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 dimethyloxazolidine dione distribution in the extracellular compartments of the frog gastric mucosa was analyzed by washout kinetics. The volumes of the two extracellular compartments, serosal and mucosal, were estimated by inulin washout as 0.435 ± 0.019 and 0.176 ± 0.018 μl/μl tissue water, respectively. In the serosal extracellular space, significant dimethyloxazolidine dione accumulations of 2.63 ± 0.25, 2.28 ± 0.16, and 1.86 ± 0.08 times that of the bathing media were found for bathing solutions with pH values of 6.9, 7.4, and 7.9 respectively. A high pH of the serosal extracellular fluid by itself could not account for the high values of dimethyloxazolidine dione accumulation. A difference in the total dimethyloxazolidine dione accumulation requires: (a) the existence of differences in the pH values and also the existence of a difference in the diffusion coefficient of the two forms of dimethyloxazolidine dione; or (b), a binding of one of the two forms, i.e., binding of dimethyloxazolidine dione form by fixed charges.  相似文献   

2.
Studies were performed to determine the pH relationships among the extracellular, intracellular, and arterial blood compartments in the brain in vivo. Resolution of the extracellular monophosphate resonance peak from the intracellular peak in 31P nuclear magnetic resonance (NMR) spectra of sheep brain with the calvarium intact enabled pH measurement in these respective compartments. Sheep were then subjected to both hyper- and hypoventilation, which resulted in a wide range of arterial PCO2 and pH values. Linear regression analysis of pH in these compartments yielded slopes of 0.56 +/- 0.05 for extracellular pH (pHe) vs. arterial pH, 0.43 +/- 0.078 for intracellular pH (pHi) vs. pHe, and 0.23 +/- 0.056 for pHi vs. arterial pH. These data indicate that CO2 buffering capacity is different and decreases from the intracellular to extracellular to arterial blood compartments. Separation of the extracellular space from the vascular space may be a function of the blood-brain barrier, which contributes to the buffering capability of the extracellular compartment. A marked decrease in the pH gradient between the extracellular and intracellular space occurs during hypercarbia and may influence mechanisms of central respiratory control.  相似文献   

3.
We have used phosphorus-31 nuclear magnetic resonance to determine intracellular pH in the cellular slime mold Dictyostelium discoideum. We devised an air-lift circulator to maintain the dense cell suspensions in a well-oxygenated and well-stirred state while causing minimal perturbation to the sample flowing through the detector coils. Cells continued to develop normally in this set-up. Spectra acquired under these conditions typically show two peaks in the inorganic phosphate region corresponding to pH values of 7.16 +/- 0.03 and 6.48 +/- 0.02. These peaks are believed to represent the mitochondrial and cytosolic compartments respectively, based on a comparison of these values with published data and the collapse of the two compartments upon addition of the mitochondrial uncoupler carbonyl cyanide 4-(trifluoromethoxy)-phenylhydrazone. Dictyostelium cells show a remarkable degree of intracellular pH homeostasis. Both mitochondrial and cytosolic pH remained unchanged as extracellular pH was varied from 4.3 to 8.1. There was also no apparent change in the pH of either compartment after up to 13.5 hours' development in suspension.  相似文献   

4.
The fluid and solute transport properties of pleural tissue were studied by using specimens of intact visceral and parietal pleura from adult sheep lungs. The samples were transferred to the laboratory in a Krebs-Ringer solution at 4 degrees C within 1 h from the death of the animal. The pleura was then mounted as a planar sheet in a Ussing-type chamber. The results that are presented in this study are the means of six different experiments. The spontaneous potential difference and the inhibitory effects of sodium nitroprusside (SNP), ouabain, and amiloride on transepithelial electrical resistance (R(TE)) were measured. The spontaneous potential difference across parietal pleura was 0.5 +/- 0.1 mV, whereas that across visceral pleura was 0.4 +/- 0.1 mV. R(TE) of both pleura was very low: 22.02 +/- 4.1 Omega. cm2 for visceral pleura and 22.02 +/- 3.5 Omega. cm2 for parietal pleura. There was an increase in the R(TE) when SNP was added to the serosal bathing solution of parietal pleura and to the serosal or mucosal bathing solution in visceral pleura. The same was observed when ouabain was added to the mucosal surface of visceral pleura and to either the mucosal or serosal surface of parietal pleura. Furthermore, there was an increase in R(TE) when amiloride was added to the serosal bathing solution of parietal pleura. Consequently, the sheep pleura appears to play a role in the fluid and solute transport between the pleural capillaries and the pleural space. There results suggest that there is a Na+ and K+ transport across both the visceral and parietal pleura.  相似文献   

5.
Plasma membrane vesicles were prepared by phase partition from a microsomal fraction of broad bean (Vicia faba L.) leaf. In order to study the effects of sodium sulfite on active uptake of sucrose, the vesicles were artificially energized by a transmembrane pH gradient (delta pH) and/or a transmembrane electrical gradient (delta psi). At 1 mM, sulfite strongly inhibited sucrose uptake but did not affect the two components of the proton motive force, delta pH (measured by dimethyloxazolidine dione) and delta psi (measured by tetraphenylphosphonium). Moreover, sulfite did not inhibit the proton-pumping ATPase of the plasma membrane vesicles. These data demonstrate that sulfite may inhibit transport of photoassimilates in plant by a direct inhibition of the sucrose carrier of the plasma membrane.  相似文献   

6.
Changes in the extracellular and blood spaces of the uterus were assessed from the distribution volumes of 51Cr-EDTA and 51Cr-labelled red blood cells during the development and regression of the artificially induced decidual cell reaction in ovariectomized, steroid-treated mice. The normally high values for uterine extracellular space (0.35-0.40 microliter/mg) fell to less than 0.20 microliter/mg in association with decidual growth. Uterine blood space increased from around 0.02 microliter/mg to 0.03-0.05 microliter/mg with decidual development. Induction of decidual regression by removal of s.c. progesterone implants caused a rapid decline in tissue blood volume to reach control values (0.01-0.02 microliter/mg) within 24 h and preceded any reduction in uterine weight. Uterine vascular permeability, as determined from the tissue accumulation of 125I-labelled human serum albumin, fell with a similar time course. Tissue extracellular space returned to the higher control values within 48 h of initiating decidual regression.  相似文献   

7.
Toad bladders sacs were placed inside quartz cuvettes. When fura-2 AM was added to the mucosal compartment, low temperature (4 degrees C) almost completely blocked the transepithelial transfer of fluorescence observed at 20 degrees C (20 degrees C = 371 +/- 56, 4 degrees C = 29 +/- 29 fluorescence intensity in arbitrary units (FIAU), excitation at 340 nm, emission at 510 nm). Simultaneously, fluorescence accumulation inside the tissue was significantly higher (20 degrees C = 25 +/- 5, 4 degrees C = 91 +/- 24% increase on basal levels (%IBL)). When fura-2 AM was added to the serosal side, low temperature also reduced the serosal to mucosal transfer (20 degrees C = 149 +/- 36, 4 degrees C = 61 +/- 35 FIAU). Nevertheless, in this situation tissue accumulation, that was significantly higher that the one observed when fura-2 AM was added to the mucosal side, was reduced at low temperature (20 degrees C = 300 +/- 30, 4 degrees C = 48 +/- 7 %IBL). Spectral analysis of the mucosal and serosal compartments indicated that free fura-2 was transferred from the intracellular to the serosal compartment, but not to the mucosal one. These results indicate that fura-2 appears as a useful tool to evaluate the cellular distribution and traffic of polycyclic charged and non-charged molecules.  相似文献   

8.
The active Na(+)-independent transport of L-alanine across the duodenal mucosa of the lizard Gallotia galloti was studied in Ussing-type chambers using a computer-controlled voltage clamp. Addition of L-alanine to the Na(+)-free bathing solutions resulted in a significant L-alanine absorption (J(net)) that was paralleled by an increase in transepithelial short-circuit current (I(sc)) and potential difference (PD) without apparent changes in the tissue conductance. The concentration dependence of J(net), PD, and I(sc) displayed Michaelis-Menten kinetics. L-alanine-induced electrical changes were completely inhibited by external alkaline pH or by the H(+)-ionophore carbonyl cyanide m-chlorophenyl-hydrazone in the bathing solution. The alanine-induced electrogenicity was dependent on the presence of extracellular K(+) and could be blocked by serosal Ba(2+) or mucosal orthovanadate. These results suggest the existence of an H(+)-coupled L-alanine cotransport at the apical membrane of enterocytes. The favorable H(+) driving force is likely to be maintained by an apical vanadate-sensitive H(+)-K(+)-ATPase, allowing the extrusion of H(+) in an exchange with K(+). Potassium exit through a basolateral barium-sensitive conductance provides the key step for the electrogenicity of L-alanine absorption.  相似文献   

9.
The purpose of this study is to assess the effect of an apparent alteration in intracellular pH and the effect of amiloride on the activity of the Na+/H+ antiporter in perfused rat kidney. Rat kidney-Na+ retention was determined using tracer 22Na in perfusate composed of HCl-glycine buffer (pH 3.80 to pH 5.92) or NH4OH-glycine buffer (pH 6.22-7.95) containing Na+ to match physiologic concentrations. Plotting renal Na+ retention for 10 min versus pH in absence of amiloride showed two classical uncompetitive activator curves for H+, one curve from pH 4.19 to 5.10 and another from pH 6.22 to 7.95. H+ acts as an uncompetitive reversible binding substrate with the receptor triggering activation of the exchanger already sequestered with Na+, thus yielding two Ka values for the exchanger suggesting non-first order kinetics. Using an equation derived for uncompetitive-activation binding of Nao+ and Hi+, plotting [mM Na+ mg protein-1 10 min-1]-1 versus [H+], two linear plots are observed on Cartesian coordinates with abscissa intersecting at 47 +/- 1 microM, pKa = 4.32 +/- 0.02 (pH 4.19-5.10) and 4.21 +/- 0.02 microM, pKa = 5.38 +/- 0.01 (pH 6.22-7.95), respectively. Perfusing buffer containing 2 mM amiloride, completely inactivated the antiporter showing stronger inhibition between pH 3.80 and 5.92. Results suggest the presence of two uncompetitive binding sites for H+ with the Na+/H+ exchanger. One is a high affinity binding site at physiological intracellular apparent pH, and another is a low affinity binding site at ischaemic apparent pH, implying the existence of two titration sites for intracellular pH regulation.  相似文献   

10.
Thrombin-induced increase in albumin permeability across the endothelium   总被引:19,自引:0,他引:19  
We studied the effect of thrombin on albumin permeability across the endothelial monolayer in vitro. Bovine pulmonary artery endothelial cells were grown on micropore membranes. Morphologic analysis confirmed the presence of a confluent monolayer with interendothelial junctions. Albumin permeability was measured by the clearance of 125I-albumin across the endothelial monolayer. The control 125I-albumin clearance was 0.273 +/- 0.02 microliter/min. The native enzyme, alpha-thrombin (10(-6) to 10(-10) M), added to the luminal side of the endothelium produced concentration-dependent increases in albumin clearance (maximum clearance of 0.586 +/- 0.08 microliter/min at 10(-6) M). Gamma (gamma) thrombin (10(-6) M and 10(-8) M), which lacks the fibrinogen recognition site, also produced a concentration-dependent increase in albumin clearance similar to that observed with alpha-thrombin. Moreover, the two proteolytically inactive forms of the native enzyme, i-Pr2 P-alpha-thrombin and D-Phe-Pro-Arg-CH2-alpha-thrombin, increased the 125I-albumin clearance (0.610 +/- 0.09 microliter/min and 0.609 +/- 0.02 microliter/min for i-Pr2 P-alpha-thrombin and D-Phe-Pro-Arg-CH2-alpha-thrombin at 10(-6) M, respectively). Since the modified forms of thrombin lack the fibrinogen recognition and active serine protease sites, the results indicate that neither site is required for increased albumin permeability. The increase in albumin clearance with alpha-thrombin was not secondary to endothelial cell lysis because lactate dehydrogenase concentration in the medium following thrombin was not significantly different from baseline values. There was also no morphological evidence of cell lysis. Moreover, the increase in 125I-albumin clearance induced by alpha-thrombin was reversible by washing thrombin from the endothelium. The basis for the increased albumin permeability following the addition of alpha-thrombin appears to be a reversible change in endothelial cell shape with formation of intercellular gaps.  相似文献   

11.
Intracellular and extracellular compartments of phospholipids in the lungs of rats were examined 28 days after intratracheal injection of silica (200 mg/kg). All compartments containing phospholipids were elevated, but the largest increases were seen in the intracellular and extracellular pulmonary surfactant. Intracellular pulmonary surfactant increased 123-fold from 1.18 +/- 0.65 to 144.9 +/- 53.8 and the extracellular surfactant increased 22-fold from 1.17 +/- 0.04 to 25.1 +/- 7.1 mg per pair of rat lungs respectively. The phospholipid composition of intracellular and extracellular surfactant did not change in response to silica, except for an almost 2-fold increase in the percentage of total phosphatidylinositol in both compartments. The phospholipid content of the lungs increased from 24.9 +/- 4.6 to 268.6 +/- 20.8 mg, with the intracellular and extracellular surfactant accounting for 59.1 and 24.6% of this total increase respectively. These data demonstrate that the major increases in the phospholipid content of the lungs induced by silica is associated with the pulmonary-surfactant system.  相似文献   

12.
In a controlled study of 11 male volunteers the following changes (means +/- SD) were observed in venous blood during (D) and 75 min after (A) a period of 20 min of voluntary hyperventilation in comparison with before (B) hyperventilation (P values referring to the difference between D and B) erythrocyte count 5.18 +/- 0.17 X 10(6) (B), 5.70 +/- 0.21 X 10(6) (D) (P less than 0.001), and 5.18 +/- 0.16 X 10(6)/microliter (A); hemoglobin 15.7 +/- 0.6 (B), 17.2 +/- 0.7 (D) (P less than 0.001), and 15.8 +/- 0.6 g/dl (A); centrifuged hematocrit 46.6 +/- 1.0 (B), 50.4 +/- 1.7 (D) (P less than 0.001), and 47.0 +/- 1.8% (A). The platelets increased from 159 +/- 30 X 10(3) (B) to 205 +/- 40 X 10(3) (D) (P less than 0.001) and returned to 157 +/- 26 X 10(3)/microliter (A). The leukocytes (WBC) were 4,210 +/- 630 (B), 6,220 +/- 1,660 (D) (P less than 0.001), and 6,190 +/- 1,870/microliter (A) (P less than 0.002, as compared with B). The rise of WBC during hyperventilation was mainly due to a 83% increase of lymphocytes, whereas a 93% increase of neutrophil leukocytes accounted for the increased WBC 75 min posthyperventilation. The increase of the ratio of band forms to segmented neutrophils from 9 (B) to 19% (A) (P less than 0.01) indicates that band forms were released from the bone marrow. The results show that WBC and platelets can be mobilized by hyperventilation by as yet unidentified mechanisms.  相似文献   

13.
The interdomain disulfide bond present in the C-lobe of all the transferrins was postulated to restrict the domain movement resulting in the slow rate of iron uptake and release. In the present study, the conformational stability and iron binding properties of a derivative of the isolated C-lobe of ovotransferrin in which the interdomain disulfide bond, Cys478-Cys671 was selectively reduced and alkylated with iodoacetamide were compared with the disulfide intact form at the endosomal pH of 5.6. Pyrophosphate and chloride mediated iron release kinetics showed no difference between the disulfide-intact and disulfide-reduced/alkylated forms; the two protein forms yielded similar observed rate constants showing an apparent hyperbolic dependency for anion concentrations. The conformational stability evaluated by unfolding and refolding experiments was greater for the disulfide-intact form than for the disulfide-reduced/alkylated form: the deltaG(D)H2O values at 30 degrees C obtained by using urea were 9.0+/-0.8 and 6.0+/-0.4 kJ/mol for the former and latter protein forms, respectively, and the corresponding values obtained by using guanidine hydrochloride were 6.2+/-0.9 and 4.3+/-0.5 kJ/mol. The dissociation constant of iron (kd) was almost the same for the two protein forms, and it varied only subtly with urea concentrations but increased markedly with GdnHCl concentrations. The nonidentical values of deltaG(D)H2O and kd for urea and GdnHCl can be attributed to the ionic nature of the later denaturant, in which chloride anion may influence the structure and iron uptake-release properties of the ovotransferrin C-lobe. Taken together, we conclude that the interdomain disulfide bond has no effect on the iron uptake and release function but significantly decreases the conformational stability in the C-lobe.  相似文献   

14.
An inducible form of extracellular laccase (EC 1.14.18.1) was isolated from the basidiomycete Coriolus hirsutus. The induction was performed with 0.11 microM syringaldazine, a substrate of laccase. The inducible form of the enzyme consisted of two isoforms, laccase I1 and laccase I2, whose molecular weights were 69 +/- 2 and 67 +/- 2 kDa, respectively. The isoelectric points of these isoenzymes were found to be 3.5 and 4.2, respectively. The optimum pH range for both laccases was 4.4-4.6, and the optimum temperature was 50 degrees C. The thermal stability of these isoenzymes was examined, and KM values for the substrates syringaldazine and pyrocatechol were determined. Our biochemical and physicochemical studies demonstrated that inducible laccase isoforms differed from constitutive forms in molecular weight, IP, KM, and thermal stability. However, their optimum pH ranges and temperatures were identical.  相似文献   

15.
A strategic assessment of the contributions of two active-site hydrogen bonds in the binding of nicotinate to recombinant ferric soybean leghaemoglobin a (rLb) was carried out by mutagenic replacement of the hydrogen-bonding residues (H61A and Y30A variants) and by complementary chemical substitution of the carboxylate functionality on the nicotinate ligand. Dissociation constants, Kd (pH 5.5, mu = 0.10 M, 25.0 +/- 0.1 degrees C), for binding of nicotinate to ferric rLb, H61A and Y30A were 1.4 +/- 0.3 microM, 19 +/- 1 microM and 11 +/- 1 microM, respectively; dissociation constants for binding of nicotinamide were, respectively, 38 +/- 1 mM, 50 +/- 2 mM and 12 +/- 1 mM, and for binding of pyridine were 260 +/- 50 microM, 4.5 +/- 0.5 microM and 66 +/- 8 microM, respectively. Binding of cyanide and azide to the H61A and Y30A variants was unaffected by the mutations. The pH-dependence of nicotinate binding for rLb and Y30A was consistent with a single titration process (pKa values 6.9 +/- 0.1 and 6.7 +/- 0.2, respectively); binding of nicotinate to H61A was independent of pH. Reduction potentials for the rLb and rLb-nicotinate derivatives were 29 +/- 2 mV (pH 5.40, 25.0 degrees C, mu = 0.10 M) and - 65 +/- 2 mV (pH 5.42, 25.0 degrees C, mu = 0.10 M), respectively. The experiments provide a quantitative assessment of the role of individual hydrogen bonds in the binding process, together with a definitive determination of the pKa of His61 and unambiguous evidence that titration of His61 controls binding in the neutral to alkaline region.  相似文献   

16.
Colonic crypts can absorb fluid, but the identity of the absorptive transporters remains speculative. Near the crypt base, the epithelial cells responsible for vectorial transport are relatively undifferentiated and often presumed to mediate only Cl- secretion. We have applied confocal microscopy in combination with an extracellular fluid marker [Lucifer yellow (LY)] or a pH-sensitive dye (2',7'-bis(2-carboxyethyl)-5(6)-carboxyfluorescein) to study mouse colonic crypt epithelial cells directly adjacent to the crypt base within an intact mucosal sheet. Measurements of intracellular pH report activation of colonocyte Na+/H+ exchange in response to luminal or serosal Na+. Studies with LY demonstrate the presence of a paracellular fluid flux, but luminal Na+ does not activate Na+/H+ exchange in the nonepithelial cells of the lamina propria, and studies with LY suggest that the fluid bathing colonocyte basolateral membranes is rapidly refreshed by serosal perfusates. The apical Na+/H+ exchange in crypt colonocytes is inhibited equivalently by luminal 20 microM ethylisopropylamiloride and 20 microM HOE-694 but is not inhibited by luminal 20 microM S-1611. Immunostaining reveals the presence of epitopes from NHE1 and NHE2, but not NHE3, in epithelial cells near the base of colonic crypts. Comparison of apical Na+/H+ exchange activity in the presence of Cl- with that in the absence of Cl- (substitution by gluconate or nitrate) revealed no evidence of the Cl--dependent Na+/H+ exchange that had been previously reported as the sole apical Na+/H+ exchange activity in the colonic crypt. Results suggest the presence of an apical Na+/H+ exchanger near the base of crypts with functional attributes similar to those of the cloned NHE2 isoform.  相似文献   

17.
The symmetry of osmotic conductivity of the canine tracheal epithelial cells was examined in vitro. When an osmotic load of 100 mosM sucrose was added to the serosal bathing solution, no change in the transepithelial potential difference was observed in 15 tissue preparations. In contrast, when the same osmotic load was added to the mucosal bathing solution, there was a rapid decrease in the transepithelial potential difference of 3.9 +/- 0.5 mV (n = 23); ouabain (10(-4) M) eliminated this change. Tissues that had been exposed to the osmotic load added to either the mucosal or serosal side were compared with the control using light and electron microscopy. When the osmotic load was added to the mucosal fluid, there was no change in the nuclear-to-cytoplasmic area ratio of the cell types examined. However, when the same osmotic load was added to the serosal fluid, a marked increase in the nuclear-to-cytoplasmic area ratio of the ciliated cells was observed. This finding indicated cell shrinkage. Dilution potentials measured by substituting NaCl with mannitol also showed asymmetry. The morphological features are probably caused by differences in the osmotic conductivity (Lp) of the basolateral and apical cell membranes, with the Lp of the apical membrane being less than that of the basolateral membrane. The basis for osmotically induced potentials remained undetermined.  相似文献   

18.
The rate of active H+ secretion (JH) across the luminal cell membrane of the turtle bladder decreases linearly with the chemical (delta pH) or electrical potential gradient (delta psi) against which secretion occurs. To examine the control of JH from the cell side of the pump, acid-base changes were imposed on the cellular compartment by increasing serosal[HCO3-] at constant PCO2 or by varying PCO2 at constant [HCO3-]. When serosal [HCO3-] was increased from 0 to 60 mM, cell [H+] decreased, as estimated by the 5,5-dimethyloxazoladine-2,4- dione method. JH was a saturable function of cell [H+], with an apparent Km of 25 nM. When PCO2 was varied between 1 and 20% at various serosal Km of 25 nM. When PCO2 was varied between 1 and 20% at various serosal [HCO3-], the PCO2 required to reach a maximal JH increased with [HCO3-] so that JH was a function of cell [H+] rather than of cell [HCO3-] or CO2. The proton pump was controlled asymmetrically with respect to the pH component of the electrochemical potential for protons, microH. On the cell side of the pump, a delta pH of < 1 U was required to vary JH between maximal and zero values, whereas on the luminal side a delta pH of 3 U was required. Cell [H+] regulates JH by determining the availability of H+ to the pump in a relationship resembling Michaelis-Menten kinetics. Increasing luminal [H+] generates an energy barrier at a luminal pH near 4.4 that equals the free energy (per H+ translocated) of the metabolic driving reaction.  相似文献   

19.
Calcium-dependence of sugar transport in rat small intestine   总被引:1,自引:0,他引:1  
The involvement of Ca2+ in the theophylline action on sugar transport was investigated in isolated rat small intestinal mucosa. Theophylline significantly increased cell water free sugar accumulation and reduced mucosal to serosal sugar fluxes both in the presence and absence of calcium, but the effects of theophylline were significantly less in calcium free media. In theophylline untreated tissues, calcium-deprived bathing solutions decreased tissue galactose accumulation and increased mucosal to serosal sugar flux. The calcium-channel blocker verapamil produced similar effects on intestinal galactose transport to those induced by low extracellular calcium activity. RMI 12330A and the calmodulin antagonist trifluoperazine abolished the theophylline-effects on intestinal galactose transport. Both drugs also affected sugar transport in basal conditions. These studies suggest that calcium might modulate sugar permeability across the basolateral boundary of rat enterocytes, and that its effect may be mediated by calmodulin.  相似文献   

20.
Pepsin inhibition by 3-alkoxy-4-arylpiperidine (substituted piperidine; (3R,4R)-3-(4-bromobenzyloxy)-4-[4-(2-naphthalen-1-yl-2-oxo-ethoxy)phenyl]piperidine) has been studied using steady-state kinetic and pre-equilibrium binding methods. Data were compared with pepstatin A, a well known competitive inhibitor of pepsin. Steady-state analysis reveals that the substituted piperidine likewise behaves as a competitive inhibitor. Pre-equilibrium binding studies indicate that the substituted piperidine can displace a fluorescently labeled statine inhibitor from the enzyme active site. Simulation of the stopped-flow fluorescence transients provided estimates of the K(d) values of 1.4 +/- 0.2 microm and 39 +/- 2 nm for the piperidine and the fluorescently labeled statine, respectively. The effects of combinations of these two inhibitors resulted in a series of parallel lines when plotted by the method of Yonetani and Theorell (Yonetani, T., and Theorell, H. (1964) Arch. Biochem. Biophys. 106, 234-251), suggesting that the two inhibitors bind in a mutually exclusive fashion to pepsin. Fitting of the entire data set to the appropriate equation yielded an alpha factor of 8 +/- 1. The magnitude of this factor ( infinity > alpha > 1) can be explained by a conformational distinction between the enzyme species that bind each inhibitor. The effects of pH on the inhibition constants for pepstatin A and the substituted piperidine also suggest that the inhibitors bind to distinct conformational forms of the enzyme. No inhibition by the piperidine was observed at acidic pH, while pepstatin A inhibition is maximal at low pH values. Inhibition by the piperidine was maximal when a group with pK 4.8 +/- 0.2 was deprotonated and another group with pK 5.9 +/- 0.2 was protonated. Most likely these two groups are the catalytic aspartates with perturbed ionization properties as a result of a significant and unique conformational change. Taken together, these data suggest that the enzyme can readily interconvert between two conformers, one capable of binding substrate and pepstatin A and the other capable of binding the substituted piperidine.  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