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.

Background

Acetate metabolism in skeletal muscle is regulated by acetylCoA synthetase (ACS). The main function of ACS is to provide cells with acetylCoA, a key molecule for numerous metabolic pathways including fatty acid and cholesterol synthesis and the Krebs cycle.

Methods

Hyperpolarized [1-13C]acetate prepared via dissolution dynamic nuclear polarization was injected intravenously at different concentrations into rats. The 13C magnetic resonance signals of [1-13C]acetate and [1-13C]acetylcarnitine were recorded in vivo for 1 min. The kinetic rate constants related to the transformation of acetate into acetylcarnitine were deduced from the 3 s time resolution measurements using two approaches, either mathematical modeling or relative metabolite ratios.

Results

Although separated by two biochemical transformations, a kinetic analysis of the 13C label flow from [1-13C]acetate to [1-13C]acetylcarnitine led to a unique determination of the activity of ACS. The in vivo Michaelis constants for ACS were KM = 0.35 ± 0.13 mM and Vmax = 0.199 ± 0.031 μmol/g/min.

Conclusions

The conversion rates from hyperpolarized acetate into acetylcarnitine were quantified in vivo and, although separated by two enzymatic reactions, these rates uniquely defined the activity of ACS. The conversion rates associated with ACS were obtained using two analytical approaches, both methods yielding similar results.

General significance

This study demonstrates the feasibility of directly measuring ACS activity in vivo and, since the activity of ACS can be affected by various pathological states such as cancer or diabetes, the proposed method could be used to non-invasively probe metabolic signatures of ACS in diseased tissue.  相似文献   

2.
A rapid method for the preparation of [1-14C]acetyl-l-carnitine is described. The method involves exchange of [1-14C]acetic acid into a pool of unlabeled acetyl-l-carnitine using the enzymes acetyl-CoA synthetase and carnitine acetyltransferase. After isotopic equilibrium is attained, radioactive acetylcarnitine is separated from the other reaction components by chromatography on Dowex 1 (Cl?) anion exchange resin. One of the procedures used to verify the product [1-14C]acetyl-l-carnitine can be used to synthesize (3S)-[5-14C]citric acid.  相似文献   

3.
Differential assay for choline acetyltransferase   总被引:7,自引:0,他引:7  
A rapid and sensitive radiochemical assay for choline acetyltransferase (EC 2.3.1.6) is reported. The assay allows for the fact that during incubation of [14C]acetyl-CoA and choline with a cell homogenate, at least one product is formed besides [14C]acetylcholine, which passes an anion exchange column. In contrast to [14C]acetylcholine, this major contaminant ([14C]acetylcarnitine) is not hydrolyzed apparently by Electrophorus acetylcholinesterase. Therefore, two types of assays are performed, the one in the presence of an acetylcholinesterase inhibitor, the other in the presence of acetylcholinesterase from Electrophorus. After passing the reaction mixtures over anion exchange columns, the radioactivities of the effluents are determined. Their difference is proportional to the choline acetyltransferase activity.  相似文献   

4.
Rat kidneys were perfused for 30 min with a Krebs-Henseleit bicarbonate buffer with 5 mM glucose. Albumin proved superior to pluronic polyols as oncotic agent with regard to carnitine reabsorption in the perfused kidney. The reabsorption of 30 μM (−)-[methyl-3H]carnitine was approx. 96% during the first 10 min. At 750 μM the reabsorption decreased to 40%. The tubular reabsorptive maximum (Tmax) was approx. 170 nmol/min per kidney. The fractional reabsorption and clearance of (+)-carnitine, γ-butyrobetaine, and carnitine esters did not deviate significantly from that of (−)-carnitine. (+)-Carnitine was not metabolized by the perfused kidney. In perfusions with (−)-carnitine or (−)-carnitine plus 10 mM α-ketoisocaproate or α-ketoisovalerate increased amounts of acetylcarnitine, isovalerylcarnitine and isobutyrylcarnitine were found. Propionate (5 mM) inhibited acetylcarnitine formation. Isovalerylcarnitine, isobutyrylcarnitine and propionylcarnitine were actively degraded to free (−)-carnitine. In urine, we found a disproportionally high excretion of carnitine or carnitine esters formed in the kidney, compared to the same derivatives when ultrafiltrated. Leakage of metabolites formed in the kidney into preurine may explain this phenomenon.  相似文献   

5.
We developed an isotopic technique to assess mitochondrial acetyl-CoA turnover (≈citric acid flux) in perfused rat hearts. Hearts are perfused with buffer containing tracer [13C2,2H3]acetate, which forms M5 + M4 + M3 acetyl-CoA. The buffer may also contain one or two labeled substrates, which generate M2 acetyl-CoA (e.g. [13C6]glucose or [1,2-13C2]palmitate) or/and M1 acetyl-CoA (e.g. [1-13C]octanoate). The total acetyl-CoA turnover and the contributions of fuels to acetyl-CoA are calculated from the uptake of the acetate tracer and the mass isotopomer distribution of acetyl-CoA. The method was applied to measurements of acetyl-CoA turnover under different conditions (glucose ± palmitate ± insulin ± dichloroacetate). The data revealed (i) substrate cycling between glycogen and glucose-6-P and between glucose-6-P and triose phosphates, (ii) the release of small excess acetyl groups as acetylcarnitine and ketone bodies, and (iii) the channeling of mitochondrial acetyl-CoA from pyruvate dehydrogenase to carnitine acetyltransferase. Because of this channeling, the labeling of acetylcarnitine and ketone bodies released by the heart are not proxies of the labeling of mitochondrial acetyl-CoA.  相似文献   

6.
Anion conductance and permeability sequences were obtained for frog skeletal muscle membranes from the changes in characteristic resistance and transmembrane potential after the replacement of one anion by another in the bathing solution. Permeability and conductance sequences are the same. The conductance sequence at pH = 7.4 is Cl- Br- > NO3 - > I- > trichloroacetate ≥ benzoate > valerate > butyrate > proprionate > formate > acetate ≥ lactate > benzenesulfonate ≥ isethionate > methylsulfonate > glutamate ≥ cysteate. The anions are divided into two classes: (a) Chloride-like anions (Cl- through trichloroacetate) have membrane conductances that decrease as pH decreases. The last six members of the complete sequence are also chloride like. (b) Benzoate-like anions (benzoate through acetate) have conductances that increase as pH decreases. At pH = 6.7 zinc ions block Cl- and benzoate conductances with inhibitory dissociation constants of 0.12 and 0.16 mM, respectively. Chloride-like and benzoate-like anions probably use the same channels. The minimum size of the channel aperture is estimated as 5.5 x 6.5 Å from the dimensions of the largest permeating anions. A simple model of the channel qualitatively explains chloride-like and benzoate-like conductance sequences and their dependence on pH.  相似文献   

7.
Ribulose-1,5-bisphosphate caryboxylase-oxygenase is activated by CO2 and Mg2+ in a process distinct from catalysis. The effect of chloroplast metabolites as they separately influenced either activation or catalysis of tobacco carboxylase was examined. Of the 28 metabolites examined, 13 effected activation of the carboxylase. The strongest positive effectors were NADPH, gluconate-6-P, glycerate-2-P, and glycerate-3-P. Negative effectors included ribose-5-P, fructose-6-P, glucose-6-P, and pyrophosphate. The concentration of CO2 or Mg2+ necessary to produce half-maximal activation is defined as Kact. NADPH and gluconate-6-P decreased the Kact(CO2) from 43 to 7.4 and 3.5 μm, respectively (pH 8.0, 5 mm MgCl2). They also decreased the Kact(M.g2+), but had little affect on the affinity of the enzyme for CO2 during the catalytic process. Increasing Mg2+ concentration decreased the Kact(CO2) and increasing CO2 concentration decreased the Kact-(Mg2+). NADP+ and gluconate-6-P also affected the pH profile of activation, shifting it toward lower pH values. Changes in activation had no effect on the pH profile for catalysis of CO2 fixation. Effectors influenced ribulose-1,5-bisphosphate oxygenase in a manner analogous to the carboxylase. At air levels of O2 and CO2, the ratio of carboxylase to oxygenase activity was not changed by the presence of effectors, including hydroxylamine.  相似文献   

8.
The interior of purified cholinergic Torpedo vesicles is acidic, pHin = 5.2 at external pH = 7.4. The internal pH changes linearily as a function of external pH yielding ΔpH = 2.0 and 2.5 at pHout = 6.3 and 9.1 respectively. The proton translocator carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP) + the ionophore valinomycin dissipate the proton gradient across the vesicular membrane and concurrently induce acetylcholine release from vesicles suspended in K+ buffer. The effect of FCCP + valinomycin is not sensitive to external pH values between 6.3 and 9.1 and is diminished at lower external pH. The possible role of intravesicular pH and of the proton electrochemical gradient in the storage of acetylcholine within cholinergic vesicles is discussed.  相似文献   

9.
4-trans-(N,N-dimethylamino)cinnamaldehyde (DACA) is a chromophoric and fluorogenic substrate of aldehyde dehydrogenase. Fluorescence of DACA is enhanced by binding to aldehyde dehydrogenase in the absence of catalysis both in the presence and absence of the coenzyme analogue 5′AMP. DACA binds to aldehyde dehydrogenase with a dissociation constant of 1–3 μM and stoichiometry of 2 mol mol−1 enzyme. Incorporation of DACA during catalysis was also investigated and found to be 2 mol DACA mol−1 enzyme. Effect of pH on the stoichiometry of DACA incorporation during catalysis has shown that DACA incorporation remained constant at 2 mol DACA mol−1 enzyme, despite a 74-fold velocity enhancement between pH 5.0 and 9.0. Increase of pH increased decomposition of enzyme–acyl intermediate without affecting the rate-limiting step of the reaction. At pH 7.0 the pH stimulated velocity enhancement was 10-fold over that at pH 5.0; further velocity enhancement (11.5-fold that of pH 7.0) was achieved by 150 μM Mg2+ ions. The velocity at pH 7.0 with Mg2+ exceeded that of pH 9.0, and that at maximal pH stimulation at pH 9.5. It was observed that level of intermediate decreased to about 1 mol mol−1 enzyme, indicating that Mg2+ ions increased the rate of decomposition of the enzyme–acyl intermediate and shifted the rate-limiting step of the reaction to another step in the reaction sequence.  相似文献   

10.
Controversy concerning acetylcholine-ATP interaction and the possible role of such binding for acetylcholine storage in synaptic vesicles has been resolved by direct binding measurements using an acetylcholine selective membrane electrode. At pH 7.4, acetylcholine was found to bind ATP?4 with a 1:1 stoichiometry and a thermodynamic formation constant of 175M?1. The interaction of acetylcholine with HATP?3 and MgATP?2 was found to be much weaker with formation constants of approximately 20M?1 and 25M?1, respectively. The data indicate that ATP binding could not account for more than 20% of acetylcholine storage under the conditions known to exist in synaptic vesicles.  相似文献   

11.
The formation of an enzyme-inhibitor adduct between bovine trypsin and [2-13C]p-amidinophenylpyruvic acid has been investigated by 13C NMR spectroscopy. The observation of a resonance at 100.8 ppm demonstrates that the hemiketal formed between the hydroxyl of serine-195 and the 2-13C carbon of p-amidinophenylpyruvic acid is sp3 hybridized with no significant deviation from tetrahedral geometry. It is shown that stabilization of the hemiketal oxyanion if it occurs is less effective than in chloromethylketone inhibitor complexes. The tetrahedral adduct is stable from pH 3 to 8. The mechanisms of breakdown of the tetrahedral adduct at pH extremes are discussed.  相似文献   

12.
We investigated the effect of carnitine analogues on the physiology of Lactobacillus plantarum subjected to salt stress. Salt stressed cells of L. plantarum accumulated exogenously provided carnitine and its structural analogues acetylcarnitine and propionylcarnitine to maximum concentrations of 466, 122 and 75 μmol (g dry weight of cells)−1, respectively. Addition of these carnitines to osmotically stressed medium increased growth rate. Furthermore, the intracellular amino acid pool, consisting of mainly aspartate and glutamate, was reduced when carnitine, acetylcarnitine or propionylcarnitine were included in the medium. This is the first study demonstrating a role for β-substituted acylcarnitine esters in osmoadaptation of a lactic acid bacterium.  相似文献   

13.
The toxicity of organophosphorus nerve agents or pesticides arises from accumulation of acetylcholine and overstimulation of both muscarinic and nicotinic acetylcholine receptors (mAChRs and nAChRs) due to inhibition of acetylcholinesterase (AChE). Standard treatment by administration of atropine and oximes, e.g., obidoxime or pralidoxime, focuses on antagonism of mAChRs and reactivation of AChE, whereas nicotinic malfunction is not directly treated. An alternative approach would be to use nAChR active substances to counteract the effects of accumulated acetylcholine. Promising in vitro and in vivo results were obtained with the bispyridinium compounds SAD-128 (1,1′-oxydimethylene bis(4-tert-butylpyridinium) dichloride) and MB327 (1,1′-(propane-1,3-diyl)bis(4-tert-butylpyridinium) di(iodide)), which were partly attributed to their interaction with nAChRs. In this study, a homologous series of unsubstituted and 4-tert-butyl-substituted bispyridinium compounds with different alkane linker lengths was investigated in competition binding experiments using [3H]epibatidine as a reporter ligand. Additionally, the effect of the well-characterised MB327 on the [3H]epibatidine equilibrium dissociation (KD) constant in different buffers was determined. This study demonstrated that divalent cations increased the affinity of [3H]epibatidine. Since quaternary ammonium molecules are known to inhibit AChE, the obtained affinity constants of the tested bispyridinium compounds were compared with the inhibition of human AChE. In competition experiments, bispyridinium derivatives of longer linker length displaced [3H]epibatidine and inhibited AChE strongly. Bispyridinium compounds with short linkers, at most, have an allosteric interaction with the [3H]epibatidine binding sites and barely inhibited AChE. In dependence on alkane linker length, the bispyridinium compounds seemed to interact at different binding sites. However, the exact binding sites of the bispyridinium compounds responsible for the positive pharmacological effects have still not been identified, making predictive drug design difficult.  相似文献   

14.
Here we show that male, but not female mice lacking expression of the GTPase M-Ras developed urinary retention with distention of the bladder that exacerbated with age but occurred in the absence of obvious anatomical outlet obstruction. There were changes in detrusor morphology in Mras -/- males: Smooth muscle tissue, which exhibited a compact organization in WT mice, appeared disorganized and became increasingly ‘layered’ with age in Mras -/- males, but was not fibrotic. Bladder tissue near the apex of bladders of Mras -/- males exhibited hypercontractility in response to the cholinergic agonist carbachol in in vitro, while responses in Mras -/- females were normal. In addition, spontaneous phasic contractions of detrusors from Mras -/- males were increased, and Mras -/- males exhibited urinary incontinence. We found that expression of the muscarinic M2 and M3 receptors that mediate the cholinergic contractile stimuli of the detrusor muscle was dysregulated in both Mras -/- males and females, although only males exhibited a urinary phenotype. Elevated expression of M2R in young males lacking M-Ras and failure to upregulate M3R with age resulted in significantly lower ratios of M3R/M2R expression that correlated with the bladder abnormalities. Our data suggests that M-Ras and M3R are functionally linked and that M-Ras is an important regulator of male bladder control in mice. Our observations also support the notion that bladder control is sexually dimorphic and is regulated through mechanisms that are largely independent of acetylcholine signaling in female mice.  相似文献   

15.
When the only solute present is a weak acid, HA, which penetrates as molecules only into a living cell according to a curve of the first order and eventually reaches a true equilibrium we may regard the rate of increase of molecules inside as See PDF for Equation where PM is the permeability of the protoplasm to molecules, Mo, denotes the external and Mi the internal concentration of molecules, Ai denotes the internal concentration of the anion A- and See PDF for Equation (It is assumed that the activity coefficients equal 1.) Putting PMFM = VM, the apparent velocity constant of the process, we have See PDF for Equation where e denotes the concentration at equilibrium. Then See PDF for Equation where t is time. The corresponding equation when ions alone enter is See PDF for Equation. where K is the dissociation constant of HA, PA is the permeability of the protoplasm to the ion pair H+ + A-, and Aie denotes the internal concentration of Ai at equilibrium. Putting PAKFM = VA, the apparent velocity constant of the process, we have See PDF for Equation and See PDF for Equation When both ions and molecules of HA enter together we have See PDF for Equation where Si = Mi + Ai and Sie is the value of Si at equilibrium. Then See PDF for Equation VM, VA, and VMA depend on FM and hence on the internal pH value but are independent of the external pH value except as it affects the internal pH value. When the ion pair Na+ + A- penetrates and Nai = BAi, we have See PDF for Equation and See PDF for Equation where P NaA is the permeability of the protoplasm to the ion pair Na+ + A-, Nao and Nai are the external and internal concentrations of Na+, See PDF for Equation, and V Na is the apparent velocity constant of the process. Equations are also given for the penetration of: (1) molecules of HA and the ion pair Na+ + A-, (2) the ion pairs H+ + A- and Na+ + A-, (3) molecules of HA and the ion pairs Na+ + A- and H+ + A-. (4) The penetration of molecules of HA together with those of a weak base ZOH. (5) Exchange of ions of the same sign. When a weak electrolyte HA is the only solute present we cannot decide whether molecules alone or molecules and ions enter by comparing the velocity constants at different pH values, since in both cases they will behave alike, remaining constant if FM is constant and falling off with increase of external pH value if FM falls off. But if a salt (e.g., NaA) is the only substance penetrating the velocity constant will increase with increase of external pH value: if molecules of HA and the ions of a salt NaA. penetrate together the velocity constant may increase or decrease while the internal pH value rises. The initial rate See PDF for Equation (i.e., the rate when Mi = 0 and Ai = 0) falls off with increase of external pH value if HA alone is present and penetrates as molecules or as ions (or in both forms). But if a salt (e.g., NaA) penetrates the initial rate may in some cases decrease and then increase as the external pH value increases. At equilibrium the value of Mi equals that of Mo (no matter whether molecules alone penetrate, or ions alone, or both together). If the total external concentration (So = Mo + Ao) be kept constant a decrease in the external pH value will increase the value of Mo and make a corresponding increase in the rate of entrance and in the value at equilibrium no matter whether molecules alone penetrate, or ions alone, or both together. What is here said of weak acids holds with suitable modifications for weak bases and for amphoteric electrolytes and may also be applied to strong electrolytes.  相似文献   

16.
The binuclear copper in the active site of Carcinus maenas hemocyanin has been substituted with one EDTA-resistant Co(II) per 75 000 Mr by reconstitution of the apo protein. Specific cobalt substitution at the copper binding site is demonstrated from the optical spectral changes directly correlated with the amount of Co(II) bound to the protein, the ellipticity in CD spectra in the near UVVis region, and the efficiency of tryptophan fluorescence quenching. The optical absorption spectrum of the cobalt-substituted protein is characterized by a band pattern attributable to d-d transitions of the metal ion. Both the position of the wavelength maximum (568 nm) and the molar extinction coefficient (≅300 M-1 cm-1) are typical of a four-coordinate, pseudo-tetrahedral Co(II) center.Optical titrations indicate that Cl-, Br-, N3-, SCN-, and CN- bind to Co(II)Hc, each with a stoichiometry of 1:1 per metal center. The apparent stability constants determined from Hill plots of titration data decrease in the order CN- » N3- ≅ SCN- >Cl->Br-. Low temperature EPR studies demonstrate that at pH 7, the cobalt is high spin both in the presence and absence of anionic ligands. A low spin species is formed at pH 9 in the presence of cyanide. The spectrum of this latter complex exhibits superhyperfine structure indicative of metal ligation to 14N supplied by the protein. Direct ligation of cyanide to cobalt is demonstrated by additional spectral splitting observed when this complex is formed using 13C-labelled CN-.  相似文献   

17.
Three covalent anthocyanin–flavonol complexes (pigments 1–3) were extracted from the violet-blue flower of Allium ‘Blue Perfume’ with 5% acetic acid-MeOH solution, in which pigment 1 was the dominant pigment. These three pigments are based on delphinidin 3-glucoside as their deacylanthocyanin and were acylated with malonyl kaempferol 3-sophoroside-7-glucosiduronic acid or malonyl-kaempferol 3-p-coumaroyl-tetraglycoside-7-glucosiduronic acid in addition to acylation with acetic acid.By spectroscopic and chemical methods, the structures of these three pigments 1–3 were determined to be: pigment 1, (6I-O-(delphinidin 3-O-(3I-O-(acetyl)-β-glucopyranosideI)))(2VI-O-(kaempferol 3-O-(2II-O-(3III-O-(β-glucopyranosylV)-β-glucopyranosylIII)-4II-O-(trans-p-coumaroyl)-6II-O-(β-glucopyranosylIV)-β-glucopyranosideII)-7-O-(β-glucosiduronic acidVI))) malonate; pigment 2, (6I-O-(delphinidin 3-O-(3I-O-(acetyl)-β-glucopyranosideI)))(2VI-O-(kaempferol 3-O-(2II-O-β-glucopyranosylIII)-β-glucopyranosideII)-7-O-(β-glucosiduronic acidVI))); and pigment 3, (6I-O-(delphinidin 3-O-(3I-O-(acetyl)-β-glucopyranosideI)))(2VI-O-(kaempferol 3-O-(2II-O-(3III-O-(β-glucopyranosylV)-β-glucopyranosylIII)-4II-O-(cis-p-coumaroyl)-6II-O-(β-glucopyranosylIV)-β-glucopyranosideII)-7-O-(β-glucosiduronic acidVI))) malonate.The structure of pigment 2 was analogous to that of a covalent anthocyanin–flavonol complex isolated from Allium schoenoprasum where delphinidin was observed in place of cyanidin. The three covalent anthocyanin–flavonol complexes (pigment 1–3) had a stable violet-blue color with three characteristic absorption maxima at 540, 547 and 618 nm in pH 5–6 buffer solution. From circular dichroism measurement of pigment 1 in the pH 6.0 buffer solution, cotton effects were observed at 533 (+), 604 (−) and 638 (−) nm. Based on these results, these covalent anthocyanin–flavonol complexes were presumed to maintain a stable intramolecular association between delphinidin and kaempferol units closely related to that observed between anthocyanin and hydroxycinnamic acid residues in polyacylated anthocyanins. Additionally, an acylated kaempferol glycoside (pigment 4) was isolated from the same flower extract, and its structure was determined to be kaempferol 3-O-sophoroside-7-O-(3-O-(malonyl)-β-glucopyranosiduronic acid).  相似文献   

18.
Slc4a4-null mice are a model of proximal renal tubular acidosis (pRTA). Slc4a4 encodes the electrogenic sodium base transporter NBCe1 that is involved in transcellular base transport and pH regulation during amelogenesis. Patients with mutations in the SLC4A4 gene and Slc4a4-null mice present with dysplastic enamel, amongst other pathologies. Loss of NBCe1 function leads to local abnormalities in enamel matrix pH regulation. Loss of NBCe1 function also results in systemic acidemic blood pH. Whether local changes in enamel pH and/or a decrease in systemic pH are the cause of the abnormal enamel phenotype is currently unknown. In the present study we addressed this question by explanting fetal wild-type and Slc4a4-null mandibles into healthy host kidney capsules to study enamel formation in the absence of systemic acidemia. Mandibular E11.5 explants from NBCe1−/− mice, maintained in host kidney capsules for 70 days, resulted in teeth with enamel and dentin with morphological and mineralization properties similar to cultured NBCe1+/+ mandibles grown under identical conditions. Ameloblasts express a number of proteins involved in dynamic changes in H+/base transport during amelogenesis. Despite the capacity of ameloblasts to dynamically modulate the local pH of the enamel matrix, at least in the NBCe1−/− mice, the systemic pH also appears to contribute to the enamel phenotype. Extrapolating these data to humans, our findings suggest that in patients with NBCe1 mutations, correction of the systemic metabolic acidosis at a sufficiently early time point may lead to amelioration of enamel abnormalities.  相似文献   

19.
A stopped flow rapid reaction apparatus capable of following changes of ±0.02 pH unit in 0.1 ml of solution in less than 0.005 sec has been developed, utilizing a commercially available pH-sensitive glass electrode. Using this instrument, extracellular pH at 37°C was followed from less than 0.025 sec to 300 sec after mixing equal volumes of the following CO2-free solutions: (A) normal human red cells, washed three times and resuspended in 150 mM NaCl at pH 7.2 with a hematocrit of 18%; and, (B) 150 mM NaCl adjusted with HCl or NaOH to pH 2.1 to pH 10.3. A minimum of 2 ml of mixture had to flow through the electrode chamber to ensure complete washout. The mixing process produced a step change in the pH of the extracellular fluid, after which exchanges across the red cell membrane and buffering by intracellular hemoglobin caused it to return toward pH 7.2 with an approximately exponential time course. Under the assumption that pH changes after mixing represent exchanges of hydroxyl for chloride ions across the cell membrane, hydroxyl ion permeabilities (P OH - in cm/sec) were calculated and found to vary from 2 x 10-4 at pH 9 to 4 x 10-1 at pH 4 according to the empirical relationship P OH - = 170 exp (-1.51 pH). The form of the dependence of P OH - on extracellular pH does not appear compatible with a simple fixed charge theory of membrane permselectivity.  相似文献   

20.
A fourth molecular from of α-galactosidase, designated LIV, an alkaline α-galactosidase, was isolated from leaves of Cucurbita pepo and purified 165-fold. It was active over a narrow pH range with optimal hydrolysis of p-nitrophenyl-α-d-galactoside and stachyose at pH 7.5. The rate of stachyose hydrolysis was 10 times that of raffinose. Km determinations in McIlvaine buffer (200 millimolar Na2-phosphate, 100 millimolar citric acid, pH 7.5) for p-nitrophenyl-α-d-galactoside, stachyose, and raffinose were 1.40, 4.5, and 36.4 millimolar, respectively. LIV was partially inhibited by Ca2+, Mg2+, and Mn2+, more so by Ni2+, Zn2+, and Co2+, and highly so by Cu2+, Ag2+, Hg2+ and by p-chloromercuribenzoate. It was not inhibited by high concentrations of the substrate p-nitrophenyl-α-d-galactoside or by myo-inositol, but α-d-galactose was a strong inhibitor. As observed for most other forms of α-galactosidase, LIV only catalyzed the hydrolysis of glycosides possessing the α-d-galactose configuration at C1, C2, and C4, and did not hydrolyze p-nitrophenyl-α-d-fucoside (α-d-galactose substituted at C6). The enzyme was highly sensitive to buffers and chelating agents. Maximum hydrolytic activity for p-nitrophenyl-α-d-galactoside was obtained in McIlvaine buffer (pH 7.5). In 10 millimolar triethanolaminehydrochloride-NaOH (pH 7.5) or 10 millimolar Hepes-NaOH (pH 7.5), hydrolytic activity was virtually eliminated, but the addition of low concentrations of either ethylenediaminetetraacetate or citrate to these buffers restored activity almost completely. Partial restoration of activity was also observed, but at higher concentrations, with pyruvate and malate. Similar effects were found for stachyose hydrolysis, but in addition some inhibition of LIV in McIlvaine buffer, possibly due to the high phosphate concentration, was observed with this substrate. It is questionable whether the organic acid anions possess any regulatory control of LIVin vivo. It was possible that the results reflected the ability of these anions, and ethylene-diaminetetraacetate, to restore LIV activity through coordination with some toxic cation introduced as a buffer contaminant.  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