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.
In this paper we introduce an important parameter called the iso-competition point (ICP), to characterize the competition binding to DNA in a two-cation-species system. By imposing the condition of charge neutralization fraction equivalence theta1 = ZthetaZ upon the two simultaneous equations in Manning's counterion condensation theory, the ICPs can be calculated. Each ICP, which refers to a particular multivalent concentration where the charge fraction on DNA neutralized from monovalent cations equals that from the multivalent cations, corresponds to a specific ionic strength condition. At fixed ionic strength, the total DNA charge neutralization fractions thetaICP are equal, no matter whether the higher valence cation is divalent, trivalent, or tetravalent. The ionic strength effect on ICP can be expressed by a semiquantitative equation as ICPZa/ICPZb = (Ia/Ib)Z, where Ia, Ib refers to the instance of ionic strengths and Z indicates the valence. The ICP can be used to interpret and characterize the ionic strength, valence, and DNA length effects on the counterion competition binding in a two-species system. Data from our previous investigations involving binding of Mg2+, Ca2+, and Co(NH3)63+ to lambda-DNA-HindIII fragments ranging from 2.0 to 23.1 kbp was used to investigate the applicability of ICP to describe counterion binding. It will be shown that the ICP parameter presents a prospective picture of the counterion competition binding to polyelectrolyte DNA under a specific ion environment condition.  相似文献   

2.
Activities of Mg and Ca in solutions of heparin with added sodium or potassium chloride have been determined by means of a previously described dye spectrophotometric method. The polyion concentration, Cp (expressed as mol univalent anionic charges/I) is 0.001 in all cases, solution total ionic strengths are 0.005, 0.0075, 0.01, 0.02 and 0.04 mol/l. Divalent metal ion concentrations are varied between 0 and 1.0 Cp. The results for the metal ion activities are expressed in the form of binding isotherms, θ2 vs. C2/Cp (θ2=C2b/Cp; C2b=bound divalent metal ion concentration) and Scatchard plots, K2 vs. θ2/(C2-C2b, at different ionic strengths. Results are correlated with a theoretical treatment of the mixed counterion system, called the ‘two variable theory’, recently developed by Manning. The ionic strength dependence of θ2 and K2 is found to agree with the prediction of the two variable theory. Mg interaction with heparin appears to be independent of the nature of the charged groups on the polyion and is in very good agreement with the two variable theory. Ca binding was found to be considerably higher than Mg binding and is in excess of theoretical predictions, suggesting a localized or specific interaction of Ca with heparin.  相似文献   

3.
The two-domain structure of streptokinase (Sk) was demonstrated by scanning calorimetric investigations at neutral pH and low ionic strength. The melting pattern of the protein is composed of two two-state transitions at TtrS1 = 45.9 +/- 0.4 degrees C with delta H1 = 431 +/- 18 kJ/mol, and TtrS2 = 60.1 +/- 1.3 degrees C with delta H2 = 306 +/- 16 kJ/mol. The partial specific heat capacity of native Sk was determined to be Cp = 1.42 +/- 0.17 J/K/g and the denaturational heat capacity change associated with the two transitions, delta Cp1 = 0.21 J/K/g and delta Cp2 = 0.38 J/K/g, respectively. The overall melting pattern of Sk remains almost unchanged at a variety of tested solvent compositions, except at pH 4 (and below) and in the presence of denaturants. The two domains show different susceptibility to urea. It is proposed that the less thermostable domain is located within the N-terminal part (residues 1-230), and the more thermostable one, within the C-terminal region.  相似文献   

4.
Polyuronates such as pectate and alginate are very well-known examples of biological polyelectrolytes undergoing, upon addition of divalent cations, an interchain association that acts as the junction of an eventually formed stable hydrogel. In the present paper, a thermodynamic model based on the counterion condensation theory has been developed to account for this cation-induced chain pairing of negatively charged polyelectrolytes. The strong interactions between cross-linking ions and uronate moieties in the specific binding site have been described in terms of chemical bonding, with complete charge annihilation between the two species. The chain-pairing process is depicted as progressively increasing with the concentration of cross-linking counterions and is thermodynamically defined by the fraction of each species. On these bases, the total Gibbs energy of the system has been expressed as the sum of the contributions of the Gibbs energy of the (single) chain stretches and of the (associated) dimers, weighted by their respective fractions 1 - theta and theta. In addition, the model assumes that the condensed divalent counterions exhibit an affinity free-energy for the chain, G(C)(aff,0), and the junction, G(D)(aff,0), respectively. Moreover, a specific Gibbs energy of chemical bonding, G(bond,0), has been introduced as the driving force for the formation of dimers. The model provides the mathematical formalism for calculating the fraction, theta, of chain dimers formed and the amount of ions condensed and bound onto the polyelectrolyte when two different types of counterions (of equal or different valence) are present. The effect of the parameter G(bond,0) has been investigated and, in particular, its difference from G(C,D)(aff,0) was found to be crucial in determining the distribution of the ions into territorial condensation and chemical bonding, respectively. Finally, the effect of the variation of the molar ratio between cross-linking ions and uronic groups in the specific binding sites, sigma0, was evaluated. In particular, a remarkable decrease in the amount of condensed counterions has been pointed out in the case of sigma0 = 1/3, with respect to the value of sigma0 = 1/4, characterizing the traditional "egg-box" structure, as a result of the drop of the charge density of the polyelectrolyte induced by complete charge annihilation.  相似文献   

5.
In order to test the diffuse ion atmosphere polarization model recently developed by us, the effects of ionic strength, titrating with Mg2+ and Co(NH3)3+6, and coion charge on the electric polarizability of short fragments of DNA are investigated. The results are consistent with the predictions of the theory and show that the diffuse ion atmosphere polarization contributes significantly to the overall orientation of DNA. At low ionic strengths, we attempt to separate the total dipole moment into two components: one that agrees well with the Debye-Hückel ion atmosphere calculations, while the other, presumably due to condensed counterion polarization, appears to be substantially independent of the ionic strength. At higher salt concentrations, however, a simple separation into dipole components is not possible, perhaps due to a significant coupling of ion flows between the diffuse atmosphere and the condensed counterion layer.  相似文献   

6.
Shim H  Raushel FM 《Biochemistry》2000,39(25):7357-7364
The active site of the bacterial phosphotriesterase (PTE) from Pseudomonas diminuta contains two divalent metal ions and a carboxylated lysine residue. The native enzyme contains two Zn(2+) ions, which can be replaced with Co(2+), Cd(2+), Ni(2+), or Mn(2+) without loss of catalytic activity. Carbon dioxide reacts with the side chain of lysine-169 to form a carbamate functional group within the active site, which then serves as a bridging ligand to the two metal ions. The activation of apo-PTE using variable concentrations of divalent metal ions and bicarbonate was measured in order to establish the mechanism by which the active site of PTE is self-assembled. The time courses for the activation of apo-PTE are pseudo-first-order, and the observed rate constants are directly proportional to the concentration of bicarbonate. In contrast, the apparent rate constants for the activation of apo-PTE decrease as the concentrations of the divalent cations are increased and then become constant at higher concentrations of the divalent metal ions. These results are consistent with a largely ordered kinetic mechanism for the assembly of the binuclear metal center where CO(2)/bicarbonate reacts with the apo-PTE prior to the binding of the two metal ions. When apo-PTE is titrated with 0-8 equiv of Co(2+), Cd(2+), or Zn(2+), the concentration of activated enzyme increases linearly until 2 equiv of metal ion is added and then remains constant at elevated levels of the divalent cations. These results are consistent with the synergistic binding of the two metal ions to the active site, and thus the second metal ion binds more tightly to the protein than does the first metal ion. Measurement of the mean dissociation constant indicates that metal binding to the binuclear metal center is strong [(K(alpha)K(beta))(1/2) = 6.0 x 10(-)(11) M and k(off) = 1.5 x 10(-)(3) min(-)(1) for Zn(2+)]. The removal of the carbamate bridge through the mutagenesis of Lys-169 demonstrates that the carbamate bridge is required for both efficient catalysis and overall stability of the metal center.  相似文献   

7.
The binding of Ca2+ to calmodulin and its two tryptic fragments has been studied using microcalorimetry. The binding process is accompanied by the uptake or release of protons, depending on the ionic strength. With no added salt, the total enthalpy change for the binding of four calcium ions to calmodulin is -41 kJ mol-1 but in the presence of 0.15 mM KCl delta Htot is +17 kJ mol-1. The mode of binding of Ca2+ is also completely different with and without added salt. It is also shown that for the C-terminal fragment of calmodulin, TR2C, the drastic reduction in delta Gtot for the binding process on increasing the ionic strength is largely an enthalpic effect. Domain interactions in calmodulin are indicated by the fact that the sum of the enthalpies of calcium binding to the two tryptic fragments is not the same as the total binding enthalpy to calmodulin itself. The binding of Ca2+ to calmodulin has also been studied calorimetrically at different temperatures in the range 21-37 degrees C. delta Cp is large and negative in this interval.  相似文献   

8.
The molecular theory of the previous paper in this series is extended to determine the effect of divalent metal ions on helix stability relative to coil at fixed ionic strength and nucleotide phosphate concentration. Specification of the state of condensed counterions, as well as their concentration, is essential for the solution of this problem, and it is assumed that they translate freely within a thin cylindrical shell close to the polynucleotide. As a function of divalent counterion concentration m the relative stability of the helix is highly nonlinear. Although the overall trend is that the helix stability increases with addition of divalent metal ion, there is a narrow concentration range for which it decreases slightly. The behavior of the relative stability as a function of m is determined by the translational degrees of freedom of the counterions, both univalent and divalent, both condensed and uncondensed. Detailed comparison of the theory with data is not given here, but it is pointed out that the calculated values of the relative stability are consistent with the order of magnitude of the observed effect Mg2+ on the melting temperature.  相似文献   

9.
By using single-channel recording techniques, we measured the conductance (gK) of the Ca(2+)-activated Maxi-K+ channel from the embryonic rat brain, and examined its dependence on K+ ions present in equimolar concentrations on both sides of the membrane patch. With ionic strength maintained constant by substitution of N-methyl-D-glucamine for K+, gK has a sigmoidal dependence upon [K+]. This result has been obscured in previous work by variations in ionic strength, which has a marked effect on single-channel conductance, especially in the limit for which this variable approaches zero. The gK versus [K+] relationship is described, theoretically, by a three-barrier, two-binding-site model in which the barrier that an ion must cross to leave the channel is decreased as [K+] is increased.  相似文献   

10.
The effect of ionic strength on the conformational equilibrium between the I(2) intermediate and the signaling state I(2)' of the photoreceptor PYP and on the rate of recovery to the dark state were investigated by time-resolved absorption and fluorescence spectroscopy. With increasing salt concentration up to approximately 600 mM, the recovery rate k(3) decreases and the I(2)/I(2)' equilibrium (K) shifts in the direction of I(2)'. At higher ionic strength both effects reverse. Experiments with mono-(KCl, NaBr) and divalent (MgCl(2), MgSO(4)) salts show that the low salt effect depends on the ionic strength and not on the cation or anion species. These observations can be described over the entire ionic strength range by considering the activity coefficients of an interdomain salt bridge. At low ionic strength the activity coefficient decreases due to counterion screening whereas at high ionic strength binding of water by the salt leads to an increase in the activity coefficient. From the initial slopes of the plots of log k(3) and log K versus the square root of the ionic strength, the product of the charges of the interacting groups was found to be -1.3 +/- 0.2, suggesting a monovalent ion pair. The conserved salt bridge K110/E12 connecting the beta-sheet of the PAS core and the N-terminal domain is a prime candidate for this ion pair. To test this hypothesis, the mutants K110A and E12A were prepared. In K110A the salt dependence of the I(2)/I(2)' equilibrium was eliminated and of the recovery rate was greatly reduced below approximately 600 mM. Moreover, at low salt the recovery rate was six times slower than in wild-type. In E12A significant salt dependence remained, which is attributed to the formation of a novel salt bridge between K110 and E9. At high salt reversal occurs in both mutants suggesting that salting out stabilizes the more compact I(2) structure. However, chaotropic anions like SCN shift the I(2)/I(2)' equilibrium toward the partially unfolded I(2)' form. The salt linkage K110/E12 stabilizes the photoreceptor in the inactive state in the dark and is broken in the light-induced formation of the signaling state, allowing the N-terminal domain to detach from the beta-scaffold PAS core.  相似文献   

11.
The human NAD-dependent isocitrate dehydrogenase (IDH), with three types of subunits present in the ratio of 2alpha:1beta:1gamma, requires a divalent metal ion to catalyze the oxidative decarboxylation of isocitrate. With the aim of identifying ligands of the enzyme-bound Mn(2+), we mutated aspartates on the alpha, beta, or gamma subunits. Mutagenesis target sites were based on crystal structures of metal-isocitrate complexes of Escherichia coli and pig mitochondrial NADP-IDH and sequence alignments. Aspartates replaced by asparagine or cysteine were 206, 230, and 234 of the alpha subunit and those corresponding to alpha-Asp-206: 217 of the beta subunit and 215 of the gamma subunit. Each expressed, purified mutant enzyme has two wild-type subunits and one subunit with a single mutation. Specific activities of WT, alpha-D206N, alpha-D230C, alpha-D234C, beta-D217N, and gamma-D215N enzymes are 22, 29, 1.4, 0.2, 7.3 and 3.7 micromol of NADH/min/mg, respectively, whereas alpha-D230N and alpha-D234N enzymes showed no activity. The K(m,Mn(2+)) for alpha-D230C and gamma-D215N are increased 32- and 100-fold, respectively, along with elevations in K(m,isocitrate). The K(m,NAD) of alpha-D230C is increased 16-fold, whereas that of beta-D217N is elevated 10-fold. For all the mutants K(m,isocitrate) is decreased by ADP, indicating that these aspartates are not needed for normal ADP activation. This study demonstrates that alpha-Asp-230 and alpha-Asp-234 are critical for catalytic activity, but alpha-Asp-206 is not needed; alpha-Asp-230 and gamma-Asp-215 may interact directly with the Mn(2+); and alpha-Asp-230 and beta-Asp-217 contribute to the affinity of the enzyme for NAD. These results suggest that the active sites of the human NAD-IDH are shared between alpha and gamma subunits and between alpha and beta subunits.  相似文献   

12.
Aggregation and fusion of unilamellar vesicles consisting of N-acyl-N-methylphosphatidylethanolamine were studied as a function of mono- and divalent cation concentrations. The aggregation reactions were irreversible processes, as demonstrated by changes in monovalent ion concentrations and by the addition of ethylenediaminetetraacetic acid (EDTA) to chelate divalent cations, suggesting the possibility of some cation-induced vesicle fusion. An increase in the NaCl ionic strength of the vesicle suspension solutions diminishes the threshold concentration for Li+ and K+ and increases that corresponding to Mn2+, Mg2+ and Ca2+. However NaCl concentrations above 300 mM yield smaller threshold values for the divalent cation-induced processes, probably due to the increased size of phospholipid vesicles as the ionic strength of the medium increases.  相似文献   

13.
Centrifuge transport, equilibrium dialysis, and electron paramagnetic resonance studies on the binding of Mn2+ to myosin revealed two sets of noninteracting binding sites which are characterized at low ionic strength (0.016 M KCl) by affinity constants of 10(6) M-1 (Class I) and 10(3) M-1 (Class II), respectively. At 0.6 M KCl concentration, the affinity of Mn2+ for both sets of sites is reduced. The maximum number of binding sites is 2 for the high affinity and 20 to 25 for the low affinity set. Other divalent metal ions displace Mn2+ from the high affinity sites in the following order of effectiveness: Ca greater than Mg = Zn = Co greater than Sr greater than Ni. The inhibitory effects of Mg2+ and Ca2+ upon the Mn2+ binding are competitive with inhibitor constants of 0.75 to 1 mM which is similar to that of the low affinity divalent metal ion binding sites. Exposure of myosin to 37 degrees partially inhibits Mn2+ binding to Class I parallel with inhibition of ATPase activity. The binding of Mn2+ to the high affinity binding sites is not significantly influenced by ADP or PPi, although Mn2+ increases the affinity of ADP binding to myosin at high ionic strength.  相似文献   

14.
The (ouabain + bumetanide + EGTA)-insensitive K+ influx (defined as residual K+ influx) in the human erythrocyte was investigated with respect to the characterization of the recently identified K+(Na+)/H+ exchanger (Richter et al. 1997). In particular, the effects of selected ion transport inhibitors on this flux in physiological ionic strength (high ionic strength, HIS) as well as low ionic strength (LIS) solutions were qstudied. The stimulation of the K+ influx observed in LIS medium was further enhanced when DIDS, phloretin, eosin-5-maleimide, furosemide, DIOA, NPPB, or DCDPC was present at a concentration of 0.1 mmol/l. This paradoxical, inhibitor-induced increase of the K+ influx was more pronounced in LIS media where chloride (7.5 mmol/l) was replaced by nitrate. For DNDS, niflumic acid, and MK-196 (0.1 mmol/l) an enhanced K+ transport could only be observed in nitrate-containing LIS solution. Bumetanide and purine riboside, at a concentration of 0.1 mmol/l, did not cause significant changes of the K+ influx in either chloride- or nitrate-containing LIS media. Dipyridamole and ruthenium red (0.1 mmol/l), which are positively charged, significantly reduced the K+ influx in both chloride- and nitrate-containing LIS media. In nitrate-containing HIS solution only dipyridamole inhibited the K+ influx. The residual K+ influx in LIS solution was significantly increased by removing internal [Mg2+], and decreased by quinacrine (1 mmol/l). In HIS solution, no effect of altering intracellular Mg2+ occurred but a stimulation of the flux by quinacrine was observed. The results are discussed in terms of a more general surface charge effect of the used inhibitors on the K+(Na+)/H+ exchanger.  相似文献   

15.
Differential scanning calorimetry, laser Raman spectroscopy, optical densitometry, and pH potentiometry have been used to investigate DNA melting profiles in the presence of the chloride salts of Ba2+, Sr2+, Mg2+, Ca2+, Mn2+, Co2+, Ni2+, and Cd2+. Metal-DNA interactions have been observed for the molar ratio [M2+]/[PO2-] = 0.6 in aqueous solutions containing 5% by weight of 160 bp mononucleosomal calf thymus DNA. All of the alkaline earth metals, plus Mn2+, elevate the melting temperature of DNA (Tm > 75.5 degrees C), whereas the transition metals Co2+, Ni2+, and Cd2+ lower Tm. Calorimetric (delta Hcal) and van't Hoff (delta HVH) enthalpies of melting range from 6.2-8.7 kcal/mol bp and 75.6-188.6 kcal/mol cooperative unit, respectively, and entropies from 17.5 to 24.7 cal/K mol bp. The average number of base pairs in a cooperative melting unit (<nmelt>) varied from 11.3 to 28.1. No dichotomy was observed between alkaline earth and transition DNA-metal complexes for any of the thermodynamic parameters other than their effects on Tm. These results complement Raman difference spectra, which reveal decreases in backbone order, base unstacking, distortion of glycosyl torsion angles, and rupture of hydrogen bonds, which occur after thermal denaturation. Raman difference spectroscopy shows that transition metals interact with the N7 atom of guanine in duplex DNA. A broader range of interaction sites with single-stranded DNA includes ionic phosphates, the N1 and N7 atoms of purines, and the N3 atom of pyrimidines. For alkaline earth metals, very little interaction was observed with duplex DNA, whereas spectra of single-stranded complexes are very similar to those of melted DNA without metal. However, difference spectra reveal some metal-specific perturbations at 1092 cm-1 (nPO2-), 1258 cm-1 (dC, dA), and 1668 cm-1 (nC==O, dNH2 dT, dG, dC). Increased spectral intensity could also be observed near 1335 cm-1 (dA, dG) for CaDNA. Optical densitometry, employed to detect DNA aggregation, reveals increased turbidity during the melting transition for all divalent DNA-metal complexes, except SrDNA and BaDNA. Turbidity was not observed for DNA in the absence of metal. A correlation was made between DNA melting, aggregation, and the ratio of Raman intensities I1335/I1374. At room temperature, DNA-metal interactions result in a pH drop of 1.2-2.2 units for alkaline earths and more than 2.5 units for transition metals. Sr2+, Ba2+, and Mg2+ cause protonated sites on the DNA to become thermally labile. These results lead to a model that describes DNA aggregation and denaturation during heating in the presence of divalent metal cations; 1) The cations initially interact with the DNA at phosphate and/or base sites, resulting in proton displacement. 2) A combination of metal-base interactions and heating disrupts the base pairing within the DNA duplex. This allows divalent metals and protons to bind to additional sites on the DNA bases during the aggregation/melting process. 3) Strands whose bases have swung open upon disruption are linked to neighboring strands by metal ion bridges. 4) Near the midpoint of the melting transition, thermal energy breaks up the aggregate. We have no evidence to indicate whether metal ion cross-bridges or direct base-base interactions rupture first. 5) Finally, all cross-links break, resulting in single-stranded DNA complexed with metal ions.  相似文献   

16.
The additivity rule of counterion activity or osmotic pressure in rodlike polyelectrolyte solutions has been discussed on the basis of the Fokker-Planck and Poisson equations in relation to the fluctuation of counterion distribution. This new theory has concluded that the additivity rule of counterion activity is less applicable than that of osmotic pressure due to the electric expansion force acting on the free-volume surface resulting from the fluctuation of counterion distribution. The theory has introduced an approximate relation between the counterion activities in the mixture solution of divalent and monovalent counterions, such that Deltaa+ = DeltaC++ - Deltaa++, in which Deltaa+ represents the increase of activity of monovalent counter-ions resulting from the addition of divalent counterionsDeltaC++, (in molar) to the solution, and Deltaa++ means the increase of the divalent counterion activity (in molar) in this process. This relation has been experimentally examined for Na-PSS solutions in the process of Cu2+ ion addition by the use of Na+ and Cu2+ sensitive electrodes, and it has been turned out that the relation is established in the low charge state of polyion.  相似文献   

17.
Measurements of thermal membrane potential across cation exchange membranes in MgCl2, CaCl2 and BaCl2 solutions and across anion exchange membranes in K2SO4, Na2SO4 and K2CO3 solutions were carried out. The magnitude of the thermal membrane potential for divalent counterions is lower than that for monovalent counterions. If the transport number of counterions in the membrane phase is unity, the slopes of the temperature coefficient of thermal membrane potential against logarithmic activities of counterion in the external solution are predicted to be--R/2F for 2-1 electrolytes with cation exchange membranes and R/2F for 1-2 electrolytes with anion exchange membranes, respectively.  相似文献   

18.
Magnetic resonance techniques have been applied to study the stability of the complexes formed between Mn(II) ions and NADP in aqueous solutions at a pH of 7.5 and 20 degrees C. The electron paramagnetic resonance (epr) data indicate that at low Mn(II) ion concentrations ([Mn(II)] less than 1 mM; [NADP] approximately 5 mM), a 1:1 complex is formed with an apparent stability constant K1 = 370 +/- 50 M-1 at an ionic strength of 0.22 in the presence of 0.20 M Cl-. At high Mn(II) ion concentrations, a Mn(II)2-NADP species, with an apparent stability constant K2 = 54 +/- 17 M-1, is present in significant amounts. When the epr data are corrected for the presence of the MnCl+ ion, the analysis of the new Scatchard plot yields stability constants for the two sites of K1 = 640 +/- 90 M-1 and K2 = 88 +/- 13 M-1, respectively. The presence of two metal ion binding sites on the NADP molecule has not been observed previously, and previous workers have always analyzed their data in terms of the 1:1 Mn(II)-NADP complex. An epr temperature study of K1 yields a value of delta H equal to 1.3 +/- 0.2 kcal/mol (1 cal = 4.187 J).  相似文献   

19.
Mannitol-1-phosphate dehydrogenase (EC 1.1.1.17) has been purified from Aspergillus parasiticus, a filamentous fungus which produces the polyketide mycotoxin, versicolorin A. Its kinetic properties have been compared with those of mannitol-1-phosphate dehydrogenase from the related non-toxin-producing fungus, A. niger. Both enzymes are inhibited by divalent transition metals, especially Zn2+ and Cd2+, but only the enzyme from A. parasiticus exhibits inhibitor-induced cooperative binding of the substrate, fructose-6 phosphate. Double reciprocal plots (1/v versus 1/Fru-6-P) are linear in the absence of Zn2+ but in the presence of Zn2+ are concave upward, with Hill coefficients of 1.5. The extent of cooperativity is inversely related to ionic strength, disappearing at 100 mM KCl. The enzymes from both organisms are relatively stable to incubation at 30 degrees C, but only the enzyme from A. parasiticus is rendered thermally unstable by the addition of divalent transition metals. A model is proposed to explain how binding of transition metal ions affects substrate binding and thermal stability of the enzyme.  相似文献   

20.
W H Beggs 《Life sciences》1979,24(8):663-668
Many independent studies have shown that inorganic salts can suppress the inhibitory activities of aminoglycoside antibiotics against various bacterial species. From studies on the nature and mechanism of this suppressive ion effect, several conclusions can be drawn. Both monovalent and divalent cation salts can inhibit aminoglycoside action. At least two different mechanisms are apparently involved. The first is divalent cation dependent and has been described with Pseudomonas aeruginosa and Mycobacterium smegmatis. With M. smegmatis, the effect aappears to be dependent more on a double positive charge than on particular metal ion species. The second mechanism is related to ionic strength and is elicited by both monovalent and divalent cation salts against many different organisms. However, based on studies with M. smegmatis, it appears that each of the two mechanisms involves interferance with the initial instantaneous electrostatic binding of drug at the cell surface.  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