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 31 毫秒
1.
The protolytic equilibria of 1,2,7,8-dibenzofluorescein in aqueous solution have been characterized by visible absorption and fluorescence spectra. The species involved are identified as dianion, monoanion, neutral form and cation. The neutral form includes both the quinoid and lactone structures. The pK(a)s were calculated by an improved procedure to be 3.14, 4.04 and 6.28, respectively. The absorption spectra for each protolytic form were resolved. The absorption maxima (molar absorption coefficient, x10(5), M(-1) cm(-1)) are 532 nm (0.87) for the dianion, 510 nm (0.39) for the monoanion, 500 nm (0.16) for the neutral form, and 494 nm (0.19) for the cation, respectively. Contrary to the assumption in the literature, we found that the monoanion is highly fluorescent (Phi(f) = 0.66, compared to Phi(f) = 0.25 for dianion) and its molar ratio can reach 50% at neutral pH. It is therefore concluded that under physiological pH conditions the monoanion plays a major role when it is used as a fluorescence probe.  相似文献   

2.
1. The variations with pH (from 36n-sulphuric acid to 10n-sodium hydroxide) of the excitation and fluorescence wavelengths and fluorescence intensity of 2-, 3- and 4-hydroxypyridine and their O- and N-methyl derivatives were investigated. 2. 4-Hydroxy- and 4-methoxy-pyridine were non-fluorescent at all pH values. 3. The cations and dipolar ions of the 3-hydroxypyridine derivatives and the anion of 3-hydroxypyridine were fluorescent, but the neutral forms were not. 4. All the forms of the 2-hydroxypyridine derivatives were fluorescent. 5. Pyridoxol, pyridoxal and its 5-phosphate, pyridoxamine and pyridoxic acid and its lactone were studied similarly. All these compounds, except pyridoxal 5-phosphate, were more fluorescent than 3-hydroxypyridine. 6. The most fluorescent forms of these compounds are the anions, except for pyridoxol, where the dipolar ion was the most fluorescent form. The least fluorescent forms are the neutral molecules. The dipolar ions were appreciably fluorescent in all cases. 7. The most fluorescent form examined was the dianion of pyridoxic acid lactone. 8. The cations were all fluorescent except the cations of 2- and 3-methoxypyridine. All the cations showed excited-state ionization. The excited pK(a) values of these cations were determined and the results are discussed with reference to Weller's (1952) equation relating ground- and excited-state dissociation constants. 9. The pK(a) values for all ionizations undergone by the compounds examined were determined from fluorescence data. 10. Stokes shifts for the various ionic and neutral species of the compounds examined were calculated and are discussed.  相似文献   

3.
Claviceps purpurea PRL 1980 grew on partially dissociated succinic acid (pH 4) but not on fully dissociated succinic acid (pH 7.2). Myeclium suspended in 42 mM solution of partially ionized succinic acid (pH 4; 60.1% nonionized, 39% monoanion, and 0.9% dianion, K+ salt) over a period of 25 min accumulated more succinic acid carbon than mycelium suspended in highly ionized solution (pH 6.8; 0.01% nonionized, 4.8% monoanion, and 95% dianion). The greater accumulation from partially ionized solution was not attributable solely to metabolism of succinic acid nor to the lower external concentration of potassium ion. Rate of uptake by sodium azide and iodoacetate-treated mycelium was proportional to external concentration at least up to 200 mumol/ml. External potassium or sodium ion was not required for uptake by inhibited or uninhibited mycelium and external sodium ion and glucose did not allow concentration of succinic acid. The internal concentrations of succinic acid carbon expressed as succinic acid in cell water were about the same as the external concentrations. Uptake was not appreciably affected by extent of ionization of external succinic acid but accumulation was markedly affected. A plot of accumulated succinic acid carbon against external pH produced a bimodal curve with the two maxima corresponding to the maximal concentrations of nonionized and monoanion succinic acid. The bimodal curve probably results from overlapping of two separate curves; the nonionized form accumulating efficiently because of one interaction with the cell and the monoanion form accumulating efficiently because of another interaction. Uptake from concentrated solution is by diffusion and efflux is rapid but not complete. Efflux is not retarded by presence of phosphate in the external solution.  相似文献   

4.
The UV/Vis absorption and fluorescence properties of dibenzofluorescein (DBFL) in organic solvents were measured and used to shed light on the possible presence of its tautomers or various prototropic forms. DBFL in aprotic solvents mainly exists in two tautomeric forms, viz. quinoid and lactone, but neither are efficiently fluorescent. In protic solvents, such as methanol and ethanol, both the monoanion and neutral quinoid are present and showed the highest fluorescence quantum yield. In contrast, DBFL is fully dissociated to the monoanion and dianion in deionized water.  相似文献   

5.
Cells of Pseudomonas sp. strain HBP1 grown on 2-hydroxy- or 2,2'-dihydroxybiphenyl contain NADH-dependent monooxygenase activity that hydroxylates 2,2'-dihydroxybiphenyl. The product of this reaction was identified as 2,2',3-trihydroxybiphenyl by 1H nuclear magnetic resonance and mass spectrometry. Furthermore, the monooxygenase activity also hydroxylates 2,2',3-trihydroxybiphenyl at the C-3' position, yielding 2,2',3,3'-tetrahydroxybiphenyl as a product. An estradiol ring cleavage dioxygenase activity that acts on both 2,2',3-tri- and 2,2',3,3'-tetrahydroxybiphenyl was partially purified. Both substrates yielded yellow meta-cleavage compounds that were identified as 2-hydroxy-6-(2-hydroxyphenyl)-6-oxo-2,4-hexadienoic acid and 2-hydroxy-6-(2,3-dihydroxyphenyl)-6-oxo-2,4-hexadienoic acid, respectively, by gas chromatography-mass spectrometry analysis of their respective trimethylsilyl derivatives. The meta-cleavage products were not stable in aqueous incubation mixtures but gave rise to their cyclization products, 3-(chroman-4-on-2-yl)pyruvate and 3-(8-hydroxychroman-4-on-2-yl)pyruvate, respectively. In contrast to the meta-cleavage compounds, which were turned over to salicylic acid and 2,3-dihydroxybenzoic acid, the cyclization products are not substrates to the meta-cleavage product hydrolase activity. NADH-dependent salicylate monooxygenase activity catalyzed the conversions of salicylic acid and 2,3-dihydroxybenzoic acid to catechol and pyrogallol, respectively. The partially purified estradiol ring cleavage dioxygenase activity that acted on the hydroxybiphenyls also produced 2-hydroxymuconic semialdehyde and 2-hydroxymuconic acid from catechol and pyrogallol, respectively.  相似文献   

6.
1. Whereas the second-order rate constants for the reaction of the thiolate ion of 2-mercaptoethanol with 4,4'-dipyridyl disulphide (k4PDS) and with 5,5'-dithiobis-2-nitrobenzoate dianion increase with decreasing dielectric constant of the solvent, or remain unchanged, the rate constant for the analogous reaction with 2,2'-dipyridyl disulphide (k2PDS) decreases. This anomalous solvent effect and other unusual physicochemical properties of 2,2'-dipyridyl disulphide are discussed. 2. The differential effect of solvent on the reactions of thiolate ion with the 2,2'- and 4,4'-dipyridyl disulphides is shown to provide a method of characterizing solvent environments of thiol groups in proteins by a reactivity-probe method that should not suffer from the usual drawback associated with the existence of steric or binding effects of unknown magnitude. Application of the method to ficin (EC 3.4.22.3) suggests that its active-centre thiol group resides in a relatively hydrophobic environment. 3. The pH-k profile for the reaction of ficin with 4,4'-dipyridyl disulphide is reported.  相似文献   

7.
The stilbenedisulfonate inhibitory site of the human erythrocyte anion-exchange system has been characterized by using serveral fluorescent stilbenedisulfonates. The covalent inhibitor 4-benzamido-4'-isothiocyanostilbene-2,2'-disulfonate (BIDS) reacts specifically with the band 3 protein of the plasma membrane when added to intact erythrocytes, and the reversible inhibitors 4,4'-dibenzamidostilbene-2,2'-disulfonate (DBDS) and 4-benzamido-4'-aminostilbene-2,2'-disulfonate (BADS) show a fluorescence enhancement upon binding to the inhibitory site on erythrocyte ghosts. The fluorescence properties of all three bound probes indicate a rigid, hydrophobic site with nearby tryptophan residues. The Triton X-100 solublized and purified band 3 protein has similar affinities for DBDS, BADS, and 4,4'-dinitrostilbene-2,2'-disulfonate (DNDS) to those observed on intact erythrocytes and erythrocyte ghosts, showing that the anion binding site is not perturbed by the solubilization procedure. The distance between the stilbenedisulfonate binding site and a group of cysteine residues on the 40 000-dalton amino-terminal cytoplasmic domain of band 3 was measured by the fluorescence resonance energy transfer technique. Four different fluorescent sulfhydryl reagents were used as either energy transfer donors or energy transfer acceptors in combination with the stilbenedisulfonates (BIDS, DBDS, BADS, and DNDS). Efficiencies of transfer were measured by sensitized emisssion, donor quenching, and donor lifetime changes. Although these sites are approachable from opposite sides of the membrane by impermeant reagents, they are separated by only 34--42 A, indicating that the anion binding site is located in a protein cleft which extends some distance into the membrane.  相似文献   

8.
Biphenyl metabolism in Aspergillus toxicarius occurs by successive hydroxylations in the 4- and 4'-positions, followed by conjugation with sulfate to produce 4-hydroxybiphenyl-O-sulfonic acid and 4,4'-dihydroxybiphenyl-O-sulfonic acid. The hydroxylation reactions normally occur only after a prolonged lag period after which the appearance of the monohydroxylated compound precedes the dihydroxylated compound. The accumulation of the monohydroxy compound is transient; therefore, it is an intermediate in the hydroxylating pathway. The onset of hydroxylating activity can be greatly accelerated when the culture is primed with the intermediate or product of the reaction (4-hydroxybiphenyl or 4,4'-dihydroxybiphenyl) at the time of biphenyl addition; a concentration of 0.05 mg 4,4'-dihydroxybiphenyl per ml produces optimal induction. Water-soluble conjugates of 4-hydroxybiphenyl and 4,4'-dihydroxybiphenyl were found in cultures of A. toxicarius grown in the presence of biphenyl plus inducer. The conjugate was shown to be the sulfate ester; no glucuronide or other conjugate species was found in any phase of the transformation. As with hydroxylating activity, the sulfotransferase activity appeared to be induced by the products of biphenyl metabolism.  相似文献   

9.
1. The reactions of 2,2'- and 4,4'-dipyridyl disulphide (2-Py-S-S-2-Py and 4-Py-S-S-4-Py) with l-ergothioneine (2-mercapto-l-histidine betaine), 2-mercaptoimidazole, 1-methyl-2-mercaptoimidazole, thiourea, thioacetamide, 2-thiopyridone (Py-2-SH) and 4-thiopyridone (Py-4-SH) were investigated spectrophotometrically in the pH range approx. 1-9. 2. These reactions involve two sequential reversible thiol-disulphide interchanges. 3. The reaction of l-ergothioneine with 2-Py-S-S-2-Py and/or with the l-ergothioneine-Py-2-SH mixed disulphide, both of which provide Py-2-SH, is characterized by at least three reactive protonic states. This provides definitive evidence that neutral l-ergothioneine is a reactive nucleophile, particularly towards the highly electrophilic protonated disulphides. 4. A similar situation appears to obtain in the reactions of l-ergothioneine and Py-2-SH with 4-Py-S-S-4-Py and in the reactions of the other 2-mercaptoimidazoles, thiourea and Py-4-SH with 2-Py-S-S-2-Py. The nucleophilic reactivity of Py-4-SH suggests that general base catalysis provided by the disulphide in a cyclic or quasi-cyclic transition state is not necessary to generate nucleophilic reactivity in the other amino-thiones whose geometry could permit such catalysis. 5. The existence of a positive deuterium isotope effect in the l-ergothioneine-2-Py-S-S-2-Py system at pH6-7 provides no evidence for general base catalysis but is in accord with a mechanism involving specific acid catalysis and post-transition-state proton transfer. 6. The pH-dependences of the overall equilibrium positions of the various thiol-disulphide interchanges are described. 7. Reaction of thioacetamide with a stoicheiometric quantity of 2-Py-S-S-2-Py at pH1 provides 2 molecules of Py-2-SH per molecule of thioacetamide and elemental sulphur; these findings can be accounted for by thiol-disulphide interchange to provide a thioacetamide-Py-2-SH mixed disulphide followed by fragmentation to provide CH(3)CN, S and Py-2-SH. 8. Provision of high reactivity in the neutral forms of the members of this series of sulphur nucleophiles by electron donation by the amino group is compared with the well known alpha effect that provides enhanced nucleophilicity in compounds containing an electronegative atom adjacent to the nucleophilic atom. 9. The decrease in the u.v. absorption of l-ergothioneine at 257nm consequent on transformation of its aminothione moiety into an S-alkyl-2-mercaptoimidazole moiety provides a convenient method of following the alkylation of l-ergothioneine by iodoacetamide. 10. The pH dependence of the extinction coefficient of l-ergothioneine at 257nm is described by epsilon(257)={8x10(3)/(1+K(a)/[H(+)]} +6x10(3)m(-1).cm(-1) in which pK(a)=10.8. 11. In the pH range 3-11 the reaction is characterized by two reactive protonic states (X and XH). 12. The X state, reaction of the ionized 2-mercaptoimidazole moiety of the l-ergothioneine dianion with neutral iodoacetamide, is characterized by the second-order rate constant 4.0m(-1).s(-1) (25.0 degrees C, I=0.05). The XH state, characterized by the second-order rate constant 0.03m(-1).s(-1), is interpreted as reaction of the thione form of the neutral 2-mercaptoimidazole moiety of the l-ergothioneine monoanion with neutral iodoacetamide. 13. The XH state of the alkylation reaction does not exhibit a deuterium isotope effect.  相似文献   

10.
Interactions of calf spleen purine nucleoside phosphorylase (PNP) with a non-typical substrate, 8-azaguanine (8-azaG), and a bisubstrate analogue inhibitor, 9-(2-phosphonylmethoxyethyl)-8-azaguanine (PME-azaG), were investigated by means of steady-state fluorescence spectroscopy. Both 8-azaG and PME-azaG form fluorescent complexes with the enzyme, and dissociation constants are comparable to the appropriate parameters (Km or Ki) obtained from kinetic measurements. PME-azaG inhibits both the phosphorolytic and synthetic pathway of the reaction in a competitive mode. The complex of 8-azaG with PNP is much weaker than the previously reported Gua-PNP complex, and its dissociation constant increases at pH > 7, where 8-azaG exists predominantly as the monoanion (pKa approximately 6.5). The fluorescence difference spectrum of the PNP/8-azaG complex points to participation of the N(7)H or/and N(8)H tautomers of the neutral substrate, and the 9-(2-phosphonylmethoxyethyl) derivative also exists as a neutral species in the complex with PNP. The latter conclusion is based on spectral characteristics of the PNP/PME-azaG complex, confirmed by fluorimetric determination of dissociation constants, which are virtually pH-independent in the range 6-7. These findings testify to involvement of the neutral purine molecule, and not its monoanion, as the substrate in the reverse, synthetic reaction. It is proposed that, in the reverse reaction pathway, the natural purine substrate is bound to the enzyme as the neutral N(7)H tautomer, which is responsible for the reported strong fluorescence of the guanine-PNP complex.  相似文献   

11.
The reaction mechanism of the oxidation of 2,6-dichlorophenol (2,6-DCP) by horseradish peroxidase (HRP) and H2O2 has been investigated and the reaction products have been characterized by UV-visible and mass spectrometry. Evidence for the dimerization of 2,6-DCP to 3,3',5,5'-tetrachloro-4,4'-dihydroxybiphenyl and the subsequent fast oxidation of this product to the corresponding 3,3',5,5'-tetrachlorodiphenoquinone have been collected. The reaction rate was found to decrease markedly as soon as the pH was raised, with a clear inflection point at pH congruent with 6.6-6.9; it also resulted independent from H2O2 concentration. Since the pK(a) for 2,6-DCP is 6.80, the reaction rate might be influenced by the protonation state of the substrate.  相似文献   

12.
The mechanism for HCO3-(-)independent proton permeability in microvillus membrane vesicles (MVV) isolated from human placenta was examined by using the entrapped pH indicator 6-carboxyfluorescein (6CF). Proton fluxes (JH) across MVV were determined in response to induced pH and anion gradients from the time course of 6CF fluorescence, the MVV buffer capacity, and the 6CF vs. pH calibration. In the absence of anions, JH was 12 +/- 2 nequiv s-1 (mg of protein)-1 (pHin 7.4, pHout 6.0, MVV voltage-clamped with K+/valinomycin, 23 degrees C), corresponding to a proton permeability coefficient of 0.02 cm/s, with an activation energy of 9.1 +/- 0.3 kcal/mol. JH was inhibited 20% by dihydro-4,4'-diisothiocyano-2,2'-stilbenedisulfonic acid (H2DIDS) with KI = 8 microM [( Cl-]out = 0 mM). For a 0.5-unit pH gradient JH increased from 1.5 to 4.6 nequiv s-1 (mg of protein)-1 as the internal MVV pH was increased (5.5-7.5). External Cl-, Br-, and I- (but not SO4(2-) and PO4-) increased JH 1.3-2.5-fold for both inwardly and outwardly directed pH gradients with KD = 1.0 +/- 0.4 mM (Br-) and greater than 100 mM (Cl-). This increase was blocked by 100 microM H2DIDS but not by amiloride or furosemide. Internal Cl- did not alter JH induced by pH gradients nor were proton fluxes induced by anion gradients in the absence of a pH gradient. Experiments in which JH was driven by membrane potentials (induced by valinomycin and K+ gradients) indicated that proton transport was voltage-sensitive. These experiments demonstrate a stilbene-sensitive electrogenic proton transport mechanism in MVV that is regulated allosterically by anions at an external binding site.  相似文献   

13.
The effect of temperature and chemical modification on the interaction of the human erythrocyte Band 3 protein (the anion transport protein) with 4-acetamido-4'-isothiocyanostilbene 2,2'-disulfonate (SITS; Ki = 10 microM)-Affi-Gel 102 resin was studied. Band 3 binds to the affinity resin in two states; weakly bound, which is eluted by 1 mM 4-benzamido-4'-aminostilbene 2,2'-disulfonate (BADS; Ki = 2 microM), and strongly bound, which is eluted only under denaturing conditions by 1% lithium dodecyl sulfate (LDS). At 4 degrees C, most of band 3 was present initially in the weakly bound form and very little in the strongly bound form. With longer incubations at 4 degrees C, the weakly bound form was slowly converted to the strongly bound form. At 37 degrees C, most of Band 3 was rapidly converted to the strongly bound form, with some Band 3 still remaining in the weakly bound form. Band 3 dimers, labelled with 4,4'-diisothiocyanostilbene 2,2'-disulfonate (DIDS) in one monomer, did bind to immobilized SITS but did not become tightly bound upon incubation at 37 degrees C. Since the covalent attachment of DIDS to one monomer prevented the adjacent monomer from becoming tightly bound to immobilized SITS ligand, this observation suggests that the inhibitor-binding sites of the two adjacent monomers must be interacting with each other. When the inhibitor site of Band 3 was selectively modified by citrate in the presence of 1-ethyl-3-(3-azonia-4,4-dimethylpentyl)carbodiimide (EAC), Band 3 bound to the resin was more easily eluted by BADS, suggesting reduced affinity for immobilized SITS. However, citrate-modified Band 3 did become tightly bound upon incubation at 37 degrees C.  相似文献   

14.
Haemolysis of red blood cells (RBC) in glycerol media may be measured spectrophotometrically. The haemolytic process in a rapid phase obeys a first order rate law. The rate constant expresses the rate of haemolysis. To gain a better understanding of the mechanism of haemolysis in glycerol media, the effects of pH and band 3 inhibitors on the rate of haemolysis in human and sheep RBC were observed. Over the pH range used (pH 5.8-10.0), the rate of haemolysis decreased with increase in pH in sheep RBC. By contrast, the rate of haemolysis increased from pH 5.8 to 6.4 and decreased above pH 6.4 in human RBC. The different effects of pH on the rate of haemolysis are due to inhibition of glycerol permeability by H(+) in human RBC but not in sheep RBC. This is supported by the different effects of temperature and Cu(2+) on the rate of haemolysis in human and sheep RBC. We did not observe complete inhibition of haemolysis by the classical band 3 inhibitor, 4, 4'-diisothiocyanatostilbene-2,2'-disulfonic acid (DIDS). Another band 3 inhibitor 4,4'-dinitrostilbene-2,2'-disulfonic acid (DNDS) showed only weak inhibition. Phenylgloxal (PG), another band 3 inhibitor, had no effect whatsoever on the rate of haemolysis. These results indicate that the anion pathway of band 3 is not the preferred route of transport of glycerol in mammalian RBC.  相似文献   

15.
A new method has been developed for the chemical modification and labeling of carboxyl groups in proteins. Carboxyl groups are activated with Woodward's reagent K (N-ethyl-5-phenylisoxazolium 3'-sulfonate), and the adducts are reduced with [3H]BH4. The method has been applied to the anion transport protein of the human red blood cell (band 3). Woodward's reagent K is a reasonably potent inhibitor of band 3-mediated anion transport; a 5-min exposure of intact cells to 2 mM reagent at pH 6.5 produces 80% inhibition of transport. The inhibition is a consequence of modification of residues that can be protected by 4,4'-dinitrostilbene-2,2'-disulfonate. Treatment of intact cells with Woodward's reagent K followed by B3H4 causes extensive labeling of band 3, with minimal labeling of intracellular proteins such as spectrin. Proteolytic digestion of the labeled protein reveals that both the 60- and the 35-kDa chymotryptic fragments are labeled and that the labeling of each is inhibitable by stilbenedisulfonate. If the reduction is performed at neutral pH the major labeled product is the primary alcohol corresponding to the original carboxylic acid. Liquid chromatography of acid hydrolysates of labeled affinity-purified band 3 shows that glutamate but not aspartate residues have been converted into the hydroxyl derivative. This is the first demonstration of the conversion of a glutamate carboxyl group to an alcohol in a protein. The labeling experiments reveal that there are two glutamate residues that are sufficiently close to the stilbenedisulfonate site for their labeling to be blocked by 4,4'-diisothiocyanodihydrostilbene-2,2'-disulfonate and 4,4'-dinitrostilbene-2,2'-disulfonate.  相似文献   

16.
Degradation experiment of model polychlorinated biphenyl (PCB) compound 4,4′-dichlorobiphenyl (4,4′-DCB) and its metabolites by the white-rot fungus Phanerochaete chrysosporium and newly isolated 4,4′-DCB-degrading white-rot fungus strain MZ142 was carried out. Although P. chrysosporium showed higher degradation of 4,4′-DCB in low-nitrogen (LN) medium than that in potato dextrose broth (PDB) medium, Phanerochaete sp. MZ142 showed higher degradation of 4,4′-DCB under PDB medium condition than that in LN medium. The metabolic pathway of 4,4′-DCB was elucidated by the identification of metabolites upon addition of 4,4′-DCB and its metabolic intermediates. 4,4′-DCB was initially metabolized to 2-hydroxy-4,4′-DCB and 3-hydroxy-4,4′-DCB by Phanerochaete sp. MZ142. On the other hand, P. chrysosporium transformed 4,4′-DCB to 3-hydroxy-4,4′-DCB and 4-hydroxy-3,4′-DCB produced via a National Institutes of Health shift of 4-chlorine. 3-Hydroxy-4,4′-DCB was transformed to 3-methoxy-4,4′-DCB; 4-chlorobenzoic acid; 4-chlorobenzaldehyde; and 4-chlorobenzyl alcohol in the culture with Phanerochaete sp. MZ142 or P. chrysosporium. LN medium condition was needed to form 4-chlorobenzoic acid, 4-chlorobenzaldehyde, and 4-chlorobenzyl alcohol from 3-hydroxy-4,4′-DCB, indicating the involvement of secondary metabolism. 2-Hydroxy-4,4′-DCB was not methylated. In this paper, we proved for the first time by characterization of intermediate that hydroxylation of PCB was a key step in the PCB degradation process by white-rot fungi.  相似文献   

17.
P Y Chen  A S Verkman 《Biochemistry》1988,27(2):655-660
The mechanisms for Cl transport across basolateral membrane vesicles (BLMV) isolated from rabbit renal cortex were examined by using the Cl-sensitive fluorescent indicator 6-methoxy-N-(3-sulfopropyl)quinolinium (SPQ). The transporters studied included Cl/base exchange, Cl/base/Na cotransport, K/Cl cotransport, and Cl conductance. Initial rates of chloride influx (JCl) were determined from the measured time course of SPQ fluorescence in BLMV following inwardly directed gradients of Cl and gradients of other ions and/or pH. For a 50 mM inwardly directed Cl gradient in BLMV which were voltage and pH clamped (7.0) using K/valinomycin and nigericin, JCl was 0.80 +/- 0.14 nmol S-1 (mg of vesicle protein)-1 (mean +/- SD, n = 8 separate preparations). In the absence of Na and CO2/HCO3 in voltage-clamped BLMV, JCl increased 56% +/- 5% in response to a 1.9 pH unit inwardly directed H gradient; the increase was further enhanced by 40% +/- 3% in the presence of CO2/HCO3 and inhibited 30% +/- 8% by 100 microM dihydro-4,4'-diisothiocyanostilbene-2,2'-disulfonic acid. Na gradients did not increase JCl in the absence of CO2/HCO3; however, an outwardly directed Na gradient in the presence of CO2/HCO3 increased JCl by 31% +/- 8% with a Na KD of 7 +/- 2 mM. These results indicate the presence of Cl/OH and Cl/HCO3 exchange, and Cl/HCO3 exchange trans-stimulated by Na. There was no significant effect of K gradients in the presence or absence of valinomycin, suggesting lack of significant K/Cl cotransport and Cl conductance under experimental conditions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
An earlier study of the reaction of phosphoimidazolide activated nucleosides (ImpN) in aqueous phosphate buffers indicated two modes of reaction of the phosphate monoanion and dianion. The first mode is catalysis of the hydrolysis of the P-N bond in ImpN's which leads to imidazole and nucleoside 5'-monophosphate. The second represents a nucleophilic substitution of the imidazole to yield the nucleoside 5'-diphosphate. This earlier study thus served as a model for the reaction of ImpN with nucleoside monophosphates (pN) because the latter can be regarded as phosphate derivatives. In the present study we investigated the reaction of guanosine 5'-phosphate-2-methylimidazolide, 2-MeImpG, in the presence of pN (N = guanosine, adenosine and uridine) in the range 6.9 less than or equal to pH less than or equal to 7.7. We observed that pN's do act as nucleophiles to form NppG, and as general base to enhance the hydrolysis of the P-N bond in 2-MeImpG, i.e. pN show the same behavior as inorganic phosphate. The kinetic analysis yields the following rate constants for the dianion pN2-: knpN = 0.17 +/- 0.02 M-1 h-1 for nucleophilic attack and khpN = 0.11 +/- 0.07 M-1 h-1 for general base catalysis of the hydrolysis. These rate constants which are independent of the nucleobase compare with kp.2 = 0.415 M-1 h-1 and khp2. = 0.217 M-1 h-1 for the reactions of HPO4(2-). In addition, this study shows that under conditions where pN presumably form stacks, the reaction mechanism remains unchanged although in quantitative terms stacked pN are somewhat less reactive. Attack by the 2'-OH and 3'-OH groups of the ribose moiety in amounts greater than or equal to 1% is not observed; this is attributed to the large difference in nucleophilicity in the neutral pH range between the phosphate group and the ribose hydroxyls. This nucleophilicity rank is not altered by stacking.  相似文献   

19.
Electrolyte transport across the basolateral membrane of the parietal cells   总被引:7,自引:0,他引:7  
The ion-transport properties of the basal lateral membranes of intact isolated parietal cells were studied at the cellular and subcellular level. The presence of an amiloride-sensitive Na+:H+ exchange was demonstrated in cells by proton gradient-driven Na+ uptake and by changes in cell pH as monitored by dimethylcarboxylfluorescein fluorescence both in a fluorimeter and on single isolated cells using a fluorescence microscope and an attached intensified photodiode array spectrophotometer. The presence of the Na+:H+ antiport in vesicles was shown both by intravesicular acidification monitored by acridine orange fluorescent quenching and by proton gradient-dependent Na+ uptake. The presence of Cl-:HCO-3 exchange was determined in intact cells by monitoring changes in cell pH due to Cl- uptake and was shown to be 4,4'-diisothiocyanostilbene-2,2'-disulfonic acid- and 4-acetamido-4'-isothiocyanostilbene-2,2'-disulfonic acid-sensitive. In vesicles, Cl-:HCO-3 exchange was demonstrated by Cl- flux measurement. The apparent affinities for both Cl- and HCO-3 on either side of the membrane were determined to be Km Cli = 20 mM, Km Clout = 17.5 mM, Km HCO-3in = 2.5 mM, and Km HCO-3out = 7.5 mM. A K+ conductance in cells and vesicles was demonstrated by monitoring K+ gradient-dependent 86Rb uptake. No evidence was found for the presence of a Cl- conductance in either cells or vesicles but a H+ conductance was found to be present in vesicles but not in intact cells. In the latter, by determining the effect of either Na+ or Cl- gradients on cell pH and by flux calculations it was concluded that the Cl-:HCO-3 exchange was the major passive flux mechanism for pH regulation in this cell type.  相似文献   

20.
A J Zaug  T R Cech 《Biochemistry》1986,25(16):4478-4482
A shortened form of the Tetrahymena intervening sequence (IVS) RNA acts as an enzyme, catalyzing nucleotidyl transfer and hydrolysis reactions with oligo(cytidylic acid) substrates [Zaug, A. J., & Cech, T. R. (1986) Science (Washington, D.C.) 231, 470-475]. These reactions involve phosphodiester substrates. We now show that the same enzyme has activity toward phosphate monoesters. The 3'-phosphate of C5p or C6p is transferred to the 3'-terminal guanosine of the enzyme. The pH dependence of the reaction (optimum at pH 5) indicates that the enzyme has activity toward the dianion and much greater activity toward the monoanion form of the 3'-phosphate of the substrate. Phosphorylation of the enzyme is reversible by C5-OH and other oligo(pyrimidines) such as UCU-OH. Thus, the RNA enzyme acts as a phosphotransferase, transferring the 3'-terminal phosphate of C5p to UCU-OH with multiple turnover. At pH 4 and 5, the phosphoenzyme undergoes slow hydrolysis to yield inorganic phosphate. Thus, the enzyme has acid phosphatase activity. The RNA enzyme dephosphorylates oligonucleotide substrates with high sequence specificity, which distinguishes it from known protein enzymes.  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