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.

1. 1. Cyanide inhibits the catalytic activity of cytochrome aa3 in both polarographic and spectrophotometric assay systems with an apparent velocity constant of 4·103 M−1·s−1 and a Ki that varies from 0.1 to 1.0 μM at 22 °C, pH 7·3.

2. 2. When cyanide is added to the ascorbate-cytochrome c-cytochromeaa3−O2 system a biphasic reduction of cytochrome c occurs corresponding to an initial Ki of 0.8 μM and a final Ki of about 0.1 μM for the cytochrome aa3−cyanide reaction.

3. 3. The inhibited species (a2+a33+HCN) is formed when a2+a33+ reacts with HCN, when a2+a32+HCN reacts with oxygen, or when a3+a33+HCN (cyano-cytochrome aa3) is reduced. Cyanide dissociates from a2+a33+HCN at a rate of 2·10−3 s−1 at 22 °C, pH 7.3.

4. 4. The results are interpreted in terms of a scheme in which one mole of cyanide binds more tightly and more rapidly to a2+a33+ than to a3+a33+.

Abbreviations: TMPD, N,N,N′,N′-tetramethyl-p-phenylenediamine  相似文献   


2.
The aim of our study was to determine whether a meal modifies the antisecretory response induced by PYY and the structural requirements to elicit antisecretory effects of analogue PYY(22–36) for potential antidiarrhea therapy. The variations in short-circuit current (Isc) due to the modification of ionic transport across the rat intestine were assessed in vitro, using Ussing chambers. In fasted rats, PYY induced a dose- and time-dependent reduction in Isc, with a sensitivity threshold at 5 × 10−11 M (ΔIsc −2 ± 0.5 μA/cm2). The reduction was maximal at 10−7 M (Isc −23 ± 2 μA/cm2), and the concentration producing half-maximal inhibition was 10−9 M. At 10−7 M, reduction of Isc by PYY reached 90% of response to 5 × 10−5 M bumetanide. The PYY effect was partly reversed by 10−5 M forskolin (Isc +13.43 ± 2.91 μA/h·cm2, p < 0.05) or 10−3 M dibutyryl adenosine 3′,5′ cyclic monophosphate (Isc +12 ± 1.69 μA/cm2, p < 0.05). Naloxone and tetrodotoxin did not alter the effect of PYY. In addition, PYY and its analogue P915 reduced net chloride ion secretion to 2.85 and 2.29 μEq/cm2 (p < 0.05), respectively. The antisecretory effect of PYY was accompanied by dose- and time-dependent desensitization when jejunum was prestimulated by a lower dose of peptide. The antisecretory potencies exhibited by PYY analogues required both a C-terminal fragment (22–36) and an aromatic amino acid residue (Trp or Phe) at position 27. At 10−7 M the biological activity of PYY was lower in fed than fasted rats (p < 0.001). Our results confirm the antisecretory effect of PYY, but show that the fed period is accompanied by desensitization, similar to the transient desensitization observed in the fasted period with cumulative doses. This suggests that PYY may act as a physiological mediator that reduces intestinal secretion.  相似文献   

3.
Recently, we demonstrated that angiotensin-(1–7) (Ang-(1–7)) stimulates the Na+-ATPase activity through a losartan-sensitive angiotensin receptor, whereas bradykinin inhibits the enzyme activity through the B2 receptor [Regul. Pept. 91 (2000) 45; Pharmacol. Rev. 32 (1980) 1]. In the present paper, the effect of bradykinin (BK) on Ang-(1–7)-stimulated Na+-ATPase activity was evaluated. Preincubation of Na+-ATPase with 10−9 M Ang-(1–7) increases enzyme activity from 7.9±0.9 to 14.1±1.5 nmol Pi mg−1 min−1, corresponding to an increase of 79% (p<0.05). This effect is reverted by bradykinin in a dose-dependent manner (10−14–10−8 M), reaching maximal inhibitory effect at 10−9 M. Des-Arg9 bradykinin (DABK), an agonist of B1 receptor, at the concentrations of 10−9–10−7 M, does not mimic the BK inhibitory effect, and des-Arg9-[Leu8]-BK (DALBK), a B1 receptor antagonist, at the concentrations of 10−10–10−7 M, does not prevent the inhibitory effect of BK on Ang-(1–7)-stimulated enzyme. On the other hand, HOE 140, an antagonist of B2 receptor, abolishes the inhibitory effect of BK on the Ang-(1–7)-stimulated enzyme in a dose-dependent manner, reaching maximal effect at 10−7 M. Taken together, these data indicate that stimulation of B2 receptors by BK can counteract the stimulatory effect of Ang-(1–7) on the proximal tubule Na+-ATPase activity.  相似文献   

4.
This article reports the electrical responses of a phosphate ionophore, the cyclic polyamine 3-decyl-1,5,8-triazacyclodecane-2,4-dione (N3-cyclic amine) incorporated into metal supported bilayer lipid membranes (s-BLM). Teflon coated silver wire was used as a support. In a potentiometric mode, the ionophore had a response that was linearly related to the logarithm of HPO42− concentration and was also dependant on pH. Selectivity coefficients for other anions compared to HPO42− ions, determined by the separate solution method, fell within the range 1.73 × 10−4 to 6.38 × 10−2.  相似文献   

5.
The rat pineal gland is known to release melatonin in response to noradrenergic stimulation. Since vasopressin (VP)- and oxytocin (OT)-containing fibers innervate the pineal gland, the effects of VP and OT on melatonin release from perifused rat pineal glands were investigated. VP (10−7 M) and OT (10−6 M) decreased the basal melatonin secretion. No dose-dependent effect was observed. At high concentrations (10−5) these peptides potentiated the isoproterenol-induced increase of melatonin secretion. Below 10−5 M no potentiation was observed. Fragments of VP {[pGlu4,Cys6]VP(4–9)} and OT {[pGlu4,Cys6]OT(4–9)} did not display any effect on the isoproterenol-induced melatonin secretion.  相似文献   

6.
Differential UV spectroscopy and thermal denaturation were used to study the Mg2+ ion effect on the conformational equilibrium in poly A · 2 poly U (A2U) and poly A · poly U (AU) solutions at low (0.01 M Na+) and high (0.1 M Na+) ionic strengths. Four complete phase diagrams were obtained for Mg2+–polynucleotide complexes in ranges of temperatures 20–96 °C and concentrations (10−5–10−2) M Mg2+. Three of them have a ‘critical’ point at which the type of the conformational transition changes. The value of the ‘critical’ concentration ([Mgt2+]cr=(4.5±1.0)×10−5 M) is nearly independent of the initial conformation of polynucleotides (AU, A2U) and of Na+ contents in the solution. Such a value is observed for Ni2+ ions too. The phase diagram of the (A2U+Mg2+) complex with 0.01 M Na+ has no ‘critical’ point: temperatures of (3→2) and (2→1) transitions increase in the whole Mg2+ range. In (AU+Mg2+) phase diagram at 0.01 M Na+ the temperature interval in which triple helices are formed and destroyed is several times larger than at 0.1 M Na+. Using the ligand theory, a qualitative thermodynamic analysis of the phase diagrams was performed.  相似文献   

7.
Cuaq+ forms stable complexes with carbon monoxide in aqueous solutions. Furthermore it reacts very fast with aliphatic radicals. The reaction of Cu(CO)maq+ with methyl radicals, CH3 was studied using the pulse-radiolysis technique. The results point out that methyl radicals react with Cu(CO)aq+ to form an unstable intermediate with a CuII-C σ bond identified as (CO)CuII-CH3+, k = (1.1±0.2) × 109 M−1 s−1. This intermediate has a strong LMCT charge transfer band (λmax = 385 nm, max = 2500 M−1 cm−1) which is similar to the absorption bands of other transient complexes with CuII-alkyl σ bonds. The coordinated carbon monoxide in (CO)CuII-CH3+ inserts into the copper—carbon bond (or rather the coordinated methyl migrates to the coordinated carbon monoxide ligand) at a rate of (3.0±0.8) × 102 s−1 to form the copperacetyl complex (CO)mCuII-C(CH3)=O+max = 480 nm, max = 2100 M−1 cm−1). The rate of formation of (CO)CuII-CH3+ and of the insertion reaction are pH independent. The complex (CO)mCuII-C(CH3)=O+ is also unstable and decomposes heterolytically to yield acetaldehyde and Cuaq2+ as the final stable products. This reaction is slightly pH dependent. The same reactivity pattern has been observed for the Cu(COnaq+ complexes (n = 2 or 3). The results clearly point out that CO remains coordinated to transient complexes of the type CuII-alkyl.  相似文献   

8.
Aqueous solutions of fractions of an extracellular linear mannan formed by Rhodotorula rubra yeast have been investigated by hydrodynamic methods (high-speed sedimentation, translation isothermic diffusion and viscometry). The molecular weight was determined according to Svedberg ( ) and the polydispersity parameters of the initial sample were also determined (Mw/Mn = 1·20 and Mz/Mw = 1·21). Relationships between the molecular weight (M) and so, Do and [η] in the range were: [η] = 2·33 × 10−2 M0.75, Do = 1·65 × 10−4 M0·58, so = 2·24 × 10−15 M0·43. The equilibrium rigidity and hydrodynamic diameter of chains representing mannan molecules were evaluated.  相似文献   

9.
A cellular suspension from rat submandibular glands was prepared with collagenase. The intracellular pH (pHi) was estimated with 2′,7′-bis-(2-carboxy-ethyl)-5(6)-carboxyfluorescein (BCECF). After exposure to NH4Cl, the pHi transiently increased (diffusion of NH3) and then dropped (influx of NH4+). Isoproterenol increased 2.5-fold the rate of NH4+ influx; bumetanide, an inhibitor of the Na+-K+-2Cl-cotransporter blocked the response to isoproterenol, confirming that the beta-adrenergic agonist stimulated the cotransporter. Forskolin (1 μmol/L) mimicked the response to isoproterenol. VIP (1 nmol/L-1 μmol/L) also increased the activity of the cotransporter. Cyclic AMP rather than calcium was the mediator of this activation since 1) carbachol which increased the [Ca2+]i fivefold increased the uptake of NH4+ by only 50%; 2) only high concentrations of VIP significantly increased the [Ca2+]i; 3) incubation in the presence of EGTA had no effect on the response to VIP; 4) low concentrations (nmol/L) of the neuropeptide increased the intracellular level of cAMP; and 5) the stimulation of the cotransporter by VIP, forskolin, and isoproterenol was inhibited by H8, an inhibitor of cAMP-dependent protein kinase. It is concluded that the Na+-K+-2Cl-cotransporter of rat submandibular glands is activated by isoproterenol, forskolin, and neuropeptides of the VIP family by a mechanism involving cAMP-dependent processes. The activation of the cotransporter by VIP could partly explain the potentiating effect of VIP on the response to sialagogues like substance P or muscarinic agonists.1  相似文献   

10.
Rates of stepwise anation of cis-Cr(ox)2(H2O2) with SCN/N3, Cr(acac)2(H2O)2+ with SCN and Cr(atda)(H2O)2 with SCN have been investigated in weakly acidic aqueous solutions. Rate constants, kI and kII for the two steps in each system, are composite as kx = kx0+kxX[X] (x = I, II; X = SCN, N3). These rate constants have been evaluated also as the corresponding ΔH and ΔS values. The results obtained and the plausible Id mechanism seem to suggest Cr---OOC bond dissociation (hence a strongly negative ΔS) generating the transition state in each system with outer-sphere association forming the precursor complex in the X dependent paths.  相似文献   

11.
Both prostaglandins (PGs) and nitric oxide (NO) have cytoprotective and hyperemic effects in the stomach. However, the effect of NO on PG synthesis in gastric mucosal cells is unclear. We examined whether sodium nitroprusside (SNP), a releaser of NO, stimulates PG synthesis in cultured rabbit gastric mucus-producing cells. These cells did not release NO themselves. Co-incubation with SNP (2 × 10−4, 5 × 10−4, 10−3 M) increased PGE2 synthesis, and SNP (10−3 M) increased PGI2 synthesis in these cells. Hemoglobin, a scavenger of NO, (10−5 M) eliminated the increase in PGE2 synthesis by SNP, but methylene blue, an inhibitor of soluble guanylate cyclase, (5 × 10−5 M) did not affect the increase in PGE2 synthesis by SNP. 8-bromo guanosine 3′ : 5′-cyclic monophosphate (8-bromo cGMP), a cGMP analogue, (10−6, 10−5, 10−4, 10−3 M) did not affect PGE2 synthesis. These findings suggest that NO increased PGE2 and PGI2 synthesis via a cGMP-independent pathway in cultured rabbit gastric cells.  相似文献   

12.
In the present paper, the modulation of the basolateral membrane (BLM) Na+-ATPase activity of inner cortex from pig kidney by angiotensin II (Ang II) and angiotensin-(1–7) (Ang-(1–7)) was evaluated. Ang II and Ang-(1–7) inhibit the Na+-ATPase activity in a dose-dependent manner (from 10−11 to 10−5 M), with maximal effect obtained at 10−7 M for both peptides. Pharmacological evidences demonstrate that the inhibitory effects of Ang II and Ang-(1–7) are mediated by AT2 receptor: The effect of both polypeptides is completely reversed by 10−8 M PD 123319, a selective AT2 receptor antagonist, but is not affected by either (10−12–10−5 M) losartan or (10−10–10−7 M) A779, selective antagonists for AT1 and AT(1–7) receptors, respectively. The following results suggest that a PTX-insensitive, cholera toxin (CTX)-sensitive G protein/adenosine 3′,5′-cyclic monophosphate (cAMP)/PKA pathway is involved in this process: (1) the inhibitory effect of both peptides is completely reversed by 10−9 M guanosine 5′-O-(2-thiodiphosphate) (GDPβS; an inhibitor of the G protein activity), and mimicked by 10−10 M guanosine 5′-O-(3-thiotriphosphate) (GTPγS; an activator of the G protein activity); (2) the effects of both peptides are mimicked by CTX but are not affected by PTX; (3) Western blot analysis reveals the presence of the Gs protein in the isolated basolateral membrane fraction; (4) (10−10–10−6 M) cAMP has a similar and non-additive effect to Ang II and Ang-(1–7); (5) PKA inhibitory peptide abolishes the effects of Ang II and Ang-(1–7); and (6) both angiotensins stimulate PKA activity.  相似文献   

13.
A possible role for endothelin in the pathogenesis of hypoglycemic brain damage in rats was evaluated using an in vitro model with which we could directly monitor the release of dopamine from striatal slices. There was no evidence of impairment in case of non-exposure of the slices to endothelin-3 during 20–40 min of hypoglycemia. The response all but disappeared in striatal slices stimulated twice with 10−5 M endothelin-3 during 20 min of hypoglycemia. Hypoglycemic damage triggered by endothelin-3 was not observed in the absence of extracellular Ca2+. Nifedipine (10−6 M), but not verapamil (10−5 M) nor diltiazem (10−5M), protected striatal tissue from this damage. Our findings provide evidence that endothelins might be etiological factors in the development of hypoglycemic/ischemic brain injury by stimulating dihydropyridine-regulated Ca2+ influx.  相似文献   

14.
Bombesin belongs to a family of peptides acting as local hormones with roles in growth regulation, neural function and secretion. Upon binding to its receptor bombesin primarily elicits an increase of inositolphosphates and diacylglycerol, events leading to increased [Ca2+]i and activation of protein kinase C. When asynchronously growing V79 Chinese hamster cells were treated with bombesin in the 10−9−10−7 M concentration range their content of inositolphosphates increased and so did the frequency of mitotic cells with abnormal chromosomal arrangements (c-mitoses). Both effects were abolished by simultaneous addition of the synthetic peptide antagonist D-Arg1,D-Phe5,D-Trpu9-Leu11-substance P that binds to certain bombesin receptors. These results demonstrate that the V79 cells most probably have receptors for bombesin and that the weak but significant c-mitotic effect is mediated by such receptors.  相似文献   

15.
Reactions of cis-diamminedichloroplatinum(II) with phosphonoformic acid (PFA), phosphonoacetic acid (PAA), and methylenediphosphonic acid (MDP) yield various phosphonatoplatinum(II) chelates which were characterized by phosphorus-31 NMR spectroscopy. The P-31 resonances for the chelates appear at 6–12 ppm downfield as compared to the uncomplexed ligands. All complexes exhibit monoprotic acidic behavior in the pH range 2–10. The chemical shift-pH profiles yielded acidity constants, 1.0 × 10−4, 1.5 × 10−4, and 1.3 × 10−6 M−1, for the PFA, PAA, and MDP chelates. In addition to the monomeric chelate, MDP formed a bridged diplatinum(II,II) complex when it reacted with cis-Pt (NH3)2(H2O)22+. The P-31 resonance for this binuclear complex appears at 22 ppm downfield from the unreacted ligand.

Rate data for the complexation reactions of the phosphonate ligands with the dichloroplatinum complex are consistent with a mechanism in which a monodentate complex is formed initially through rate-limiting aquation process of the platinum complex, followed by a rapid chelation. For the PFA and PAA complexes, initial binding sites are the carboxylato oxygens. Implications of the various binding modes of the phosphonates in relationship to their antiviral activities are discussed.  相似文献   


16.
The structure of the 7S globulin from Phaseoulus vulgaris L in dilatue solutions has been studied by small angle X-ray scattering (SAXS), by quasi-elastic light scattering (Q ELS), by circular dichroism spectroscopy (c.d.), and by precise density measurements. The molar mass, the radius of gyration, the volume, the maximum dimension and the diffusion coefficient were determined as M = 1.45 × 105 g mol−1, RG = 4.05 nm, V = 300- nm3, L = 13.0 nm and D20,w0 = 4.5 × 10−7 cm2 s−1, respectively. The molecule has an asymmetrical shape with the dimensions 12.5 × 12.5 × 3.75 nm. The secondary structure of the 7S globulin is characterized by a small portion of -helical structure (14%) and a marked content of β-structure (18%).  相似文献   

17.
Data are reported for the binding of Ni2+, Co2+, and Mg2+ to the B-form of double-stranded poly(dG-dC) at ionic strength conditions I = 0.001 M, 0.01 M, and 0.1 M. The apparent binding constants for Ni2+ and Co2+ are about the same and are 2- to 3-fold higher than those for Mg2+. Kinetic studies indicate that Mg2+ binds to the polynucleotide mainly (or solely) as a mobile cloud (electrostatically, outer-sphere), whereas the transition metal ions undergo site binding (inner-sphere coordination) with poly(dG-dC). The kinetic data suggest that an Ni2+ ion coordinates to more than one binding site at the polynucleotide, presumably to G-N7 and a phosphate group.

At low ionic strength conditions the addition of Ni2+ induces a B → Z conformational transition in poly(dG-dC). As demonstrated by UV absorption and CD spectroscopy, the transition occurs at I = 0.001 M already when 3 × 10−5 – 7 × 10−5 M of Ni2+ are added to 8 × 10−5 M (in monomeric units) of poly(dG-dC), and at I = 0.01 M between 2.5 × 10−4 and 4.5 × 10−4 M of Ni2+. Using murexide as an indicator of the concentration of free Ni2+ ions, the amount of Ni2+ which is bound to the polynucleotide could be determined. At I = 0.001 M it was established that the B → Z transition begins when 1 Ni2+ is bound coordinatively per four base pairs, and the transition is complete when 1 Ni2+ is bound coordinatively per three base pairs. It is this coordinated Ni2+ which induces the B → Z transition.  相似文献   


18.
Steroid 21-hydroxylase activity has been identified in many tissues, including liver. But it is possible that the enzyme found in the liver is different from adrenal 21-hydroxylase. In the adrenal cortex, steroid 21-hydroxylase activity is increased by corticotropin (ACTH); the effect of ACTH is mediated by cyclic AMP (cAMP), and presumably involves a cAMP-dependent protein kinase (PKA). It is not yet clear, however, how extra-adrenal steroid 21-hydroxylase activity is regulated. In the present study, we examined the effect of N6, 2′-O-dibutyryl adenosine 3′,5′-cyclic monophosphate (dbcAMP), forskolin, N-[2-(methylamino)ethyl]5-isoquinolinesulfonamide (H-8) and 12-O-tetradecanoylphorbol-13-acetate (TPA) on steroid 21-hydroxylase activity in primary cultures of rat hepatocytes to determine the nature of regulation of extra-adrenal steroid 21-hydroxylase activity. Steroid 21-hydroxylase activity in hepatocytes incubated with 10−11M dbcAMP for 24 h was 1.6 times higher than that in control hepatocytes untreated with dbcAMP. On the other hand, steroid 21-hydroxylase activity decreased by 20 and 50% when the cells were incubated with 10−5 and 10−3 M dbcAMP, respectively. The stimulatory effect of 10−11 M dbcAMP was not blocked by 10−5 M H-8 (PKA inhibitor), but the inhibitory effect of 10−5 or 10−3 M cAMP was. TPA did not alter the activity of steroid 21-hydroxylase. These findings indicate that the steroid 21-hydroxylase in rat liver is regulated by mechanisms different from those in the adrenal glands.  相似文献   

19.
Gary Bailin   《BBA》1977,462(3):689-699
A human skeletal actin · tropomyosin · troponin complex was phosphorylated in the presence of [γ-32P]ATP, Mg2+, adenosine 3′:5′-monophosphate (cyclic AMP) and cyclic AMP-dependent protein kinase (protein kinase). Phosphorylation was not observed when the actin complex was incubated in the absence of protein kinase or 1 μM cyclic AMP. In the presence of 10−7 M Ca2+ and protein kinase 0.1 mole of [32P]phosphate per 196 000 g of protein was incorporated. This was two-fold higher than the [32P]phosphate content of a rabbit skeletal actin complex but two-fold lower than that of a bovine cardiac actin complex. At high Ca2+, 5 · 10−5 M, little change in the phosphorylation of a human skeletal actin complex occurred. Phosphoserine and phosphothreonine were identified in the [32P]phosphorylated actin complex. Polyacrylamide gel electrophoresis in sodium dodecyl sulfate showed that 60% of the label was associated with the tropomyosin binding component of troponin. The inhibitory component of troponin contained 16% of the bound [32P]phosphate. Increasing the Ca2+ concentration did not significantly decrease the [32P]phosphate content of the phosphorylated proteins in the actin complex. No change in the distribution of phosphoserine or phosphothreonine was observed. Half maximal calcium activation of the ATPase activity of reconstituted human skeletal actomyosin made with the [32P]phosphorylated human skeletal actin complex was the same as a reconstituted actomyosin made with an actin complex incubated in the absence of protein kinase at low or high Ca2+.  相似文献   

20.
Spontaneous and evoked secretion of acetylcholinesterase from the hippocampus in vivo has been demonstrated by the use of push-pull cannulae. Local perfusion with 10−5M carbachol evoked an increase of 104% in acetylcholinesterase release with no accompanying change in butyrylcholinesterase or lactic dehydrogenase. Local or systemic atropine sulphate blocked the carbachol-evoked increase in acetylcholinesterase release, whilst gallamine had no effect. Local perfusion of γ-aminobutyric acid (10−4M) also blocked the carbachol-evoked release of acetylcholinesterase but had no effect on the spontaneous release.

It is concluded that a soluble form of acetylcholinesterase is secreted from the hippocampus in response to stimulation of muscarinic receptors; this secretion can be influenced by γ-aminobutyric acid, which is present in interneurones in the hippocampus.  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