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 812 毫秒
1.
Ono T  Rompel A  Mino H  Chiba N 《Biophysical journal》2001,81(4):1831-1840
Effects of adding monovalent alkali metal cations to Ca(2+)-depleted photosystem (PS)II membranes on the biochemical and spectroscopic properties of the oxygen-evolving complex were studied. The Ca(2+)-dependent oxygen evolution was competitively inhibited by K(+), Rb(+), and Cs(+), the ionic radii of which are larger than the radius of Ca(2+) but not inhibited significantly by Li(+) and Na(+), the ionic radii of which are smaller than that of Ca(2+). Ca(2+)-depleted membranes without metal cation supplementation showed normal S(2) multiline electron paramagnetic resonance (EPR) signal and an S(2)Q(A)(-) thermoluminescence (TL) band with a normal peak temperature after illumination under conditions for single turnover of PSII. Membranes supplemented with Li(+) or Na(+) showed properties similar to those of the Ca(2+)-depleted membranes, except for a small difference in the TL peak temperatures. The peak temperature of the TL band of membranes supplemented with K(+), Rb(+), or Cs(+) was elevated to approximately 38 degrees C which coincided with that of Y(D)(+)Q(A)(-) TL band, and no S(2) EPR signals were detected. The K(+)-induced high-temperature TL band and the S(2)Q(A)(-) TL band were interconvertible by the addition of K(+) or Ca(2+) in the dark. Both the Ca(2+)-depleted and the K(+)-substituted membranes showed the narrow EPR signal corresponding to the S(2)Y(Z)(+) state at g = 2 by illuminating the membranes under multiple turnover conditions. These results indicate that the ionic radii of the cations occupying Ca(2+)-binding site crucially affect the properties of the manganese cluster.  相似文献   

2.
Apoflavodoxin from the sulfate reducing bacteria Desulfovibrio desulfuricans is a small, acidic protein with a net charge of -19 at neutral pH. Here, we show that monovalent cations in biologically relevant amounts have dramatic effects on apoflavodoxin stability. The effect is largest for Gdm(+) and decreases as a function of increased cation charge density (Gdm(+)>NH(4)(+)K(+) approximately Cs(+) approximately Na(+)>Li(+)). A linear correlation of stabilizing effects with cation hydration properties suggests an important role of dehydration in efficient cation interaction with the protein. The effects on stability are due to preferential binding of one cation to native apoflavodoxin and results in an increase in thermal midpoint of 20 degrees C and the free energy of unfolding (at 20 degrees C) increases fivefold. Tuning of biophysical properties (such as folding and ligand/cofactor binding) of acidic proteins by cation binding may be important in vivo.  相似文献   

3.
Salts inhibit the activity of sweet almond beta-glucosidase. For cations (Cl(-) salts) the effectiveness follows the series: Cu(+2), Fe(+2)>Zn(+2)>Li(+)>Ca(+2)>Mg(+2)>Cs(+)>NH(4)(+)>Rb(+)>K(+)>Na(+) and for anions (Na(+) salts) the series is: I(-)>ClO(4)(-)>(-)SCN>Br(-) approximately NO(3)(-)>Cl(-) approximately (-)OAc>F(-) approximately SO(4)(-2). The activity of the enzyme, like that of most glycohydrolases, depends on a deprotonated carboxylate (nucleophile) and a protonated carboxylic acid for optimal activity. The resulting pH-profile of k(cat)/K(m) for the beta-glucosidase-catalyzed hydrolysis of p-nitrophenyl glucoside is characterized by a width at half height that is strongly sensitive to the nature and concentration of the salt. Most of the inhibition is due to a shift in the enzymic pK(a)s and not to an effect on the pH-independent second-order rate constant, (k(cat)/K(m))(lim). For example, as the NaCl concentration is increased from 0.01 M to 1.0 M the apparent pK(a1)increases (from 3.7 to 4.9) and the apparent pK(a2)decreases (from 7.2 to 5.9). With p-nitrophenyl glucoside, the value of the pH-independent (k(cat)/K(m))(lim) (=9 x 10(4) M(-1) s(-1)) is reduced by less than 4% as the NaCl concentration is increased. There is a similar shift in the pK(a)s when the LiCl concentration is increased to 1.0 M. The results of these salt-induced pK(a) shifts rule out a significant contribution of reverse protonation to the catalytic efficiency of the enzyme. At low salt concentration, the fraction of the catalytically active monoprotonated enzyme in the reverse protonated form (i.e., proton on the group with a pK(a) of 3.7 and dissociated from the group with a pK(a) of 7.2) is very small ( approximately 0.03%). At higher salt concentrations, where the two pK(a)s become closer, the fraction of the monoprotonated enzyme in the reverse protonated form increases over 300-fold. However, there is no increase in the intrinsic reactivity, (k(cat)/K(m))(lim), of the monoprotonated species. For other enzymes which may show such salt-induced pK(a) shifts, this provides a convenient test for the role of reverse protonation.  相似文献   

4.
The aim of the present work was to characterize Na(+) currents through nonselective cation channels (NSCCs) in protoplasts derived from root cells of Arabidopsis. The procedure of the protoplast isolation was modified to increase the stability of Arabidopsis root protoplasts in low external Ca(2+) by digesting tissue in elevated Ca(2+). Experiments in whole-cell and outside-out modes were carried out. We found that Na(+) currents in Arabidopsis root protoplasts were mediated by cation channels that were insensitive to externally applied tetraethylammonium(+) and verapamil, had no time-dependent activation (permanently opened or completely activated within 1-2 ms), were voltage independent, and were weakly selective for monovalent cations. The selectivity sequence was as follows: K(+) (1.49) > NH(4)(+) (1.24) > Rb(+) (1.15) approximately equal to Cs(+) (1.10) approximately equal to Na(+) (1.00) > Li(+) (0.73) > tetraethylammonium(+) (0.47). Arabidopsis root NSCCs were blocked by H(+) (pK approximately equal to 6.0), Ca(2+) (K(1/2) approximately equal to 0.1 mM), Ba(2+), Zn(2+), La(3+), Gd(3+), quinine, and the His modifier diethylpyrocarbonate. They were insensitive to most organic blockers (nifedipine, verapamil, flufenamate, and amiloride) and to the SH-group modifier p-chloromercuriphenyl sulfonic acid. Voltage-insensitive, Ca(2+)-sensitive single channels were also resolved. Properties of Arabidopsis root NSCCs are discussed and compared with characteristics of similar conductances studied previously in plants and animals. It is suggested that NSCCs present a distinct group of plant ion channels, mediating toxic Na(+) influx to the cell and probably having other important roles in physiological processes of plants.  相似文献   

5.
The effect of polyelectrolyte cation specificity, charge density, and conformation on the interaction between furcelleran, kappa, iota, and lambda-carrageenan, respectively, and amitriptyline, an amphiphilic cationic drug molecule, was studied by means of a dialysis equilibrium technique. The carrageenans used in this study-furcelleran, kappa, iota, and lambda-carrageenan-had a charge density corresponding to 0.69, 0.92, 1.53, and 2.07 sulfate groups per disaccharide. In general, the binding isotherms followed the order Li(+) < Na(+) < N(CH3)(4)(+) < K(+) < or = Cs(+) approximately Rb(+), i.e., the binding isotherms were shifted to higher concentrations of free amphiphile according to the sequence indicated. This affinity sequence correlates well with that found for the dextran sulfate-amitriptyline system (A. Hugerth and L.-O. Sundel?f, Langmuir 2000, 16, 313-317). The factor determining the critical aggregation concentration in the presence of Na(+) compared to K(+) was found to be as follows: the flexibility (conformation) in the case of the lower charged carrageenans, i.e., furcelleran and kappa-carrageenan, charge density for iota-carrageenan, and in the lambda-carrageenan case the difference in the ROSO(3)(-)-alkali ion specificity.  相似文献   

6.
Ba(2+), a doubly charged analogue of K(+), specifically blocks K(+) channels by virtue of electrostatic stabilization in the permeation pathway. Ba(2+) block is used here as a tool to determine the equilibrium binding affinity for various monovalent cations at specific sites in the selectivity filter of a noninactivating mutant of KcsA. At high concentrations of external K(+), the block-time distribution is double exponential, marking at least two Ba(2+) sites in the selectivity filter, in accord with a Ba(2+)-containing crystal structure of KcsA. By analyzing block as a function of extracellular K(+), we determined the equilibrium dissociation constant of K(+) and of other monovalent cations at an extracellular site, presumably S1, to arrive at a selectivity sequence for binding at this site: Rb(+) (3 μM) > Cs(+) (23 μM) > K(+) (29 μM) > NH(4)(+) (440 μM) > Na(+) and Li(+) (>1 M). This represents an unusually high selectivity for K(+) over Na(+), with |ΔΔG(0)| of at least 7 kcal mol(-1). These results fit well with other kinetic measurements of selectivity as well as with the many crystal structures of KcsA in various ionic conditions.  相似文献   

7.
Both Cs(+) and NH(4)(+) alter neuronal Cl(-) homeostasis, yet the mechanisms have not been clearly elucidated. We hypothesized that these two cations altered the operation of the neuronal K(+)-Cl(-) cotransporter (KCC2). Using exogenously expressed KCC2 protein, we first examined the interaction of cations at the transport site of KCC2 by monitoring furosemide-sensitive (86)Rb(+) influx as a function of external Rb(+) concentration at different fixed external cation concentrations (Na(+), Li(+), K(+), Cs(+), and NH(4)(+)). Neither Na(+) nor Li(+) affected furosemide-sensitive (86)Rb(+) influx, indicating their inability to interact at the cation translocation site of KCC2. As expected for an enzyme that accepts Rb(+) and K(+) as alternate substrates, K(+) was a competitive inhibitor of Rb(+) transport by KCC2. Like K(+), both Cs(+) and NH(4)(+) behaved as competitive inhibitors of Rb(+) transport by KCC2, indicating their potential as transport substrates. Using ion chromatography to measure unidirectional Rb(+) and Cs(+) influxes, we determined that although KCC2 was capable of transporting Cs(+), it did so with a lower apparent affinity and maximal velocity compared with Rb(+). To assess NH(4)(+) transport by KCC2, we monitored intracellular pH (pH(i)) with a pH-sensitive fluorescent dye after an NH(4)(+)-induced alkaline load. Cells expressing KCC2 protein recovered pH(i) much more rapidly than untransfected cells, indicating that KCC2 can mediate net NH(4)(+) uptake. Consistent with KCC2-mediated NH(4)(+) transport, pH(i) recovery in KCC2-expressing cells could be inhibited by furosemide (200 microM) or removal of external [Cl(-)]. Thermodynamic and kinetic considerations of KCC2 operating in alternate transport modes can explain altered neuronal Cl(-) homeostasis in the presence of Cs(+) and NH(4)(+).  相似文献   

8.
Synthesis and photophysical properties of three Ru(bpy)(3)(2+)-Ptz (bpy = 2,2'-bipyridine and Ptz = phenothiazine) dyads, where the number of Ptz groups increased from one to three, are reported. The MLCT absorption bands of these compounds were slightly red shifted compared to Ru(bpy)(3)(2+). The emission, however, was highly quenched and this is attributed to electron transfer from the Ptz moiety to the excited Ru(bpy)(3)(2+) to generate the charge separated state Ru(bpy)(3)(+)-Ptz (+). Observed electron transfer rates (k(et) > 10(8) s(-1)) were much faster than those previously reported (k(et) < 10(7) s(-1)) for linked Ru(bpy)(3)(2+)-Ptz systems. Compared to the previous systems, back electron transfer rates in these systems were about 100 times slower. This has enabled us to observe the charge separated state in nanosecond flash photolysis experiments. Transient absorptions assignable to Ru(bpy)(3)(+) and Ptz (+), having lifetimes in the range of 10-30 ns were observed. In order to explain the fast charge separation and slow charge recombination rates, formation of a folded conformer where the Ptz group attached to one bpy residue comes closer to and associates with another bpy moiety was invoked. A scheme which explains the fast electron transfer and slow recombination in this pre-associated state is proposed.  相似文献   

9.
The NH(4)(+) and NO(3)(-) uptake kinetics by Typha latifolia L. were studied after prolonged hydroponics growth at constant pH 3.5, 5.0, 6.5 or 7.0 and with NH(4)(+) or NO(3)(-) as the sole N-source. In addition, the effects of pH and N source on H(+) extrusion and adenine nucleotide content were examined. Typha latifolia was able to grow with both N sources at near neutral pH levels, but the plants had higher relative growth rates, higher tissue concentrations of the major nutrients, higher contents of adenine nucleotides, and higher affinity for uptake of inorganic nitrogen when grown on NH(4)(+). Growth almost completely stopped at pH 3.5, irrespective of N source, probably as a consequence of pH effects on plasma membrane integrity and H(+) influx into the root cells. Tissue concentrations of the major nutrients and adenine nucleotides were severely reduced at low pH, and the uptake capacity for inorganic nitrogen was low, and more so for NO(3)(-)-fed than for NH(4)(+)-fed plants. The maximum uptake rate, V(max), was highest for NH(4)(+) at pH 6.5 (30.9 micro mol h(-1) g(-1) root dry weight) and for NO(3)(-) at pH 5.0 (31.7 micro mol h(-1) g(-1) root dry weight), and less than 10% of these values at pH 3.5. The affinity for uptake as estimated by the half saturation constant, K((1/2)), was lowest at low pH for NH(4)(+) and at high pH for NO(3)(-). The changes in V(max) and K((1/2)) were thus consistent with the theory of increasing competition between cations and H(+) at low pH and between anions and OH(-) at high pH. C(min) was independent of pH, but slightly higher for NO(3)(-) than for NH(4)(+) (C(min)(NH(4)(+)) approximately 0.8 mmol m(-3); C(min)(NO(3)(-)) approximately 2.8 mmol m(-3)). The growth inhibition at low pH was probably due to a reduced nutrient uptake and a consequential limitation of growth by nutrient stress. Typha latifolia seems to be well adapted to growth in wetland soils where NH(4)(+) is the prevailing nitrogen compound, but very low pH levels around the roots are very stressful for the plant. The common occurrence of T. latifolia in very acidic areas is probably only possible because of the plant's ability to modify pH-conditions in the rhizosphere.  相似文献   

10.
Cho M  Kim Y  Han SY  Min K  Rahman MA  Shim YB  Ban C 《BMB reports》2008,41(2):126-131
The folding of aptamer immobilized on an Au electrode was successfully detected using label-free electrochemical methods. A thrombin binding DNA aptamer was used as a model system in the presence of various monovalent cations. Impedance spectra showed that the extent to which monovalent cations assist in folding of aptamer is ordered as K(+) > NH(4)(+) > Na(+) > Cs(+). Our XPS analysis also showed that K(+) and NH(4)(+) caused a conformational change of the aptamer in which it forms a stable complex with these monovalent ions. Impedance results for the interaction between aptamer and thrombin indicated that thrombin interacts more with folded aptamer than with unfolded aptamer. The EQCM technique provided a quantitative analysis of these results. In particular, the present impedance results showed that thrombin participates a folding of aptamer to some extent, and XPS analysis confirmed that thrombin stabilizes and induces the folding of aptamer.  相似文献   

11.
S Zhang  S J Kehl    D Fedida 《Biophysical journal》2001,81(1):125-136
Zinc ions are known to induce a variable depolarizing shift of the ionic current half-activation potential and substantially slow the activation kinetics of most K(+) channels. In Kv1.5, Zn(2+) also reduces ionic current, and this is relieved by increasing the external K(+) or Cs(+) concentration. Here we have investigated the actions of Zn(2+) on the gating currents of Kv1.5 channels expressed in HEK cells. Zn(2+) shifted the midpoint of the charge-voltage (Q-V) curve substantially more (approximately 2 times) than it shifted the V(1/2) of the g-V curve, and this amounted to +60 mV at 1 mM Zn(2+). Both Q1 and Q2 activation charge components were similarly affected by Zn(2+), which indicated free access of Zn(2+) to channel closed states. The maximal charge movement was also reduced by 1 mM Zn(2+) by approximately 15%, from 1.6 +/- 0.5 to 1.4 +/- 0.47 pC (n = 4). Addition of external K(+) or Cs(+), which relieved the Zn(2+)-induced ionic current reduction, decreased the extent of the Zn(2+)-induced Q-V shift. In 135 mM extracellular Cs(+), 200 microM Zn(2+) reduced ionic current by only 8 +/- 1%, compared with 71% reduction in 0 mM extracellular Cs(+), and caused a comparable shift in both the g-V and Q-V relations (17.9 +/- 0.6 mV vs. 20.8 +/- 2.1 mV, n = 6). Our results confirm the presence of two independent binding sites involved in the Zn(2+) actions. Whereas binding to one site accounts for reduction of current and binding to the other site accounts for the gating shift in ionic current recordings, both sites contribute to the Zn(2+)-induced Q-V shift.  相似文献   

12.
The thionucleoside 2-thiocytidine (C2S) occurs in nature in transfer RNAs; it receives attention in diverse fields like drug research and nanotechnology. By potentiometric pH titrations we measured the acidity constants of H(C2S)(+) and the stability constants of the M(C2S)(2+) and M(C2S-H)(+) complexes (M(2+) = Zn(2+), Cd(2+)), and we compared these results with those obtained previously for its parent nucleoside, cytidine (Cyd). Replacement of the (C2)=O unit by (C2)=S facilitates the release of the proton from (N3)H(+) in H(C2S)(+) (pK (a) = 3.44) somewhat, compared with H(Cyd)(+) (pK (a) = 4.24). This moderate effect of about 0.8 pK units contrasts with the strong acidification of about 4 pK units of the (C4)NH(2) group in C2S (pK (a) = 12.65) compared with Cyd (pK (a) approximately 16.7); the reason for this result is that the amino-thione tautomer, which dominates for the neutral C2S molecule, is transformed upon deprotonation into the imino-thioate form with the negative charge largely located on the sulfur. In the M(C2S)(2+) complexes the (C2)S group is the primary binding site rather than N3 as is the case in the M(Cyd)(2+) complexes, though owing to chelate formation N3 is to some extent still involved in metal ion binding. Similarly, in the Zn(C2S-H)(+) and Cd(C2S-H)(+) complexes the main metal ion binding site is the (C2)S(-) unit (formation degree above 99.99% compared with that of N3). However, again a large degree of chelate formation with N3 must be surmised for the M(C2S-H)(+) species in accord with previous solid-state studies of related ligands. Upon metal ion binding, the deprotonation of the (C4)NH(2) group (pK (a) = 12.65) is dramatically acidified (pK (a) approximately 3), confirming the very high stability of the M(C2S-H)(+) complexes. To conclude, the hydrogen-bonding and metal ion complex forming capabilities of C2S differ strongly from those of its parent Cyd; this must have consequences for the properties of those RNAs which contain this thionucleoside.  相似文献   

13.
Wang Z  Fedida D 《Biophysical journal》2001,81(5):2614-2627
Sustained Na(+) or Li(+) conductance is a feature of the inactivated state in wild-type (WT) and nonconducting Shaker and Kv1.5 channels, and has been used here to investigate the cause of off-gating charge immobilization in WT and Kv1.5-W472F nonconducting mutant channels. Off-gating immobilization in response to brief pulses in cells perfused with NMG/NMG is the result of a more negative voltage dependence of charge recovery (V(1/2) is -96 mV) compared with on-gating charge movement (V(1/2) is -6.3 mV). This shift is known to be associated with slow inactivation in Shaker channels and the disparity is reduced by 40 mV, or approximately 50% in the presence of 135 mM Cs. Off-gating charge immobilization is voltage-dependent with a V(1/2) of -12 mV, and correlates well with the development of Na(+) conductance on repolarization through C-type inactivated channels (V(1/2) is -11 mV). As well, the time-dependent development of the inward Na(+) tail current and gating charge immobilization after depolarizing pulses of different durations has the same time constant (tau = 2.7 ms). These results indicate that in Kv1.5 channels the transition to a stable C-type inactivated state takes only 2-3 ms and results in strong charge immobilization in the absence of Group IA metal cations, or even in the presence of Na. Inclusion of low concentrations of Cs delays the appearance of Na(+) tail currents in WT channels, prevents transition to inactivated states in Kv1.5-W472F nonconducting mutant channels, and removes charge immobilization. Higher concentrations of Cs are able to modulate the deactivating transition in Kv1.5 channels and prevent the residual slowing of charge return.  相似文献   

14.
Organic cation transporters of the OCT family mediate downhill transport of organic cations, compatible with carrier, pore, or gate-lumen-gate mechanisms. We studied rat OCT2 expressed in Xenopus oocytes by the two-electrode voltage-clamp technique, including membrane capacitance (C(m)) monitoring. Choline, a transported cationic substrate, elicited the expected inward currents but also elicited decreases of C(m). Similar C(m) decreases were caused by the non-transported inhibitors tetrabutylammonium (a cation) and corticosterone (uncharged). Effects on C(m) were voltage-dependent, with a maximum at -140 mV. These findings suggest that the empty rOCT2 protein can undergo an electrogenic conformation change, with one conformation highly favored at physiological voltage. Moreover, alkali cations elicited considerable inward currents and inhibited uptake of [(14)C]tetraethylammonium with a sequence Cs(+) > Rb(+) > K(+) > Na(+) approximately Li(+). Cs(+) affected current and capacitance with similar affinity (K(0.5) approximately 50 mm). Tetraethylammonium inhibited Cs(+) currents in a concentration-dependent manner. Conversely, Cs(+) inhibited tetraethylammonium uptake by a competitive mechanism. Activation energy of the currents estimated from measurements between 12 degrees C and 32 degrees C was approximately 81 kJ/mol for Cs(+) and 39 kJ/mol for tetramethylammonium, compatible with permeation of Cs(+) through rOCT2 along the same path as organic substrates and by a mechanism different from simple electrodiffusion. Rationalization of Cs(+) selectivity in terms of a pore pointed to a pore diameter of approximately 4 A. Intriguingly, that value matches the known selectivity of rOCT2 for organic compounds. Our data show that selective permeability of rOCT2 is not determined by ligand affinity but might rather be understood in terms of the ion channel concept of a distinct "selectivity filter."  相似文献   

15.
The thermal stability of the triplexes d(C(+)-T)(6):d(A-G)(6);d(C-T)(6) and d(T)(21):d(A)(21);d(T)(21) was studied in the presence of high concentrations of the anions Cl(-), HPO(4)(2-), CH(3)COO(-), SO(4)(2-) and ClO(4)(-). Thermally-induced triplex and duplex transitions were identified by UV- and CD-spectroscopy and T(m) values were determined from melting profiles. A thermodynamic analysis of triplex transitions shows the limitations of commonly used treatments for determining the associated release or uptake of salt, solute or water. Enhancement of the stability of these triplexes follows the rank order of the Hofmeister series for anions of sodium and ammonium salts, whereas water structure-breaking solutes have the opposite effect. The rank order for the Hofmeister series ClO(4)(-)相似文献   

16.
Liu W  Toney MD 《Biochemistry》2004,43(17):4998-5010
Dialkylglycine decarboxylase (DGD) is a tetrameric pyridoxal phosphate (PLP)-dependent enzyme that catalyzes both decarboxylation and transamination in its normal catalytic cycle. Its activity is dependent on cations. Metal-free DGD and DGD complexes with seven monovalent cations (Li(+), Na(+), K(+), Rb(+), Cs(+), NH(4)(+), and Tl(+)) and three divalent cations (Mg(2+), Ca(2+), and Ba(2+)) have been studied. The catalytic rate constants for cation-bound enzyme (ck(cat) and ck(cat)/bK(AIB)) are cation-size-dependent, K(+) being the monovalent cation with the optimal size for catalytic activity. The divalent alkaline earth cations (Mg(2+), Ca(2+), and Ba(2+)) all give approximately 10-fold lower activity compared to monovalent alkali cations of similar ionic radius. The Michaelis constant for aminoisobutyrate (AIB) binding to DGD-PLP complexes with cations (bK(AIB)) varies with ionic radius. The larger cations (K(+), Rb(+), Cs(+), NH(4)(+), and Tl(+)) give smaller bK(AIB) ( approximately 4 mM), while smaller cations (Li(+), Na(+)) give larger values (approximately 10 mM). Cation size and charge dependence is also found with the dissociation constant for PLP binding to DGD-cation complexes (aK(PLP)). K(+) and Rb(+) possess the optimal ionic radius, giving the lowest values of aK(PLP). The divalent alkaline earth cations give aK(PLP) values approximately 10-fold higher than alkali cations of similar ionic radius. The cation dissociation constant for DGD-PLP-AIB-cation complexes (betaK(M)z+) was determined and also shown to be cation-size-dependent, K(+) and Rb(+) yielding the lowest values. The kinetics of PLP association and dissociation from metal-free DGD and its complexes with cations (Na(+), K(+), and Ba(2+)) were analyzed. All three cations tested increase PLP association and decrease PLP dissociation rate constants. Kinetic studies of cation binding show saturation kinetics for the association reaction. The half-life for association with saturating Rb(+) is approximately 24 s, while the half-life for dissociation of Rb(+) from the DGD-PLP-AIB-Rb(+) complex is approximately 12 min.  相似文献   

17.
Although store-operated calcium release-activated Ca(2+) (CRAC) channels are highly Ca(2+)-selective under physiological ionic conditions, removal of extracellular divalent cations makes them freely permeable to monovalent cations. Several past studies have concluded that under these conditions CRAC channels conduct Na(+) and Cs(+) with a unitary conductance of approximately 40 pS, and that intracellular Mg(2+) modulates their activity and selectivity. These results have important implications for understanding ion permeation through CRAC channels and for screening potential CRAC channel genes. We find that the observed 40-pS channels are not CRAC channels, but are instead Mg(2+)-inhibited cation (MIC) channels that open as Mg(2+) is washed out of the cytosol. MIC channels differ from CRAC channels in several critical respects. Store depletion does not activate MIC channels, nor does store refilling deactivate them. Unlike CRAC channels, MIC channels are not blocked by SKF 96365, are not potentiated by low doses of 2-APB, and are less sensitive to block by high doses of the drug. By applying 8-10 mM intracellular Mg(2+) to inhibit MIC channels, we examined monovalent permeation through CRAC channels in isolation. A rapid switch from 20 mM Ca(2+) to divalent-free extracellular solution evokes Na(+) current through open CRAC channels (Na(+)-I(CRAC)) that is initially eightfold larger than the preceding Ca(2+) current and declines by approximately 80% over 20 s. Unlike MIC channels, CRAC channels are largely impermeable to Cs(+) (P(Cs)/P(Na) = 0.13 vs. 1.2 for MIC). Neither the decline in Na(+)-I(CRAC) nor its low Cs(+) permeability are affected by intracellular Mg(2+) (90 microM to 10 mM). Single openings of monovalent CRAC channels were not detectable in whole-cell recordings, but a unitary conductance of 0.2 pS was estimated from noise analysis. This new information about the selectivity, conductance, and regulation of CRAC channels forces a revision of the biophysical fingerprint of CRAC channels, and reveals intriguing similarities and differences in permeation mechanisms of voltage-gated and store-operated Ca(2+) channels.  相似文献   

18.
Das S  Reusch RN 《Biochemistry》2001,40(7):2075-2079
Poly-(R)-3-hydroxybutyrate/polyphosphate (PHB/polyP) complexes, whether isolated from the plasma membranes of bacteria or prepared from the synthetic polymers, form ion channels in planar lipid bilayers that are highly selective for Ca(2+) over Na(+) at physiological pH. This preference for divalent over monovalent cations is attributed to a high density of negative charge along the polyP backbone and the higher binding energies of divalent cations. Here we modify the charge density of polyP by varying the pH, and observe the effect on cation selectivity. PHB/polyP complexes, isolated from E. coli, were incorporated into planar lipid bilayers, and unitary current-voltage relations were determined as a function of pH. When Ca(2+) was the sole permeant cation, conductance diminished steadily from 97 +/- 6 pS at pH 7.4 to 47 +/- 3 pS at pH 5.5. However, in asymmetric solutions of Ca(2+) and Na(+), there was a moderate increase in conductance from 98 +/- 4 at pH 7.4 to 129 +/- 4 pS at pH 6.5, and a substantially larger increase to 178 +/- 6 pS at pH 5.6, signifying an increase in Na(+) permeability or disorganization of channel structure. Reversal potentials point to a sharp decrease in preference for Ca(2+) over Na(+) over a relatively small decrease in pH. Ca(2+) was strongly favored over Na(+) at physiological pH, but the channels became nonselective near the pK(2) of phosphate (approximately 6.8), and displayed weak selectivity for Na(+) over Ca(2+) at acidic pH. Evidently, PHB/polyP complexes are versatile ion carriers whose selectivity may be modulated by small adjustments of the local pH. The results may be relevant to the physiological function of PHB/polyP channels in bacteria and the role of PHB and polyP in the Streptomyces lividans potassium channel.  相似文献   

19.
The effects of monovalent cations (Li(+), Na(+), K(+), Rb(+), Cs(+), and NH4(+)) on the thermal stability of RNA tertiary structure were investigated by UV melting. We show that with the RNA used here (nucleotides 1051-1108 of Escherichia coli 23 S rRNA with four base substitutions), monovalent cations and Mg(2+) compete in stabilizing the RNA tertiary structure, and that the competition takes place between two boundaries: one where Mg(2+) concentration is zero and the other where it is maximally stabilizing ("saturating"). The pattern of competition is the same for all monovalent cations and depends on the cation's ability to displace Mg(2+) from the RNA, its ability to stabilize tertiary structure in the absence of Mg(2+), and its ability to stabilize tertiary structure at saturating Mg(2+) concentrations. The stabilizing ability of a monovalent cation depends on its unhydrated ionic radius, and at a low monovalent cation concentration and saturating Mg(2+), there is a (calculated) net release of a single monovalent cation/RNA molecule when tertiary structure is denatured. The implications are that under these conditions there is at least one binding site for monovalent cations on the RNA, the site is specifically associated with formation of stable tertiary structure, K(+) is the most effective of the tested cations, and Mg(2+) appears ineffective at this site. At high ionic strength, and in the absence of Mg(2+), stabilization of tertiary structure is still monovalent-cation specific and ionic-radius dependent, but a larger number of cations ( approximately eight) are released upon RNA tertiary structure denaturation, and NH(4)(+) appears to be the most effective cation in stabilizing tertiary structure under these conditions. In the majority of the experiments, methanol was added as a cosolvent to the buffer. Its use allowed the examination of the behavior of monovalent ions under conditions where their effects would otherwise have been too weak to be observed. Methanol stabilizes tertiary but not secondary structure of the RNA. There was no evidence that it either causes qualitative changes in cation-binding properties of the RNA or a change in the pattern of monovalent cation/Mg(2+) competition.  相似文献   

20.
G M Ananyev  A Murphy  Y Abe  G C Dismukes 《Biochemistry》1999,38(22):7200-7209
The size and charge density requirements for metal ion binding to the high-affinity Mn2+ site of the apo-water oxidizing complex (WOC) of spinach photosystem II (PSII) were studied by comparing the relative binding affinities of alkali metal cations, divalent metals (Mg2+, Ca2+, Mn2+, Sr2+), and the oxo-cation UO22+. Cation binding to the apo-WOC-PSII protein was measured by: (1) inhibition of the rate and yield of photoactivation, the light-induced recovery of O2 evolution by assembly of the functional Mn4Ca1Clx, core from its constituent inorganic cofactors (Mn2+, Ca2+, and Cl-); and by (2) inhibition of the PSII-mediated light-induced electron transfer from Mn2+ to an electron acceptor (DCIP). Together, these methods enable discrimination between inhibition at the high- and low-affinity Mn2+ sites and the Ca2+ site of the apo-WOC-PSII. Unexpectedly strong binding of large alkali cations (Cs+ > Rb+ > K+ > Na+ > Li+) was found to smoothly correlate with decreasing cation charge density, exhibiting one of the largest Cs+/Li+ selectivities (>/=5000) for any known chelator. Both photoactivation and electron-transfer measurements at selected Mn2+ and Ca2+ concentrations reveal that Cs+ binds to the high-affinity Mn2+ site with a slightly greater affinity (2-3-fold at pH 6.0) than Mn2+, while binding about 10(4)-fold more weakly to the Ca2+-specific site required for reassembly of functional O2 evolving centers. In contrast to Cs+, divalent cations larger than Mn2+ bind considerably more weakly to the high-affinity Mn2+ site (Mn2+ > Ca2+ > Sr2+). Their affinities correlate with the hydrolysis constant for formation of the metal hydroxide by hydrolysis of water: Me2+aq --> [MeOH]+aq + H+aq. Along with the strong stimulation of the rate of photoactivation by alkaline pH, these metal cation trends support the interpretation that [MnOH]+ is the active species that forms upon binding of Mn2+aq to apo-WOC. Further support for this interpretation is found by the unusually strong inhibition of Mn2+ photooxidation by the linear uranyl cation (UO22+). The intrinsic binding constant for [MnOH]+ to apo-WOC was determined using a thermodynamic cycle to be K = 4.0 x 10(15) M-1 (at pH 6.0), consistent with a high-affinity, preorganized, multidentate coordination site. We propose that the selectivity for binding [MnOH]+, a linear low charge-density monocation, vs symmetrical Me2+ dications is functionally important for assembly of the WOC by enabling: (1) discrimination against higher charge density alkaline earth cations (Mg2+ and Ca2+) and smaller alkali metal cations (Na+ and K+) that are present in considerably greater abundance in vivo, and thus would suppress photoactivation; and (2) higher affinity binding of the one Ca2+ ion or the remaining three Mn2+ ions via coordination to form mu-hydroxo-bridged intermediates, apo-WOC-[Mn(mu-OH)2Mn]3+ or apo-WOC-[Mn(mu-OH)Ca]3+, during subsequent assembly steps of the native Mn4Ca1Clx core. In contrast to more acidic Me2+ divalent ion inhibitors of the high-affinity Mn2+ site, like Ca2+ and Sr2+, Cs+ does not accelerate the decay of the first light-induced intermediate, IM1, formed during photoactivation (attributed to apo-WOC-[Mn(OH)2]+). The inability of Cs+ to promote decay of IM1, despite having comparable affinity as Mn2+, is consistent with its considerably weaker Lewis acidity, resulting in the reprotonation of IM1 by water becoming the rate-limiting step for decay prior to displacement of Mn2+. All four different lines of evidence provide a self-consistent picture indicating that the initial step in assembly of the WOC involves high-affinity binding of [MnOH]+.  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