首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Nucleosome core particle structure and structural changes in solution   总被引:1,自引:0,他引:1  
The radius of gyration, Rg, of chicken erythrocyte nucleosome core particles, was found to be 4.56 (+/- 0.07) nm by small-angle X-ray scattering, independent of particle concentration and of NaCl concentration between 0.1 M and 0.6 M-NaCl. The large, positive, second virial coefficient, A2, from particle concentration dependence (but independent of NaCl concentration) in small-angle X-ray scattering may indicate non-electrostatic repulsive ordering over large distances. Density contrast variation in equilibrium sedimentation with a small probe (sucrose) and a larger probe (gamma-cyclodextrin) yields good results for core particle hydration in the first instance, and for an estimate of the total particle volume in the second instance.  相似文献   

2.
The hydration and temperature dependencies of the 23Na+, 133Cs+, and 7Li+ quadrupolar splitting have been determined in hydrated, macroscopically oriented DNA fibers. At low water contents the quadrupolar splitting is found to decrease as the water content increases, regardless of counterion, while at high water contents the hydration dependence is reversed. The 23Na+ and 133Cs+ quadrupolar splittings decrease as the temperature increases, while the 7Li+ splitting shows the opposite behavior. At high water contents the 23Na+ and 133Cs+ splittings decrease, and then, after passing zero splitting, increase as the temperature increases. The interpretation of the temperature dependence is discussed in terms of a two-site model (free and bound ions) and a three-site model (free ions and specifically or nonspecifically bound ions). It is suggested that a three-site model is more consistent with the data for the present system. At high water contents, the temperature dependence of the 7Li+ splitting vanishes, indicating counterion condensation. The behavior of the 7Li+ splitting is confirmed by measurements on DNA fibers in equilibrium with a C2H5OD-D2O-LiCl solution. The salt dependence in this system is weak. The counterion quadrupolar splitting is seen to be very sensitive to structural transitions in double-helical DNA.  相似文献   

3.
R E Harrington 《Biochemistry》1985,24(8):2011-2021
Fractionated chicken erythrocyte chromatin fibers consisting of 10-mer and 75-mer polynucleosomes have been studied by flow birefringence and viscosity over a range of Na+ and Mg2+ ion concentrations sufficient to span the 10-30-nm fiber transition. Negative intrinsic flow bifringence was observed under all solvent conditions investigated. The intrinsic birefringence, obtained from the reduced birefringence to intrinsic viscosity ratio, was used to evaluate various optical models for the DNA conformation in the fiber. Results are consistent with an extended chromatosome-linker "necklace" model for the unfolded, low-salt fiber and with a solenoidal model of edge-stacked chromatosomes for the condensed fiber at high salts. These results are consistent with and independently corroborative of similar models based upon electric dichroism and neutron scattering reported by others.  相似文献   

4.
Structural transition in chromatin induced by ions in solution   总被引:4,自引:3,他引:1       下载免费PDF全文
Structural transition in chromatin was measured as a function of counter ions in solution (NaCl or MgCl(2)) and of histones bound on the DNA. The addition of counter ions to aqueous solutions of chromatin, partially dehistonized chromatin, and DNA caused a drastic reduction in viscosity and a significant increase in sedimentation coefficient. Transitions occurred primarily at about 2 x 10(-3) M NaCl and 1 x 10(-5) M MgCl(2) and are interpreted as a change in structure of chromatin induced by tight binding of cations (Na(+) or Mg(++)) to DNA, either free or bound by histones, and is an intrinsic property of DNA rather than of the type of histone bound. At a given ionic condition, removal of histone H1 from chromatin had only a minor effect on the hydrodynamic properties of chromatin while removal of other histones caused a drastic change in these properties. An increase in the sedimentation coefficient of DNA was observed also for protamine. DNA complexes wherein the bound protein contains only unordered coil rather than the alpha-helices found in histones.  相似文献   

5.
The tetrameric (H3/H4)2 146 base pair (bp) DNA and hexameric (H3/H4)2(H2A/H2B)1 146 bp DNA subnucleosomal particles have been prepared by depletion of chicken erythrocyte core particles using 3 or 4 M urea, 250 mM sodium chloride, and a cation-exchange resin. The particles have been characterized by cross-linking and sedimentation equilibrium. The structures of the particles, particularly the tetrameric, have been studied by sedimentation velocity, low-angle neutron scattering, circular dichroism, optical melting, and nuclease digestion with DNase I, micrococcal nuclease, and exonuclease III. It is concluded that since the radius of gyration of the DNA in the tetramer particle and its maximum dimension are very close to those of the core particle, no expansion occurs on removal of all the H2A and H2B. Nuclease digestion results indicate that histones H3/H4 in the tetramer particle protect a total of 70 bp of DNA that are centrally located within the 146 bp. Within the 70 bp DNA length, the two terminal regions of 10 bp are, however, not strongly protected from digestion. The optical melting profile of both particles can be resolved into three components and is consistent with the model of histone protection of DNA proposed from nuclease digestion. The structure proposed for the tetrameric histone complex bound to DNA is that of a compact particle containing 1.75 superhelical turns of DNA, in which the H3 and H4 histone location is the same as found for the core particle in chromatin by histone/DNA cross-linking [Shick, V. V., Belyavsky, A. V., Bavykin, S. G., & Mirzabekov, A. D. (1980) J. Mol. Biol. 139, 491-517]. Optical melting of the hexamer particle shows that each (H2A/H2B)1 dimer of the core particle protects about 22 base pairs of DNA.  相似文献   

6.
Using small angle neutron scattering we have measured the static form factor of two different superhelical DNAs, p1868 (1868 bp) and pUC18 (2686 bp), in dilute aqueous solution at salt concentrations between 0 and 1.5 M Na+ in 10 mM Tris at 0% and 100% D2O. For both DNA molecules, the theoretical static form factor was also calculated from an ensemble of Monte Carlo configurations generated by a previously described model. Simulated and measured form factors of both DNAs showed the same behavior between 10 and 100 mM salt concentration: An undulation in the scattering curve at a momentum transfer q = 0.5 nm-1 present at lower concentration disappears above 100 mM. The position of the undulation corresponds to a distance of approximately 10-20 nm. This indicated a change in the DNA superhelix diameter, as the undulation is not present in the scattering curve of the relaxed DNA. From the measured scattering curves of superhelical DNA we estimated the superhelix diameter as a function of Na+ concentration by a quantitative comparison with the scattering curve of relaxed DNA. The ratio of the scattering curves of superhelical and relaxed DNA is very similar to the form factor of a pair of point scatterers. We concluded that the distance of this pair corresponds to the interstrand separation in the superhelix. The computed superhelix diameter of 16.0 +/- 0.9 nm at 10 mM decreased to 9.0 +/- 0.7 nm at 100 mM salt concentration. Measured and simulated scattering curves agreed almost quantitatively, therefore we also calculated the superhelix diameter from the simulated conformations. It decreased from 18.0 +/- 1.5 nm at 10 mM to 9.4 +/- 1.5 nm at 100 mM salt concentration. This value did not significantly change to lower values at higher Na+ concentration, in agreement with results obtained by electron microscopy, scanning force microscopy imaging in aqueous solution, and recent MC simulations, but in contrast to the observation of a lateral collapse of the DNA superhelix as indicated by cryo-electron microscopy studies.  相似文献   

7.
Ionic effects on the structure of nucleoprotein cores from adenovirus   总被引:2,自引:0,他引:2  
Nucleoprotein cores, prepared from adenovirus type 5 with a deoxycholate/heat treatment, consist of the viral DNA and two major internal proteins. The core particles exhibit structural characteristics that are highly reproducible and dependent on their ionic environment. In low-ionic-strength buffer, the cores had a sedimentation coefficient of 180 S and appeared in the electron microscope as homogeneous particles with distinct centers from which numerous arms and loops radiated. Condensation of the cores was induced by Mg2+ or Ca2+ over the range 0 to 1 mM. The sedimentation coefficient increased monotonically with divalent cation concentration, reaching a maximum of 405 S in 1 mM Mg2+. A corresponding condensation in the core structure was observed by electron microscopy. Increasing concentrations of NaCl also produced a conformational change in the cores, with an almost linear increase in sedimentation velocity up to 274 S in 0.04 M NaCl. Between 0.05 and 1.0 M NaCl, the cores were insoluble. In 2.0 M NaCl, the cores were again soluble with an s20,w of 228 S. Under all ionic strength conditions in which the cores were soluble, both core proteins remained bound to the DNA.  相似文献   

8.
The electrophoretic mobilities of 24 single-stranded DNA oligomers, each containing 26 nucleotide residues, have been measured in polyacrylamide gels and in free solution. The mobilities observed at 20 degrees C differed by approximately 20% in polyacrylamide gels and by approximately 10% in free solution, even though the oligomers contained the same number of bases. Increasing the temperature or adding urea to the solution equalized the mobilities of the oligomers, suggesting that the variable mobilities observed at 20 degrees C are due to the formation of stable secondary structures, most likely hairpins. Thermal melting profiles were measured for eight oligomers in 40 mM Tris acetate buffer. The observed melting temperatures of most oligomers correlated roughly with the mobilities observed at 20 degrees C; however, one oligomer was much more stable than the others. The melting temperatures of four of the oligomers were close to the values predicted by DINAMelt [Markham, N. R., and Zuker, M. (2005) Nucleic Acids Res. 33, W577-W581]; melting temperatures of the other oligomers differed significantly from the predicted values. Thermal melting profiles were also measured for two oligomers as a function of the Tris acetate buffer concentration. The salt concentration dependence of the melting temperatures suggests that 0.15 Tris+ ion per phosphate is released upon denaturation. Because the apparent number of Tris+ ions released is greater than that observed by others for the release of Na+ ions from similar hairpins, the results suggest that DNA hairpins (and, presumably, duplexes) bind more Tris+ ions than Na+ ions in solution.  相似文献   

9.
Interactions of divalent polyamines with double-helical DNA in aqueous solution are investigated by monitoring the decrease in 23Na NMR relaxation rates as NaDNA is titrated with H3N(+)-(CH2)m-+NH3, where m = 3, 4, 5, or 6. Analogous measurements are made for the same homologous series of methylated polyamines (methonium ions). The dependence of the 23Na relaxation rates on the amount of added divalent cation (M2+) is analyzed quantitatively in terms of a two-state model. The sodium ions are assumed to be in rapid exchange between a "bound" state, where they are close enough to DNA so that it affects their relaxation rate, and a "free" state in bulk solution, where their relaxation rate is the same as in solutions containing no DNA. The distribution of Na+ and M2+ between these states is described quantitatively in terms of an ion-exchange parameter: DM = (pMB)(1-pNaB)n/(pNaB)n(1-pMB), where pNaB and pMB are the fractions of Na+ and M2+ that are close enough to DNA to be considered bound (by the NMR criterion), and n is the number of sodium ions displaced from DNA by the binding of one M2+ ion. For each of the polyamines and methonium ions investigated here, equations derived from this two-state model yield acceptable fittings of the titration curves if roNa, the number of sodium ions bound per DNA phosphate when no competing cations are present, is assigned a value between 0.6 and 1.00. Within this range, changing the value assigned to roNa does change the best-fitted values of DM determined for these polyamines (DH) and for the methonium ions (DMe) but does not alter the following conclusions about the trends in these parameters. (1) For polyamines and methonium ions of the same m, DH exceeds DMe by factors that are significantly larger for m = 3 and 4 than for m = 5 and 6. (2) DH for m = 3 and 4 is larger than DH for m = 5 and 6. (3) DMe for m = 3 and 4 is smaller than DMe for m = 5 and 6.  相似文献   

10.
R W Wilson  V A Bloomfield 《Biochemistry》1979,18(11):2192-2196
The addition of the trivalent or tetravalent cations spermidine or spermine to a solution of T7 DNA in aqueous solution causes an alteration of the DNA from its extended coil form to a condensed form. If performed at low DNA concentration and at low ionic strengths, this transformation results in a monomolecular collapse to form a particle with a hydrodynamic radius of about 500 A. We have monitored this change using quasielastic and total intensity light scattering. In a solution of 50% methanol in water, the divalent cations Mg2+ and putrescine also can cause the condensation of DNA. Using Manning's (1978) counterion condensation theory, we calculate a striking unity among these disparate ions: the collapse occurs in each case when from 89 to 90% of the DNA phosphate charges are neutralized by condensed counterions.  相似文献   

11.
Nanostructures of complexes of DNA with single-chain surfactant of octadecyltrimethylammonium (OTA) and double-chain surfactant of didodecyldimethylammonium (DDA) in aqueous NaCl solution at concentration, Cs, from 0 to 500 mM were studied using small-angle-scattering techniques (SAXS). SAXS profiles of the DNA-OTA complex show two SAXS peaks with a spacing ratio of 1:3(1/2) in the solution at Cs below 150 mM and three peaks with a spacing ratio of 1:3(1/2):4(1/2) at Cs above 250 mM. Contents of Na+ and Cl- ions in the complexes evaluated from the atomic absorbance for Na+ and the potentiometry for Cl- revealed charge molar ratios of OTA/DNA = 1 and DDA/DNA = 1.25. Contents of Na+ and Cl- ions per ionic unit of DNA molecule in the DNA-OTA complex equilibrating with the solution at Cs below 100 mM were much less than 0.1, while they increased with NaCl concentration at Cs above 200 mM. The DNA-OTA complex in the solution at Cs above 260 mM exhibited an endothermic peak in the DSC measurements, and the others did not. On the basis of the experimental results, the salt concentration dependent nanostructures are discussed.  相似文献   

12.
A variation of affinity capillary electrophoresis, called the replacement ion (RI) method, has been developed to measure the binding of monovalent cations to random sequence, double-stranded (ds) DNA. In this method, the ionic strength is kept constant by gradually replacing a non-binding ion in the solution with a binding ion and measuring the mobility of binding and non-binding analytes as a function of binding ion concentration. The method was validated by measuring the binding of Li+ ions to adenosine nucleotides; the apparent dissociation constants obtained by the RI method are comparable to literature values obtained by other methods. The binding of Tris+, NH4+, Li+, Na+, and K+ to dsDNA was then investigated. The apparent dissociation constants observed for counterion binding to a random-sequence 26-base pair (bp) oligomer ranged from 71 mM for Tris+ to 173 mM for Na+ and K+. Hence, positively charged Tris buffer ions will compete with other monovalent cations in Tris-buffered solutions. The bound cations identified in this study may correspond to the strongly correlated, tightly bound ions recently postulated to exist as a class of ions near the surface of dsDNA (Tan, Z.-J., and Chen, S.-J. (2006) Biophys. J. 91, 518-536). Monovalent cation binding to random-sequence dsDNA would be expected to occur in addition to any site-specific binding of cations to A-tracts or other DNA sequence motifs. Single-stranded DNA oligomers do not bind the five tested cations under the conditions investigated here.  相似文献   

13.
The conformation in solution of a small bipartite I-III hammerhead ribozyme has been deduced from the photo-crosslinks formed between cleavable ribo-deoxysubstrates appropriately substituted with the probe deoxy-4-thiouridine and ribozyme residues. The ribozyme-substrate complex is able to adopt a Y-like structure with stems I and II in close proximity in the presence of 400 mM Na+ only. Indeed, a cross-link joining stem I (1.6) to loop II (AL2.4) forms in significant amount under these conditions. This cross-linked complex furthermore elicits, upon Mg2+ addition, a catalytic activity similar to that exhibited by the complexes cross-linked at the distal ends of either stem I or stem III or of the non-substituted bipartite complex. This shows that the reaction mechanism is fully compatible with a strong structural constraint between stems I and II and that sodium ions at high concentration (400 mM) are able to promote a proper folding of hammerhead ribozymes. None of the multiple cross-links formed within the ribozyme core (probe in position 16.1 or 1.1) was found catalytically active. The cross-link patterns nevertheless indicate a higher flexibility of the core in Na+ than in Mg2+. While most of the cross-links can be accommodated by the Y solution structure, some of them (16.1 to U4 and 2.1) definitely can not, suggesting that additional alternative inactive conformations exist in solution.  相似文献   

14.
Nanosecond scale molecular dynamics simulations have been performed on antiparallel Greek key type d(G7) quadruplex structures with different coordinated ions, namely Na+ and K+ ion, water and Na+ counter ions, using the AMBER force field and Particle Mesh Ewald technique for electrostatic interactions. Antiparallel structures are stable during the simulation, with root mean square deviation values of approximately 1.5 A from the initial structures. Hydrogen bonding patterns within the G-tetrads depend on the nature of the coordinated ion, with the G-tetrad undergoing local structural variation to accommodate different cations. However, alternating syn-anti arrangement of bases along a chain as well as in a quartet is maintained through out the MD simulation. Coordinated Na+ ions, within the quadruplex cavity are quite mobile within the central channel and can even enter or exit from the quadruplex core, whereas coordinated K+ ions are quite immobile. MD studies at 400K indicate that K+ ion cannot come out from the quadruplex core without breaking the terminal G-tetrads. Smaller grooves in antiparallel structures are better binding sites for hydrated counter ions, while a string of hydrogen bonded water molecules are observed within both the small and large grooves. The hydration free energy for the K+ ion coordinated structure is more favourable than that for the Na+ ion coordinated antiparallel quadruplex structure.  相似文献   

15.
Neutron scattering studies of nucleosome structure at low ionic strength   总被引:1,自引:0,他引:1  
Ionic strength studies using homogeneous preparations of chicken erythrocyte nucleosomes containing either 146 or 175 base pairs of DNA show a single unfolding transition at about 1.5 mM ionic strength as determined by small-angle neutron scattering. The transition seen by some investigators at between 2.9 and 7.5 mM ionic strength is not observed by small-angle neutron scattering in either type of nucleosome particle. The two contrasts measured (H2O and D2O) indicate that only small conformational changes occur in the protein core, but the DNA is partially unfolded below the transition point. Patterson inversion of the data and analysis of models indicate that the DNA in both types of particle is unwinding from the ends, leaving about one turn of supercoiled DNA bound to the histone core in approximately its normal (compact) conformation. The mechanism of unfolding appears to be similar for both types of particles and in both cases occurs at the same ionic strength. The unfolding observed for nucleosomes in this study is in definite disagreement with extended superhelical models for the DNA and also disagrees with models incorporating an unfolded histone core.  相似文献   

16.
Using density gradient centrifugation and [3H]ryanodine as a specific marker, the ryanodine receptor-Ca2+ release channel complex from Chaps-solubilized canine cardiac sarcoplasmic reticulum (SR) has been purified in the form of an approximately 30 S complex, comprised of Mr approximately 400,000 polypeptides. Purification resulted in a specific activity of approximately 450 pmol bound ryanodine/mg of protein, a 60-70% recovery of ryanodine binding activity, and retention of the high affinity ryanodine binding site (KD = 3 nM). Negative stain electron microscopy revealed a 4-fold symmetric, four-leaf clover structure, which could fill a box approximately 30 x 30 nm and was thus morphologically similar to the SR-transverse-tubule, junctionally associated foot structure. The structural, sedimentation, and ryanodine binding data strongly suggest there is one high affinity ryanodine binding site/30 S complex, comprised of four Mr approximately 400,000 subunits. Upon reconstitution into planar lipid bilayers, the purified complex exhibited a Ca2+ conductance (70 pS in 50 mM Ca2+) similar to that of the native cardiac Ca2+ release channel (75 pS). The reconstituted complex was also found to conduct Na+ (550 pS in 500 mM Na+) and often to display complex Na+ subconducting states. The purified channel could be activated by micromolar Ca2+ or millimolar ATP, inhibited by millimolar Mg2+ or micromolar ruthenium red, and modified to a long-lived open subconducting state by ryanodine. The sedimentation, subunit composition, morphological, and ryanodine binding characteristics of the purified cardiac ryanodine receptor-Ca2+ release channel complex were similar to those previously described for the purified ryanodine receptor-Ca2+ release channel complex from fast-twitch skeletal muscle.  相似文献   

17.
A congenitally abnormal fibrinogen (Vlissingen) was isolated from the blood of a young woman suffering from massive pulmonary embolism. Fibrinogen Vlissingen showed an abnormal clotting time with both thrombin and Reptilase. The release of the fibrino-peptides A and B by thrombin was normal, but fibrin polymerization was impaired both in the presence and absence of Ca2+ ions. On sodium dodecyl sulfate-polyacrylamide gel electrophoresis performed according to Laemmli the gamma-chain of fibrinogen Vlissingen showed two bands, one normal and one having an apparently lower molecular mass of about 1,500 daltons. The previously described protective effect of Ca2+ ions on plasmin degradation of the carboxyl terminus of the gamma-chain of normal fibrinogen was only partially detectable in fibrinogen Vlissingen. In addition the binding of Ca2+ ions was decreased. Fibrinogen Vlissingen bound 2.4 Ca2+ ions per fibrinogen molecule at pH 7.4, whereas normal fibrinogen bound 3.1 Ca2+ ions. At pH 5.8 fibrinogen Vlissingen bound 1.1 Ca2+ ions, whereas normal fibrinogen bound 2.0 Ca2+ ions per molecule fibrinogen in the D-domains, again indicating a structural change in the carboxyl terminus of fibrinogen. The structural defect was determined by sequence analysis of DNA amplified by use of the polymerase chain reaction. Exons VIII, IX, and X of the gamma-chain gene were amplified and the DNA sequence of the amplified fragments was determined. A 6-base deletion was found in 50% of the fragments corresponding to exon VIII, indicating that the patient was heterozygous for the mutation. This deletion codes for amino acids Asn-319 and Asp-320 in the normal fibrinogen gamma-chain. The data indicate that Asn-319 and Asp-320 are crucial for maintaining the integrity of the carboxyl-terminal polymerization sites, the protective effect of Ca2+ ions on plasmin degradation of the carboxyl terminus of the gamma-chain, and the calcium binding domain at the carboxyl terminus of fibrinogen.  相似文献   

18.
23Na spin-lattice relaxation rate (1/T1 = R1) measurements have been used to study the intercalation of a series of 9-aminoacridine derivatives in DNA. The 23Na relaxation rate is strongly dependent upon the amount of intercalator added to a sodium DNA solution. The results are analyzed by a combined use of the ion condensation theory and the quadrupolar relaxation theory of polyelectrolyte solutions. This interpretation shows that the major effect in lowering the relaxation rate by intercalation is not due to the release of sodium ions but is caused by a substantial decrease in the relaxation rate Rb for the remaining bound sodium ions. Likewise, titration of NaDNA solutions with MgCl2 shows that condensation of Mg2+ on the DNA double helix reduces Rb. A good agreement between experiment and theory is found if the average lengthening following intercalation of a 9-aminoacridine moiety is assumed to be approximately 2.7 A. The distinction between mono- and bis-intercalation is clearly indicated by the results. The two bis-intercalating drugs examined are found to bis-intercalate only up to r less than or equal to 0.02. For r greater than 0.02 the drugs apparently mono-intercalate.  相似文献   

19.
23Na NMR relaxation rate measurements show that Na+ binds specifically to phosphatidylserine vesicles and is displaced partially from the binding site by K+ and Ca2+ but to a considerably less extent by tetraethylammonium ion. The data indicate that tetraethylammonium ion affects the binding of Na+ only slightly, by affecting the surface potential through its presence in the double layer, without competing for a phosphatidylserine binding site. Values for the intrinsic binding constant for the Na+-phosphatidylserine complex that would be consistent with the competition experiments (and the dependence of the relaxation rate on concentration of free Na+) fall in the range 0.4--1.2 M-1 with a better fit towards the higher values. We conclude that in the absence of competing cations in solution an appreciable fraction of the phosphatidylserine sites could be associated with bound Na+ at 0.1 M Na+ concentration.  相似文献   

20.
The objective of the present investigation was to characterize the ATP-dependent Na+-Na+ exchange, with respect to cation sensitivity on the two aspects of the Na+/K+-pump protein. In order to accomplish this, we used Na+/K+-ATPase reconstituted with known orientation in the proteoliposomes. Activation by cytoplasmic Na+ shows cooperative interaction between three sites. The apparent intrinsic site constants displayed transmembrane dependence on the extracellular Na+ concentration. However, the apparent K0.5 for cytoplasmic Na+ is independent of the extracellular Na+ concentration. The activation by extracellular Na+ at a fixed cytoplasmic Na+ concentration is biphasic with a component which saturates at a concentration of about 1-2 mM extracellular Na+, a plateau phase up to 20 mM, and another component which tends to saturate at about 80 mM followed by a slight deactivation at higher concentrations of Na+. The apparent K0.5 value for extracellular Na+ is also found to be independent of the Na+ concentration on the opposite side of the membrane. The activation by extracellular Na+ can be explained by the negative cooperativity in the binding of extracellular Na+, but positive cooperativity in the rate of dephosphorylation of enzyme species with one and three sodium ions bound extracellularly. Na+ bound to E2-PNa has a transmembrane effect on the cooperativity between binding of cytoplasmic Na+, and E2-PNa2 does not dephosphorylate. K0.5/Vm for cytoplasmic as well as for extracellular Na+ decreases with an increase in the trans Na+ concentration in the non-saturating concentration range. The experiments indicate that at a step in the reaction simultaneous binding of extracellular and cytoplasmic Na+ occurs.  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