (2)beta(c) (2) differs by four protonic charges and alpha(c) (2)beta(2), alpha(2)beta(c) (2) by two protonic charges from normal haemoglobin, indicating that the expected number of ionizing groups have been removed. 4. alpha(c) (2)beta(2) and alpha(c) (2)beta(c) (2) show a 25% decrease in the alkaline Bohr effect, in contrast with alpha(2)beta(c) (2), which has the same Bohr effect as normal haemoglobin. 5. The deoxy form of alpha(c) (2)beta(c) (2) does not bind more CO(2) than the oxy form of alpha(c) (2)beta(c) (2), whereas alpha(c) (2)beta(2) and alpha(2)beta(c) (2) show intermediate binding. 6. The results reported confirm the hypothesis that, under physiological conditions, haemoglobin binds CO(2) through the four terminal alpha-amino groups and that the two terminal alpha-amino groups of alpha-chains are involved in the Bohr effect.  相似文献   

5.
It has been claimed that the sole H(2)O(2)-scavenging system in the cyanobacterium Synechococcus sp. PCC 7942 is a cytosolic catalase-peroxidase. We have measured in vivo activity of a light-dependent peroxidase in Synechococcus sp. PCC 7942 and UTEX 625. The addition of small amounts of H(2)O(2) (2.5 microM) to illuminated cells caused photochemical quenching (qP) of chlorophyll fluorescence that was relieved as the H(2)O(2) was consumed. The qP was maximal at about 50 microM H(2)O(2) with a Michaelis constant of about 7 microM. The H(2)O(2)-dependent qP strongly indicates that photoreduction can be involved in H(2)O(2) decomposition. Catalase-peroxidase activity was found to be almost completely inhibited by 10 microM NH(2)OH with no inhibition of the H(2)O(2)-dependent qP, which actually increased, presumably due to the light-dependent reaction now being the only route for H(2)O(2)-decomposition. When (18)O-labeled H(2)O(2) was presented to cells in the light there was an evolution of (16)O(2), indicative of H(2)(16)O oxidation by PS 2 and formation of photoreductant. In the dark (18)O(2) was evolved from added H(2)(18)O(2) as expected for decomposition by the catalase-peroxidase. This evolution was completely blocked by NH(2)OH, whereas the light-dependent evolution of (16)O(2) during H(2)(18)O(2) decomposition was unaffected.  相似文献   

6.
Taking into account published contradictory results concerning the regulation of fatty acid synthase (Fas) by H(2)O(2), we carried out a systematic study where two methods of H(2)O(2) delivery (steady-state and bolus addition) and the effect of a wide range of H(2)O(2) concentrations were investigated. A decrease in Fas activity was observed for cells exposed to 100 and 150μM H(2)O(2) in a steady-state, while a bolus addition of the same H(2)O(2) concentrations did not alter Fas activity. Similar results were observed for the mRNA levels of FAS1, the gene that encodes Fas subunit β. However, the exposure to a steady-state 50μM H(2)O(2) dose lead to an increase in FAS1 mRNA levels, showing a biphasic modulation of Fas by H(2)O(2). The results obtained emphasize that cellular effects of H(2)O(2) can vary over a narrow range of concentrations. Therefore, a tight control of H(2)O(2) exposure, which can be achieved by exposing H(2)O(2) in a steady-state, is important for cellular studies of H(2)O(2)-dependent redox regulation.  相似文献   

7.
Ex vivo ?(13)C, (2)H? NMR spectroscopy allowed to estimate the relative sizes of neuronal and glial glutamate pools and the relative contributions of (1-(13)C) glucose and (2-(13)C, 2-(2)H(3)) acetate to the neuronal and glial tricarboxylic acid cycles of the adult rat brain. Rats were infused during 60 min in the right jugular vein with solutions containing (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose or (2-(13)C, 2-(2)H(3)) acetate only. At the end of the infusion the brains were frozen in situ and perchloric acid extracts were prepared and analyzed by high resolution (13)C NMR spectroscopy (90.5 MHz). The relative sizes of the neuronal and glial glutamate pools and the contributions of acetyl-CoA molecules derived from (2-(13)C, (2)H(3)) acetate or (1-(13)C) glucose entering the tricarboxylic acid cycles of both compartments, could be determined by the analysis of (2)H-(13)C multiplets and (2)H induced isotopic shifts observed in the C4 carbon resonances of glutamate and glutamine. During the infusions with (2-(13)C, 2-(2)H(3)) acetate and (1-(13)C) glucose, the glial glutamate pool contributed 9% of total cerebral glutamate being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (4%), (2-(13)C) acetyl-CoA (3%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (2%). The neuronal glutamate pool accounted for 91% of the total cerebral glutamate being mainly originated from (2-(13)C) acetyl-CoA (86%) and (2-(13)C, 2-(2)H) acetyl-CoA (5%). During the infusions of (2-(13)C, 2-(2)H(3)) acetate only, the glial glutamate pool contributed 73% of the cerebral glutamate, being derived from (2-(13)C, 2-(2)H(3)) acetyl-CoA (36%), (2-(13)C, 2-(2)H) acetyl-CoA (27%) and (2-(13)C) acetyl-CoA (10%). The neuronal pool contributed 27% of cerebral glutamate being formed from (2-(13)C) acetyl-CoA (11%) and recycled (2-(13)C, 2-(2)H) acetyl-CoA (16%). These results illustrate the potential of ?(13)C, (2)H? NMR spectroscopy as a novel approach to investigate substrate selection and metabolic compartmentation in the adult mammalian brain.  相似文献   

8.
The reactions of phosphine Ph(2)P(CH(2))(2)SO(3)Na with Cp(2)M'Cl(2) (M'=Ti, Zr) in aqueous solution give the metallophosphines, Cp(2)Ti(OSO(2)(CH(2))(2)PPh(2))(2) (Cp=cyclopentadienyl) and CpZr(OH)(OSO(2)(CH(2))(2)PPh(2))(2). These react with CODM"Cl(2) (M"=Pd, Pt) (COD=1,5-cyclooctadiene) in dichloromethane to give heterobimetallic complexes Cp(2)Ti(OSO(2)(CH(2))(2)PPh(2))(2)M"Cl(2) and CpZr(OH)(OSO(2)(CH(2))(2) PPh(2))(2)M"Cl(2) respectively. The compounds are characterised by infrared and NMR spectroscopies and elemental analysis. Electrospray mass spectra of the complexes are reported and compared to those of Cp(2)M'Cl(2) in water and dimethylsulfoxide (DMSO). For zirconocene dichloride and its product heterobimetallic complexes, the addition of ethylenediamine tetraacetic acid disodium salt (Na(2)H(2)EDTA) was found to be an effective ionisation enhancement agent for the electrospray mass spectral studies. Cytotoxicity studies for the previously reported Cl(2)Pt(PPh(2)(CH(2))(2)SO(3)H)(2).3.5H(2)O (Wedgwood et al., Inorg. Chim. Acta 290 (1999) 189), and the compounds Cp(2)Ti(OSO(2)(CH(2))(2) PPh(2))(2).1.5H(2)O and Cp(2)Ti(OSO(2)(CH(2))(2)PPh(2))(2)PtCl(2).4H(2)O reported here, have been evaluated by colony formation assay against cisplatin-sensitive and -resistant cell lines L929 and L929/R to highlight potential chemotherapeutic activity. The compound Cl(2)Pt(PPh(2)(CH(2))(2)SO(3)H)(2).3.5H(2)O overcomes cisplatin resistance.  相似文献   

9.
Adenosine A(2a) receptor (A(2a)R) colocalizes with dopamine D(2) receptor (D(2)R) in the basal ganglia and modulates D(2)R-mediated dopaminergic activities. A(2a)R and D(2)R couple to stimulatory and inhibitory G proteins, respectively. Their opposing roles in regulating neuronal activities, such as locomotion and alcohol consumption, are mediated by their opposite actions on adenylate cyclase, which often serves as "co-incidence detector" of various activators. On the other hand, the neural actions of A(2a)R and D(2)R are also, at least partially, independent of each other, as indicated by studies using D(2)R and A(2a)R knock-out mice. Here we co-expressed human A(2a)R and human D(2L)R in CHO cells and examined their signaling characteristics. Human A(2a)R desensitized rapidly upon agonist stimulation. A(2a)R activity (80%) was diminished after 2 hr of pretreatment with its agonist CGS21680. In contrast, human D(2L)R activity was sustained even after 2 hr and 18 hr pretreatment with its agonist quinpirole. Long-term (18 hr) stimulation of human D(2L)R also increased basal cAMP levels in CHO cells, whereas long-term (18 hr) activation of human A(2a)R did not affect basal cAMP levels. Furthermore, long-term (18 hr) activation of D(2L)R dramatically sensitized A(2a)R-induced stimulation of adenylate cyclase in a pertussis toxin-sensitive way. Forskolin-induced cAMP accumulation was significantly increased after short-term (2 hr) human D(2L)R stimulation and further elevated after long-term (18 hr) D(2L)R activation. However, neither short-term (2 hr) nor long-term (18 hr) stimulation of A(2a)R affected the inhibitory effects of D(2L)R on adenylate cyclase. Co-stimulation of A(2a)R and D(2L)R could not induce desensitization or sensitization of D(2L)R either. In summary, signaling through A(2a)R and D(2L)R is distinctive and synergistic, supporting their unique and yet integrative roles in regulating neuronal functions when both receptors are present.  相似文献   

10.
The DNA-binding proteins from starved cells (Dps) are a family of proteins induced in microorganisms by oxidative or nutritional stress. Escherichia coli Dps, a structural analog of the 12-subunit Listeria innocua ferritin, binds and protects DNA against oxidative damage mediated by H(2)O(2). Dps is shown to be a Fe-binding and storage protein where Fe(II) oxidation is most effectively accomplished by H(2)O(2) rather than by O(2) as in ferritins. Two Fe(2+) ions bind at each of the 12 putative dinuclear ferroxidase sites (P(Z)) in the protein according to the equation, 2Fe(2+) + P(Z) --> [(Fe(II)(2)-P](FS)(Z+2) + 2H(+). The ferroxidase site (FS) bound iron is then oxidized according to the equation, [(Fe(II)(2)-P](FS)(Z+2) + H(2)O(2) + H(2)O --> [Fe(III)(2)O(2)(OH)-P](FS)(Z-1) + 3H(+), where two Fe(II) are oxidized per H(2)O(2) reduced, thus avoiding hydroxyl radical production through Fenton chemistry. Dps acquires a ferric core of approximately 500 Fe(III) according to the mineralization equation, 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2Fe(III)OOH((core)) + 4H(+), again with a 2 Fe(II)/H(2)O(2) stoichiometry. The protein forms a similar ferric core with O(2) as the oxidant, albeit at a slower rate. In the absence of H(2)O(2) and O(2), Dps forms a ferrous core of approximately 400 Fe(II) by the reaction Fe(2+) + H(2)O + Cl(-) --> Fe(II)OHCl((core)) + H(+). The ferrous core also undergoes oxidation with a stoichiometry of 2 Fe(II)/H(2)O(2). Spin trapping experiments demonstrate that Dps greatly attenuates hydroxyl radical production during Fe(II) oxidation by H(2)O(2). These results and in vitro DNA damage assays indicate that the protective effect of Dps on DNA most likely is exerted through a dual action, the physical association with DNA and the ability to nullify the toxic combination of Fe(II) and H(2)O(2). In the latter process a hydrous ferric oxide mineral core is produced within the protein, thus avoiding oxidative damage mediated by Fenton chemistry.  相似文献   

11.
Binuclear rhodium (II) complexes, [Rh2(OOCPh)2(phen)2(H2O)2] (OOCPh)2 (1), [Rh2(OOCPh)2(bpy)2(H2O)2] (OOCPh)2 (2), [Rh2(OOCBu(n))2 (bpy)2(H2O)2] (OOCBu(n)2 (3), and [Rh2(OOCPr(n)2 (phen)2(H2O)2] (OOCPr(n)2 (4) (Phen = 1,10-phenanthroline and bpy = 2,2'-bipyridine), have been synthesized and characterized using NMR, IR and electronic spectra. Activity of these compounds against Gram-positive bacteria decreases in the order: 1?2?3 > 4. Complex 1 is active against many Staphylococcus strains resistant to commonly used antibiotics. The complexes 1-4 are much less active agents against Gram-negative bacteria.  相似文献   

12.
Given that coactivation of adenosine A(2A) (A(2A)R) and dopamine D(2) (D(2)R) receptors results in the coaggregation, cointernalization, and codesensitization of the A(2A)R and D(2)R and the role of scaffolding protein β-arrestin2 in the desensitization, internalization, and signaling of G-protein-coupled receptors, in this study we explored the ability of the A(2A)R agonist CGS21680 in A(2A)R-D(2)R-coexpressing cells to modulate the D(2)R agonist-induced recruitment of β-arrestin2 to the D(2)R by means of proximity-based bioluminescence resonance energy transfer (BRET(2)) and co-trafficking analysis. We found evidence that CGS21680 can increase the maximal BRET(2) signal between β-arrestin2(RLuc) and D(2L)R(GFP2) upon D(2)R activation, by increasing the potency of the D(2)R agonist to exert this action. In addition, this change was associated with an increased formation of cytoplasmic clusters containing β-arrestin2(GFP2) and D(2L)R(YFP) as seen from the co-trafficking analysis. Furthermore, the A(2A)R agonist advanced the time for the increase in Akt phosphorylation obtained with the D(2)R agonist. Finally, using a novel bioinformatics approach to predict the protein-protein interface, we have also found that amino acid pro-triplets TNY, LLS, RAF, and VSR may be crucial for the -induced β-arrestin2 recruitment by A(2A)R-D(2)R heteromers. Taken together, the results indicate that the antagonistic A(2A)R-D(2)R allosteric receptor-receptor interaction in A(2A)R-D(2)R heteromers favors β-arrestin2 recruitment to the D(2L)R protomer with subsequent cointernalization associated with a reduced time onset of Akt phosphorylation followed by a rapid dephosphorylation. Thus, β-arrestin2 action becomes more rapid and short-lasting and, in this way, mimics G-protein-mediated signaling.  相似文献   

13.
Bacterioferritin (EcBFR) of Escherichia coli is an iron-mineralizing hemoprotein composed of 24 identical subunits, each containing a dinuclear metal-binding site known as the "ferroxidase center." The chemistry of Fe(II) binding and oxidation and Fe(III) hydrolysis using H(2)O(2) as oxidant was studied by electrode oximetry, pH-stat, UV-visible spectrophotometry, and electron paramagnetic resonance spin trapping experiments. Absorption spectroscopy data demonstrate the oxidation of two Fe(II) per H(2)O(2) at the ferroxidase center, thus avoiding hydroxyl radical production via Fenton chemistry. The oxidation reaction with H(2)O(2) corresponds to [Fe(II)(2)-P](Z) + H(2)O(2) --> [Fe(III)(2)O-P](Z) + H(2)O, where [Fe(II)(2)-P](Z) represents a diferrous ferroxidase center complex of the protein P with net charge Z and [Fe(III)(2)O-P](Z) a micro-oxo-bridged diferric ferroxidase complex. The mineralization reaction is given by 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2FeOOH((core)) + 4H(+), where two Fe(II) are again oxidized by one H(2)O(2). Hydrogen peroxide is shown to be an intermediate product of dioxygen reduction when O(2) is used as the oxidant in both the ferroxidation and mineralization reactions. Most of the H(2)O(2) produced from O(2) is rapidly consumed in a subsequent ferroxidase reaction with Fe(II) to produce H(2)O. EPR spin trapping experiments show that the presence of EcBFR greatly attenuates the production of hydroxyl radical during Fe(II) oxidation by H(2)O(2), consistent with the ability of the bacterioferritin to facilitate the pairwise oxidation of Fe(II) by H(2)O(2), thus avoiding odd electron reduction products of oxygen and therefore oxidative damage to the protein and cellular components through oxygen radical chemistry.  相似文献   

14.
Group IVA cytosolic phospholipase A(2) (cPLA(2)alpha) initiates eicosanoid production; however, this pathway is not completely ablated in cPLA(2)alpha(-/-) lung fibroblasts stimulated with A23187 or serum. cPLA(2)alpha(+/+) fibroblasts preferentially released arachidonic acid, but A23187-stimulated cPLA(2)alpha(-/-) fibroblasts nonspecifically released multiple fatty acids. Arachidonic acid release from cPLA(2) alpha(-/-) fibroblasts was inhibited by the cPLA(2)alpha inhibitors pyrrolidine-2 (IC(50), 0.03 microM) and Wyeth-1 (IC(50), 0.1 microM), implicating another C2 domain-containing group IV PLA(2). cPLA(2) alpha(-/-) fibroblasts contain cPLA(2)beta and cPLA(2)zeta but not cPLA(2)epsilon or cPLA(2)delta. Purified cPLA(2)zeta exhibited much higher lysophospholipase and PLA(2) activity than cPLA(2)beta and was potently inhibited by pyrrolidine-2 and Wyeth-1, which did not inhibit cPLA(2)beta. In contrast to cPLA(2)beta, cPLA(2)zeta expressed in Sf9 cells mediated A23187-induced arachidonic acid release, which was inhibited by pyrrolidine-2 and Wyeth-1. cPLA(2)zeta exhibits specific activity, inhibitor sensitivity, and low micromolar calcium dependence similar to cPLA(2)alpha and has been identified as the PLA(2) responsible for calcium-induced fatty acid release and prostaglandin E(2) production from cPLA(2) alpha(-/-) lung fibroblasts. In response to ionomycin, EGFP-cPLA(2)zeta translocated to ruffles and dynamic vesicular structures, whereas EGFP-cPLA(2)alpha translocated to the Golgi and endoplasmic reticulum, suggesting distinct mechanisms of regulation for the two enzymes.  相似文献   

15.
Series of dinuclear species [Ru(2)(L)(2)(LH)(2)-mu-S(2)Cl(2)], [Ru(2)(L)(2)(LH)(4)-mu-Se(2)Cl(2)] (L=L(1)H and L(3)H) and [Ru(2)(L(2))(2)(L(2)H)-mu-Se(2)Cl(2)].2H(2)O, where L(1)H, L(2)H and L(3)H represent for 2-mercapto-5-phenyl-1,3,4-oxadiazole, 2-mercapto-benzimidazole and 2-mercapto-benzothiazole, respectively, have been prepared and characterized by their elemental analyses and spectral (IR, UV-visible, 1H NMR and FAB mass) data. The diamagnetism of these complexes are indicative of an exchange coupled dinuclear ruthenium(III) species containing S(2)(2-) and Se(2)(2-) bridges. The complexes along with free heterocycles (L(1)H-L(3)H) as well as RuCl(3) x 3H(2)0 were tested for their nitrate and nitrite reductase activities. The electrochemical behaviour of the complexes showed irreversible oxidation peaks at +(1.04-1.40) V quite comparable to those reported for sulphido bridged complexes.  相似文献   

16.
Apoptosis was studied under conditions that mimic the steady state of H(2)O(2) in vivo. This is at variance with previous studies involving a bolus addition of H(2)O(2), a procedure that disrupts the cellular homeostasis. The results allowed us to define three phases for H(2)O(2)-induced apoptosis in Jurkat T-cells with reference to cytosolic steady state concentrations of H(2)O(2) [(H(2)O(2))(ss)]: (H(2)O(2))(ss) values below 0.7 microM elicited no effects; (H(2)O(2))(ss) approximately 0.7-3 microM induced apoptosis; and (H(2)O(2))(ss) > 3 microM yielded no additional apoptosis and a gradual shift towards necrosis as the mode of cell death were observed. H(2)O(2)-induced apoptosis was not affected by either BCNU, an inhibitor of glutathione reductase, or diamide, a compound that reacts both with low-molecular weight and protein thiols, or selenols. Glutathione depletion, accomplished by incubating cells either with buthionine sulfoximine or in cystine-free medium, rendered cells more sensitive to H(2)O(2)-induced apoptosis, but did not change the threshold and saturating concentrations of H(2)O(2) that induced apoptosis. Two unrelated metal chelators, desferrioxamine and dipyridyl, strongly protected against H(2)O(2)-induced apoptosis. It may be concluded that, under conditions of H(2)O(2) delivery that mimic in vivo situations, the oxidative event that triggers the induction of apoptosis by H(2)O(2) is a Fenton-type reaction and is independent of the thiol or selenium states of the cell.  相似文献   

17.
Purinergic receptors (P2XRs) activate and desensitize in response to the binding of extracellular nucleotides in a receptor- and ligand-specific manner, but the structural bases of their ligand preferences and channel kinetics have been incompletely characterized. Here we tested the hypothesis that affinity of agonists for binding domain accounts for a ligand-specific desensitization pattern. We generated chimeras using receptors with variable sensitivity to ATP in order: P2X(4)R > P2X(2a)R = P2X(2b)R P2X(7)R. Chimeras having the ectodomain Ile(66)-Tyr(310) sequence of P2X(2)R and Val(61)-Phe(313) sequence of P2X(7)R in the backbone of P2X(4)R were expressed but were non-functioning channels. P2X(2a) + X(4)R and P2X(2b) + X(4)R chimeras having the Val(66)-Tyr(315) ectodomain sequence of P2X(4)R in the backbones of P2X(2a)R and P2X(2b)R were functional and exhibited increased sensitivity to ligands as compared with both parental receptors. These chimeras also desensitized faster than parental receptors and in a ligand-nonspecific manner. However, like parental P2X(2b)R and P2X(2a)R, chimeric P2X(2b) + X(4)R desensitized more rapidly than P2X(2a) + X(4)R, and the rate of desensitization of P2X(2a)+X(4)R increased by substituting its Arg(371)-Pro(376) intracellular C-terminal sequence with the Glu(376)-Gly(381) sequence of P2X(4)R. These results indicate the relevance of interaction between the ectodomain and flanking regions around the transmembrane domains on ligand potency and receptor activation. Furthermore, the ligand potency positively correlates with the rate of receptor desensitization but does not affect the C-terminal-specific pattern of desensitization.  相似文献   

18.
The solution-state conformations of eight proline-containing peptide fragments found in human salivary proline-rich glycoprotein (PRG) were investigated in 2 x distilled water (treated with metal ion chelating resin) using 13C-nuclear magnetic resonance (NMR) and circular dichroism (CD) spectroscopy. The peptide sequences and acronyms were as follows: PRG9-2 = NH2-G(1)-P(2)-CONH2, PRG9-3 = NH2-G(1)P(2)-P(3)-CONH2, PRG9-4 = NH2-G(1)-P(2)-P(3)-P(4)-CONH2, PRG9-5 = NH2-G(1)-P(2)-P(3)-P(4)-H(5)-CONH2, PRG9-6 = NH2-G(1)-P(2)-P(3)-P(4)-H(5)-P(6)-CONH2, PRG9-7 = NH2-G(1)-P(2)-P(3)-P(4)-H(5)-P(6)-G(7)-CONH2, PRG9-8 = NH2-G(1)-P(2)-P(3)-P(4)-H(5)-P(6)-G(7)-K(8)-CONH2 and PRG9-9 = NH2-G(1)-P(2)-P(3)-P(4)-H(5)-P(6)-G(7)-K(8)-P(9)-CONH2. Sequence-specific resonance assignments from the 13C-NMR spectra indicated that the trans proline isomer dominated the conformations of the peptides. CD results clearly showed the presence of the poly-L-proline II helix as the major conformation in PRG9-3----PRG9-5, supplemented by beta- and/or gamma-turns in PRG9-6----PRG9-9. These data suggest that in "metal free" water, native PRG could contain several small poly-L-proline II helices along with beta- and/or gamma-turns. Since proline is the major amino acid present in native PRG, these localized conformations may contribute to PRG's global conformation and act as a primary force in determining its biological activities.  相似文献   

19.
Classical benzodiazepines, for example diazepam, interact with alpha(x)beta(2)gamma(2) GABA(A) receptors, x = 1, 2, 3, 5. Little is known about effects of alpha subunits on the structure of the binding pocket. We studied here the interaction of the covalently reacting diazepam analog 7-Isothiocyanato-5-phenyl-1,3-dihydro-2H-1,4-benzodiazepin-2-one (NCS compound) with alpha(1)H101Cbeta(2)gamma(2) and with receptors containing the homologous mutation, alpha(2)H101Cbeta(2)gamma(2), alpha(3)H126Cbeta(2)gamma(2) and alpha(5)H105Cbeta(2)gamma(2). This comparison was extended to alpha(6)R100Cbeta(2)gamma(2) receptors as this mutation conveys to these receptors high affinity towards classical benzodiazepines. The interaction was studied at the ligand binding level and at the functional level using electrophysiological techniques. Results indicate that the geometry of alpha(6)R100Cbeta(2)gamma(2) enables best interaction with NCS compound, followed by alpha(3)H126Cbeta(2)gamma(2), alpha(1)H101Cbeta(2)gamma(2) and alpha(2)H101Cbeta(2)gamma(2), while alpha(5)H105Cbeta(2)gamma(2) receptors show little interaction. Our results allow conclusions about the relative apposition of alpha(1)H101 and homologous positions in alpha(2), alpha(3), alpha(5) and alpha(6) with the position occupied by -Cl in diazepam. During this study we found evidence for the presence of a novel site for benzodiazepines that prevents modulation of GABA(A) receptors via the classical benzodiazepine site. The novel site potentially contributes to the high degree of safety to some of these drugs. Our results indicate that this site may be located at the alpha/beta subunit interface pseudo-symmetrically to the site for classical benzodiazepines located at the alpha/gamma interface.  相似文献   

20.
Sulfide oxidation at halo-alkaline conditions in a fed-batch bioreactor   总被引:1,自引:0,他引:1  
A biotechnological process is described to remove hydrogen sulfide (H(2)S) from high-pressure natural gas and sour gases produced in the petrochemical industry. The process operates at halo-alkaline conditions and combines an aerobic sulfide-oxidizing reactor with an anaerobic sulfate (SO(4) (2-)) and thiosulfate (S(2)O(3) (2-)) reducing reactor. The feasibility of biological H(2)S oxidation at pH around 10 and total sodium concentration of 2 mol L(-1) was studied in gas-lift bioreactors, using halo-alkaliphilic sulfur-oxidizing bacteria (HA-SOB). Reactor operation at different oxygen to sulfide (O(2):H(2)S) supply ratios resulted in a stable low redox potential that was directly related with the polysulfide (S(x) (2-)) and total sulfide concentration in the bioreactor. Selectivity for SO(4) (2-) formation decreased with increasing S(x) (2-) and total sulfide concentrations. At total sulfide concentrations above 0.25 mmol L(-1), selectivity for SO(4) (2-) formation approached zero and the end products of H(2)S oxidation were elemental sulfur (S(0)) and S(2)O(3) (2-). Maximum selectivity for S(0) formation (83.3+/-0.7%) during stable reactor operation was obtained at a molar O(2):H(2)S supply ratio of 0.65. Under these conditions, intermediary S(x) (2-) plays a major role in the process. Instead of dissolved sulfide (HS(-)), S(x) (2-) seemed to be the most important electron donor for HA-SOB under S(0) producing conditions. In addition, abiotic oxidation of S(x) (2-) was the main cause of undesirable formation of S(2)O(3) (2-). The observed biomass growth yield under SO(4) (2-) producing conditions was 0.86 g N mol(-1) H(2)S. When selectivity for SO(4) (2-) formation was below 5%, almost no biomass growth was observed.  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