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 rate of peroxidation of unconjugated bilirubin (UCB), catalyzed by horseradish peroxidase (HRP), has been employed by Jacobsen (1969. FEBS Lett. 5: 112-114) to assess the fraction of unbound UCB in the presence of serum albumin. We used this method to examine the interactions of UCB with taurocholate (TC) at pH 8.2, assuming solubilization of UCB by TC is due to pigment binding and/or to partitioning into the micelle, thus rendering UCB unavailable for peroxidation. Inhibition of UCB peroxidation conformed with predictions based on these assumptions and demonstrated significant interaction of UCB with both monomeric and micellar TC. Although significant inhibition of UCB peroxidation was seen with TC monomer, inhibition was even greater with TC micelles. In contrast, pyrogallol, another substrate of HRP, acted very differently in the presence of TC. Inhibition of pyrogallol peroxidation by TC was much less than with UCB and occurred primarily with monomeric TC, with little further inhibition in the micellar range. The results of this study suggest that at taurocholate concentrations above 50 mM, similar to the physiologic bile salt concentrations in human bile, at least 99% of UCB is bound to bile salt, dramatically decreasing the concentration of unbound UCB. Since bile salts also bind Ca2+, they play a dual role in protection against the precipitation of calcium bilirubinate from bile. Therefore, bile salts are a major factor in the prevention of the formation and growth of pigment gallstones.  相似文献   

2.
The interactions of the bile salt sodium taurocholate (TC) in 50 mM Trizma-HCl buffer and 150 mM NaCl (pH 9) at 37 degrees C with membranes composed of sphingomyelin (SM) were studied by dynamic light scattering, cryogenic transmission electron microscopy (cryo-TEM) and turbidity measurements. Small unilamellar SM vesicles were prepared by extrusion. Below the CMC of TC, taurocholate addition leads to vesicle growth due to incorporation of the taurocholate molecules into the vesicle bilayer. At around half the CMC of the bile salt, the SM vesicles are transformed into SM/TC mixed worm-like micelles, which are visualized by cryo-TEM for the first time. Further increase in the taurocholate concentration leads to the rupture of these structures into small spherical micelles. Interestingly, large non-spherical micelles were also identified for pure taurocholate solutions. Similar threadlike structures have been reported earlier for the bile salt sodium taurodeoxycholate [Rich, A., Blow, D., 1958. Nature 182, 1777; Blow, D.M., Rich, A., 1960. J. Am. Chem. Soc. 82, 3566-3571; Galantini, L., Giglio, E., La Mesa, C., Viorel-Pavel, N., Punzo, F., 2002. Langmuir 18, 2812] and for mixtures of taurocholate and phosphatidylcholate [Ulmius, J., Lindblom, G., Wennerstr?m, H., Johansson, L.B.-A., Fontel, K., S?derman, O., Ardvisson, G., 1982. Biochemistry 21, 1553; Hjelm, R.P., Thiyagarajan, P., Alkan-Onyuksel, H., 1992. J. Phys. Chem. 96, 8653] as determined by various scattering methods.  相似文献   

3.
We describe the use and validation of Superose 6, a high performance gel filtration medium for rapid, high resolution separation and sizing of coexisting simple micelles, mixed micelles, and vesicles in bile. We fractionated model biles (1.7-4.2 g/dl total lipid concentration, 0.15 M NaCl) composed of lecithin (L), cholesterol (Ch), and the common bile salt taurocholate (TC) using Superose 6 gel filtration columns (1.0 cm diameter, 30 cm length, 0.5 ml model bile application, 1.0 ml fractions) pre-equilibrated and eluted with 2.5-10.0 mM TC. Lipid particle sizes were determined by quasielastic light scattering and lipid compositions by conventional analyses. In the absence of L and Ch, pure TC "biles" (32.2 mM), when eluted in the presence of 7.5 mM TC, yielded a single peak of particles (mean hydrodynamic radii, Rh values of 11-15 A), consistent with simple TC micelles. Model biles containing L and TC ([L] = 13.8 mM, [TC] = 32.2 mM) were fractionated with baseline resolution into TC-L mixed micelles, (Rh values of 30-40 A) and simple TC micelles. In agreement with the ternary TC-L-H2O phase diagram (Mazer, N. A., et al. 1980. Biochemistry. 19: 601-615), the proportions of simple and mixed micelles were inversely related to L concentrations ([L] = 0-32.2 mM) and correlated positively with eluant TC concentration. Superose 6 gel fractionation of model biles "super-saturated" with Ch (TC:L:Ch molar ratio 27:63:10, total lipid concentration 3 g/dl) yielded high resolution separation of vesicles (Rh value of 320 A) from mixed micelles of TC-L-Ch (Rh values of 40-50 A) and simple TC micelles (Rh values of 11-15 A). At an eluant TC concentration of 7.5 mM, Ch-rich vesicles (Ch/L molar ratio = 1.6) separated that contained 40% of total Ch, 9% of total L, and no TC, accurately reflecting predictions of the quaternary L-Ch-TC-H2O metastable phase diagram (Mazer, N. A., and M. C. Carey. 1983 Biochemistry. 22: 426-442). This suggested that a 7.5 mM TC concentration approximated the intermicellar concentration under the experimental conditions. We also fractionated an identical model bile using conventional Sephacryl S-300, a medium generally used to study model and native biles. Compared with Superose 6, the Sephacryl S-300 column of equivalent size yielded particle separations with lower resolution and speed (30 h v l h).(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
Our studies of equilibrium solubilization of crystals of unconjugated bilirubin (UCB) in buffered aqueous NaCl (1988. J. Lipid Res. 29: 335-348) suggested that the two carboxylic pKa values were 6.8 and 9.3 and the solubility of UCB diacid was 0.1 microM. These data, however, were not ideal, due to possible effects of crystal size, metastability, 96-h incubation times with formation of polar derivatives, impurities in the bilirubin, and imprecision of analyses at low concentrations of UCB ([UCB]). In the present study, designed to determine the pKa values and self-association of UCB, these problems were minimized by solvent partition of UCB from solution in CHCl3 into buffered aqueous NaCl. There was no crystal phase. Equilibrium was attained rapidly (10 min); UCB and CHCl3 were highly purified; and accurate diazo assay of low [UCB] in the aqueous phase, [Bw], was achieved by concentrating the UCB through back-extraction into a small volume of CHCl3. By determining effects on partition rations of varying the [UCB] in the CHCl3 phase, [Bc], we could assess also the self-association of UCB species in the aqueous phase. Partition ratios (P = Bw/Bc) did not differ between initial and repeat extractions, indicating insignificant concentrations of polar UCB derivatives. Similar P ratios were obtained when equilibrium was approached from a supersaturated aqueous phase. At 21-25 degrees C, mu = 0.15, the data (n = 76) fit the equation: log P = log Po + log[1 + 10(pH-A) + 10(2pH-B) + Bc.10(4pH-D)]; the bracketed terms reflect P for H2Bo (diacid), HB- (monoanion), B= (dianion), and (B=)2 dimer, respectively. Computer-fitted values for constants (+/- SD) were: Po = P for H2Bo = 5.79 x 10(-5); A = pK1 = 8.12 +/- 0.23; B = pK1 + pK2 = 16.56 +/- 0.10; pK2 = 8.44 +/- 0.33; D = pk22 + 2(pK1 + pK2) -log(2Po) = 37.64 +/- 0.07, and k22 = 0.26 microM-1 [formation constant of (B=)2 dimer]. In ancillary studies, multiple cycles of direct dissolution of UCB crystals revealed a progressive decrease in aqueous solubility of UCB as fine crystals were removed; this effect was minimal in CHCl3. Unlike in water, moreover, varied UCB crystal forms had similar solubilities in CHCl3, with [Bc] = 1.14 mM at saturation. As determined from [Bc]sat.Po, the aqueous solubility of H2Bo was 66 nM.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

5.
The carbon dioxide solubility coefficient, alphaCO2, and the apparent carbonic acid dissociation constants, K'1 and K'2 were estimated in the serum of the crab Carcinus maenas at various temperatures and ionic strengths. At 15 degrees C, the indirectly determined alphaCO2 value is 0-0499 m-mole l-1 torr-1 for crabs living in normal sea water (salinity ca. 35 percent). It is apparently independent of the serum protein concentration and of the stage of the moulting cycle. For crabs living in undiluted sea water, the mean pK'1 value, determined either gasometrically or titrimetrically, is 6-027 at 15 degrees C. At the same temperature, pK'2=9-29. These values approximate to those of sea water at 35 percent salinity. pK'1 drops as temperature rises; the measured deltapK'1/deltat is -0-0053 pH unit degrees C-1 between 10 and 30 degrees C. PK'1 rises as the ionic strength is lowered. Alignment nomograms have been constructed for the determination of alphaCO2, pK'1 and pK'2 values in relation to various conditions of temperature and salinity.  相似文献   

6.
We have analyzed the manner of incorporation of bile acid into lipid bilayers and resultant perturbation of the bilayer structure with lower bile acid/lipid ratios relevant to the physiological conditions (approximately 1 mM) by 2H and 31P NMR methods, as an aid to understanding the possible role as an endogenous tumor promoter in colon cancer besides the primary physiological function of solubilizing lipids. On the basis of the 2H quadrupole splittings of [6,6,7,7,8-2H5]deoxycholate and [11,11,12,12-2H4]chenodeoxycholate in the presence of lamellar multibilayers of egg yolk lecithin, these bile acids were found to be incorporated in such a manner that the B-D rings lie parallel with the normal of the bilayers when the ratio of the bile acid to lipid is low (less than 0.11). When the ratio is increased, these bile acid molecules are not dispersed entirely in the bilayer but aggregate to form micelles with lipids. Further, we studied the resultant perturbation of the multibilayers of egg yolk lecithin analyzed by using the 2H quadrupole splitting of [18,18,18-2H3]stearic acid as a probe and by 31P chemical shift anisotropy. We found that the bilayer structure is retained even at the bile acid-to-lipid ratio of 0.25, although a small amount of an isotropic phase appeared such as small vesicles and micelles. The molecular ordering of fatty acyl chains was rather enhanced by the presence of 1 mM deoxycholate in erythrocyte ghosts as seen from the 2H quadrupole splitting of [16,16,16-2H3]palmitic acid, although deoxycholate caused hemolysis in this condition. The former observation can be explained by the way the lipid-protein interaction is modified by deoxycholate located in the interface between the lipids and proteins.  相似文献   

7.
Cooperative interaction of histone H1 with DNA.   总被引:4,自引:1,他引:3       下载免费PDF全文
The cooperative binding of histone H1 with DNA was studied using a fluorescently labelled histone H1. The titration data were analysed in terms of the large ligand model. The stoichiometric number, n = 65 +/- 10 bases/H1, was independent of NaCl concentration (0.02 - 0.35 M). The nucleation and the cooperative binding constants, K' and K, and the cooperativity parameter q were sensitive to salt concentration; K = 3.6 +/- 0.8 X 10(7) M-1 and q = 1.1 +/- 0.4 X 10(3) at 0.2 M NaCl. The dependence of K' on NaCl concentration revealed that 6 Na+ ions were released from DNA upon complex formation. An extrapolation of K' to 1M NaCl yielded a small value, K' = 5 +/- 2 M-1. Thus the binding of H1 is essentially electrostatic, being compatible with its independence of temperature. A calculation of K' based on the counterion release reproduced the salt concentration dependence of K'. Therefore, the binding of H1 is of an electrostatic territorial type. Thus, H1 may move along the DNA chain to a certain extent, when both salt concentration and the degree of saturation are sufficiently low. The condition is so restricted that the sliding would not play an important role in vivo. It was concluded from the DNA concentration independent binding isotherm that H1 can cooperatively bind onto a single DNA molecule. A simple power law dependence of the cooperativity parameter q upon NaCl concentration was found; q oc[NaCl]h with h = 0.72, though the physical basis of this dependence remains unknown.  相似文献   

8.
The steady-state distribution of inorganic phosphate and malate between the intra- and extramitochondrial spaces was measured in suspensions of nonrespiring and respiring rat liver mitochondria in which the transmembrane pH difference was incrementally varied. In respiration-inhibited mitochondria, the slope of log [Pi]in/[Pi]out (ordinate) versus delta pH approached 2 by either chemical or isotopic determination of [Pi], and the slope of log [malate]in/[malate]out versus delta pH was 2.0 with an extrapolated log [Pi]in/[Pi]out value of 0.3 at delta pH = 0. We conclude that the distribution of Pi and malate for nonrespiring mitochondria were quantitatively consistent with those predicted by exchange of Pi- for OH- (or cotransport with H+) and of malate 2- for Pi2-. In respiring mitochondria using glutamate + malate as substrate, there was very little pH dependence of Pi or malate accumulation (the slopes were less than 0.5) unless n-butylmalonate (inhibitor of Pi-dicarboxylate exchange) was added before the glutamate and malate, in which case the distribution patterns at delta pH less than 0.4 were similar to those in nonrespiring mitochondria. In either case, however, after reaching a maximal value of 1.1, log [Pi]in/[Pi]out did not further increase with increasing delta pH. Thus, in normally metabolizing mitochondria, the distributions of Pi and malate are not directly correlated with the difference in pH across the membrane.  相似文献   

9.
The control of biliary phospholipid and cholesterol secretions by bile acid was studied by using the technique of retrograde intrabiliary injection. Taurocholate (TC), a moderately hydrophobic bile acid, taurodehydrocholate (TDHC), a hydrophilic non-micelle-forming bile acid, and 3-[(3-cholamidopropyl)-dimethylammonio]propane-1-sulphonate (CHAPS), a detergent, were individually administered by retrograde intrabiliary injection (RII) into the biliary tree, and bile acids, phospholipids and cholesterol subsequently appearing in the bile were measured. TC (1.3 mumol; 45 microliters) injected retrogradely provoked a 3.5-fold increase in biliary phospholipid output for 40 min, as compared with the saline control. Injection of 2.7 mumol of TC (90 microliters) caused a 7.5-fold increase in phospholipid output, which reached a peak at 12 min after RII, and phospholipid output continued for 40 min. Cholesterol output was also elicited under these conditions, showing both dose-dependency and extended secretion. Injection of 1.8 mumol of TDHC caused very little increase in either biliary phospholipid or cholesterol. Injection of 0.9 mumol of CHAPS (45 microliters) provoked a single substantial peak of phospholipid output in the 3 min bile sample. T.l.c. analysis of the phospholipid extracts of the bile collected after each compound showed, for TC, a single compound which co-migrated with the phosphatidylcholine standard, whereas for CHAPS substantial amounts of other phospholipids were present.  相似文献   

10.
Mg2+ increased but Na+ and GTP decrease [3H]substance P (SP) binding to rat cerebral cortical membranes and to 10 mM 3-[(3-cholamidopropyl)dimethylammonio]-1-propane sulfonate (CHAPS)-solubilized membrane fraction. To determine the binding parameters that are modified by the cations and GTP, inhibition experiments of [3H]SP binding by unlabeled SP were performed in both of the preparations. Nonlinear least-squares regression analysis of data in the membrane fraction indicated that optimal fitting of the inhibition curves in the presence of 10 mM MgCl2 was attained with a two-site model, corresponding to a "high-affinity (H)" and a "low-affinity (L)" state. By omitting MgCl2, or by addition of NaCl and GTP, the [3H]SP specific binding was decreased, the H state disappeared, and the L state and a new "super-low affinity (SL)" state observed. The SP/[3H]SP inhibition curves in the cerebral cortical membranes by in vivo treatment with pertussis toxin (islet-activating protein) were similar to that in the presence of GTP in control membranes. The effects of MgCl2, NaCl, and GTP were greater in the CHAPS-solubilized fraction than in the membrane fraction. In contrast to the membrane fraction, the inhibition curves of [3H]SP binding by unlabeled SP in the presence of MgCl2 in the CHAPS-solubilized fraction were best fitted to a one-site model. The KD value was relatively close to that of the low-affinity state in the membrane fraction. Even with the addition of NaCl or GTP, or by reducing MgCl2 concentration to 1 mM, although the inhibition curves consistently fit the one-site model, the KD values changed only slightly.  相似文献   

11.
Noradrenaline-induced accumulation of 3H-labeled inositol mono-, bis-, and trisphosphate (IP1, IP2, and IP3, respectively) in lithium-treated slices of rat cerebral cortex preincubated with [3H]inositol was potentiated by gamma-aminobutyric acid (GABA). However, the effect on [3H]IP2 accumulation was much greater than that on [3H]IP1 or [3H]IP3 accumulation. The principal effect of GABA on noradrenaline concentration-response curves for both [3H]IP1 and [3H]IP2 was to cause an increase in the maximal response attainable. However, whereas the EC50 for GABA potentiation of [3H]IP1 formation was 0.5 mM, the curve for the potentiation of [3H]IP2 formation showed a marked upturn at GABA concentrations of greater than 1 mM. Prazosin (1 microM) blocked the noradrenaline-induced formation of all three inositol phosphates (IPs), in both the presence and the absence of 2 mM GABA. 3H-IP formation induced by phenylephrine and methoxamine was also potentiated by GABA, and again the greatest effect was on [3H]IP2 accumulation. The ratio of [3H]IP2/[3H]IP1 formed in response to 100 microM noradrenaline was increased by 2 mM GABA at all times from 10 to 60 min, whereas the ratio of [3H]IP3/[3H]IP1 was little altered. The effect of GABA was not mimicked by the GABAA agonists isoguvacine and 3-aminopropanesulphonic acid and was not blocked by bicuculline methiodide. (-)-Baclofen, a GABAB agonist, did produce some stimulation of the response to noradrenaline, but to a much lesser extent than GABA. Of the agents tested, nipecotic acid came nearest to reproducing the effect of GABA, in that the major effect was on [3H]IP2 accumulation. The effects of 2 mM GABA and 2 mM nipecotic acid were not additive. GABA potentiation of noradrenaline-induced 3H-IP formation was still apparent in the absence of Li+, but the increase of [3H]IP2 content was less than that of [3H]IP1 content.  相似文献   

12.
Spectrophotometric measurements of bilirubin-IX alpha in water and in aqueous/organic solvent mixtures at pH 10.0 as a function of bilirubin-IX alpha concentration (approx. 0.6--400 microM) are consistent with the formation of dimers (KD - 1.5 microM) in dilute (less than 10 microM) aqueous solution and further self-aggregation to multimers at higher concentrations. Added urea (to 10M) and increases in temperature (to 62 degrees C) obliterate the dimer-multimer transition at 10 microM, but added NaCl (to 0.30 M) promotes strong aggregation of dimers over a narrow concentration range, suggesting a 'micellization' phenomenon. Concentrations of dioxan or ethanol greater than 60% (v/v) in water were required to obtain the absorption spectrum of bilirubin-IX alpha monomers, suggesting that both hydrophobic and electrostatic (pi-orbital) interactions are involved in stabilizing the dimeric state in water. Micellar concentrations of sodium dodecyl sulphate induced spectrophotometric shifts in the dimer absorption spectrum of bilirubin-IX alpha consistent with progressive partitioning of bilirubin-IX alpha monomers into a relatively non-polar region of the micelles and allowed a deduction of the apparent critical micellar concentration that closely approximated the literature values. The pattern of bilirubin IX alpha association with bile salts is complex, since the absorption spectrum shifts hypsochromically below and bathochromically above the critical micellar concentration of the bile salts. Consistent with these observations, bilirubin IX alpha appears to bind to the polar face of bile salt monomers and to the polar perimeter of small bile salt micelles. At higher bile salt concentrations some-bilirubin-IX alpha monomers partition into the hydrophobic interior of the bile salt micelles. Our results suggest that under physiological conditions the natural conjugates of bilirubin-IX alpha may exhibit similar physical chemical properties in bile, in that dimers, highly aggregated multimers and bile salt-associated monomers may co-exist.  相似文献   

13.
Pigment gallstones contain considerable amounts of unconjugated bilirubin (UCB) in the form of calcium bilirubinate and/or bilirubin polymers. Since more than 98% of bile pigments are excreted as conjugates of bilirubin, the source of this UCB needs to be identified. By using a rapid h.p.l.c. method, we compared the non-enzymic hydrolysis of bilirubin monoglucuronide (BMG) and bilirubin diglucuronide (BDG) to UCB in model bile and in native guinea-pig bile. Model biles containing 50 microM solutions of pure BMG and BDG were individually incubated in 25 mM-sodium taurocholate (NaTC) and 0.4 M-imidazole/5 mM-ascorbate buffer (TC-BUF) at 37 degrees C. Over an 8 h period, BMG hydrolysis produced 4-6 times more UCB than BDG hydrolysis. At pH 7.4, 25% of the BMG was converted into UCB, whereas only 4.5% of BDG was converted into UCB. Hydrolysis rates for both BMG and BDG followed the pH order 7.8 greater than 7.6 approximately equal to 7.4 greater than 7.1 Incubation with Ca2+ (6.2 mM) at pH 7.4 in TC-BUF resulted in precipitated bile pigment which, at 100 X magnification, appeared similar to precipitates seen in the bile of patients with pigment gallstones. At pH 7.4, lecithin (crude phosphatidylcholine) (4.2 mM) was a potent inhibitor of hydrolysis of BMG and BDG. The addition of a concentration of cholesterol equimolar with that of lecithin eliminated this inhibitory effect. Guinea-pig gallbladder bile incubated with glucaro-1,4-lactone (an inhibitor of beta-glucuronidase) underwent hydrolysis similar to the model bile systems. The non-enzymic hydrolysis of bile pigments, especially BMG, may be an important mechanism of bile-pigment precipitation and, ultimately, of gallstone formation.  相似文献   

14.
The pathway by which L-lactate (Lac) crosses the plasma membrane of isolated human neutrophils was investigated. The influx of [14C]Lac from a 2 mM Lac, 145 mM Cl-, 5.6 mM glucose medium was approximately 1.5 meq/liter of cell water.min and was sensitive to the organomercurial agent mersalyl (apparent Ki approximately 20 microM), to alpha-cyano-4-hydroxycinnamate (CHC), the classical inhibitor of monocarboxylate transport in mitochondria, and to UK-5099 (apparent Ki approximately 40 microM), a more potent analogue of CHC. Transport was also strongly blocked (greater than 80%) by 1 mM of either 3,5-diiodosalicylic acid, MK-473 (an indanyloxyacetate derivative), or diphenyl-amine-2-carboxylate, and by 0.4 mM pentachlorophenol, but not by 1 mM ethacrynic acid, furosemide, or the disulfonic stilbenes SITS or H2DIDS. One-way [14C]Lac efflux from steady-state cells amounted to approximately 6 meq/liter.min and was likewise affected by the agents listed above. Influx, which was membrane potential insensitive and Na+ independent, displayed a strong pH dependence: extracellular acidification enhanced uptake while alkalinization inhibited the process (pK' approximately 5.7 at 2 mM external Lac). The rate of [14C]Lac influx was a saturable function of external Lac, the Km being approximately 7 mM. Steady-state cells exhibited an intracellular Lac content of approximately 5 mM and secreted lactic acid into the bathing medium a a rate of approximately 4 meq/liter.min. Secretion was completely suppressed by 1 mM mersalyl which inactivates the carrier, leading to an internal accumulation of Lac. That the Lac carrier truly mediates an H+ + Lac- cotransport (or formally equivalent Lac-/OH- exchange) was documented by pH-stat techniques wherein an alkalinization of poorly buffered medium could be detected upon the addition of Lac; these pH changes were sensitive to mersalyl. Thus, the Lac carrier of neutrophils possesses several features in common with other monocarboxylate transport systems in erythrocytes and epithelia.  相似文献   

15.
We investigated effects of low Mg and moderately raised NaCl, as occurs in plantations irrigated with tertiary municipal effluent in New Zealand, on growth, Mg nutrition and photosynthetic activity of Pinus radiata D. Don seedlings grown in nutrient solutions with a Perlite medium. Seedlings were grown with either sufficient (0.35 mM; H[Mg]) or limited (0.033 mM; L[Mg]) Mg supply, without NaCl or with NaCl addition (8.7 mM; +[NaCl]). After 30 weeks, seedlings grown at L[Mg] displayed severe Mg deficiency symptoms, and had significantly less biomass than those at H[Mg]. While NaCl addition had an adverse effect on seedling growth at H[Mg], it increased growth at L[Mg]. The +[NaCl] treatment greatly increased the Mg uptake rates, which were associated with increased stomatal conductance and increased root to shoot ratio. Magnesium deficiency reduced the rates of light-saturated photosynthesis and stomatal conductance, but not the quantum efficiency of photosystem II, which was reduced mainly by the +[NaCl] treatment, especially at H[Mg]. Our study clearly indicated that NaCl addition could counteract the impact of low Mg supply by enhancing Mg uptake from the rooting medium.  相似文献   

16.
We have previously shown that cloned rat multidrug resistance-associated protein 3 (Mrp3) has the ability to transport organic anions such as 17beta-estradiol 17-beta-D-glucuronide (E(2)17betaG) and has a different substrate specificity from MRP1 and MRP2 in that glutathione conjugates are poor substrates for Mrp3 (Hirohashi, T., Suzuki, H., and Sugiyama, Y. (1999) J. Biol. Chem. 274, 15181-15185). In the present study, the involvement of Mrp3 in the transport of endogenous bile salts was investigated using membrane vesicles from LLC-PK1 cells transfected with rat Mrp3 cDNA. The ATP-dependent uptake of [(3)H]taurocholate (TC), [(14)C]glycocholate (GC), [(3)H]taurochenodeoxycholate-3-sulfate (TCDC-S), and [(3)H]taurolithocholate-3-sulfate (TLC-S) was markedly stimulated by Mrp3 transfection in LLC-PK1 cells. The extent of Mrp3-mediated transport of bile salts was in the order, TLC-S > TCDC-S > TC > GC. The K(m) and V(max) values for the uptake of TC and TLC-S were K(m) = 15.9 +/- 4.9 microM and V(max) = 50.1 +/- 9.3 pmol/min/mg of protein and K(m) = 3.06 +/- 0.57 microM and V(max) = 161.9 +/- 21.7 pmol/min/mg of protein, respectively. At 55 nM [(3)H]E(2)17betaG and 1.2 microM [(3)H]TC, the apparent K(m) values for ATP were 1.36 and 0.66 mM, respectively. TC, GC, and TCDC-S inhibited the transport of [(3)H]E(2)17betaG and [(3)H]TC to the same extent with an apparent IC(50) of approximately 10 microM. TLC-S inhibited the uptake of [(3)H]E(2)17betaG and [(3)H]TC most potently (IC(50) of approximately 1 microM) among the bile salts examined, whereas cholate weakly inhibited the uptake (IC(50) approximately 75 microM). Although TC and GC are transported by bile salt export pump/sister of P-glycoprotein, but not by MRP2, and TCDC-S and TLC-S are transported by MRP2, but not by bile salt export pump/sister of P-glycoprotein, it was found that Mrp3 accepts all these bile salts as substrates. This information, together with the finding that MRP3 is extensively expressed on the basolateral membrane of human cholangiocytes, suggests that MRP3/Mrp3 plays a significant role in the cholehepatic circulation of bile salts.  相似文献   

17.
The mechanism of increasing effect of CuCl2 on specific [3H]cimetidine binding was examined in brain membranes of rats. CuCl2-Induced elevation of [3H]cimetidine binding was high in Krebs-Ringer solution (pH 7.4) compared to those in 50 mM Na, K-phosphate buffer (pH 7.4) and in 50 mM Tris-HCl buffer (pH 7.4). CaCl2 (5–50 mM) inhibited effect of CuCl2, but NaCl (25–200 mM), KCl (5–100 mM) or MgCl2 (5–50 mM) did not. CuCl2 (50 μM) elevated 9.3- and 2.5-fold the binding in phosphate- and Tris—HCl buffer, respectively. EDTA-2Na decreased the binding elevated by 50 μM CuCl2 in phosphate buffer to the similar level in Tris-HCl buffer, whereas it did not affect those in Tris-HCl buffer. The absorption spectra of cimetidine and CuCl2 mixture showed a peak at 317 nm in phosphate buffer that was not observed in Tris-HCl buffer. It is suggested that cimetidine-Cu2+ chelate complex could be formed in phosphate buffer, resulting in higher amount of binding in phosphate buffer than in Tris-HCl buffer. PdCl2 also caused a marked elevation in [3H]cimetidine binding, seeming to be due to formation of cimetidine-Pd2+ chelate complex. There were two types of [3H]cimetidine binding in the presence of 20 nM PdCl2: high affinity binding with Kd = 0.7 ± 0.1 nM and low affinity binding with Kd = 44.3 ± 3.0 nM. It is suggested that cimetidine-Cu2+ complex binds to cimetidine binding sites in brain with higher affinity than cimetidine alone.  相似文献   

18.
The uptake of dibucaine into large unilamellar vesicles in response to proton gradients (delta pH; inside acidic) or membrane potentials (delta psi; inside negative) has been investigated. Dibucaine uptake in response to delta pH proceeds rapidly in a manner consistent with permeation of the neutral (deprotonated) form of the drug, reaching a Henderson-Hasselbach equilibrium where [dibucaine]in/[dibucaine]out = [H+]in/[H+]out and where the absolute amount of drug accumulated is sensitive to the buffering capacity of the interior environment. Under appropriate conditions, high absolute interior concentrations of the drug can be achieved (approximately 120 mM) in combination with high trapping efficiencies (in excess of 90%). Dibucaine uptake in response to delta psi proceeds more than an order of magnitude more slowly and cannot be directly attributed to uptake in response to the delta pH induced by delta psi. This induced delta pH is too small (less than or equal to 1.5 pH units) to account for the transmembrane dibucaine concentration gradients achieved and does not come to electrochemical equilibrium with delta psi. Results supporting the possibility that the charged (protonated) form of dibucaine can be accumulated in response to delta psi were obtained by employing a permanently positively charged dibucaine analogue (N-methyldibucaine). Further, the results suggest that delta psi-dependent uptake may depend on formation of a precipitate of the drug in the vesicle interior. The uptake of dibucaine into vesicles in response to ion gradients is of direct utility in drug delivery and controlled release applications and is related to processes of drug sequestration by cells and organelles in vivo.  相似文献   

19.
The velocity of the oxidative renaturation of reduced ribonuclease A catalyzed by protein disulfide isomerase (PDI) is strongly dependent on the composition of a glutathione/glutathione disulfide redox buffer. As with the uncatalyzed, glutathione-mediated oxidative folding of ribonuclease, the steady-state velocity of the PDI-catalyzed reaction displays a distinct optimum with respect to both the glutathione (GSH) and glutathione disulfide (GSSG) concentrations. Optimum activity is observed at [GSH] = 1.0 mM and [GSSG] = 0.2 mM. The apparent kcat at saturating RNase concentration is 0.46 +/- 0.05 mumol of RNase renatured min-1 (mumol of PDI)-1 compared to the apparent first-order rate constant for the uncatalyzed reaction of 0.02 +/- 0.01 min-1. Changes in GSH and GSSG concentration have a similar effect on the rate of both the PDI-catalyzed and uncatalyzed reactions except under the more oxidizing conditions employed, where the catalytic effectiveness of PDI is diminished. The ratio of the velocity of the catalyzed reaction to that of the uncatalyzed reaction increases as the quantity [GSH]2/[GSSG] increases and approaches a constant, limiting value at [GSH]2/[GSSG] greater than 1 mM, suggesting that a reduced, dithiol form of PDI is required for optimum activity. As long as the glutathione redox buffer is sufficiently reducing to maintain PDI in an active form [( GSH]2/[GSSG] greater than 1 mM), the rate acceleration provided by PDI is reasonably constant, although the actual rate may vary by more than an order of magnitude. PDI exhibits half of the maximum rate acceleration at a [GSH]2/[GSSG] of 0.06 +/- 0.01 mM.  相似文献   

20.
1. As used in the Henderson-Hasselbalch equation, [HCO3], [CO2] and pH may all be variously defined; values of pK'1 must be chosen accordingly. 2. In common usage, "HCO3" may include CO3, carbamate, various ion pairs and possibly other bound CO2, as well as free HCO3 ions. 3. pH measurements may be systematically affected by the choice of standard buffers and by proteins and blood cells, and the errors in pH may be pH-dependent. 4. According to how it is expressed, the solubility coefficient for CO2 (S) may be influenced by sample water content, proteins and lipids. However, it need not feature in the calculation. 5. pK'1 is often found to decrease with increasing pH. This may be partly due to inclusion of CO3 and carbamate, but not of H2CO3.HCO3-, in "HCO3" and partly, perhaps, to errors in pH measurement. 6. To the extent that pH measurements are reliable, concentrations or activities of true HCO3 are calculable from pH and PCO2, but, if pH measurements are likely to be systematically erroneous, it may be preferable to define "HCO3" as "total bound CO2" and to base pK'1 on gasometric or titrimetric determinations of that.  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