首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The binding of the restriction endonuclease EcoRI to DNA is exceptionally specific. Even a single basepair change ("star" sequence) from the recognition sequence, GAATTC, decreases the binding free energy of EcoRI to values nearly indistinguishable from nonspecific binding. The difference in the number of waters sequestered by the protein-DNA complexes of the "star" sequences TAATTC and CAATTC and by the specific sequence complex determined from the dependence of binding free energy on water activity is also practically indistinguishable at low osmotic pressures from the 110 water molecules sequestered by nonspecific sequence complexes. Novel measurements of the dissociation rates of noncognate sequence complexes and competition equilibrium show that sequestered water can be removed from "star" sequence complexes by high osmotic pressure, but not from a nonspecific complex. By 5 Osm, the TAATTC "star" sequence complex has lost almost 90 of the approximately 110 waters initially present. It is more difficult to remove water from the CAATTC "star" sequence complex. The sequence dependence of water loss correlates with the known sequence dependence of "star" cleavage activity.  相似文献   

2.
The DNA binding stringency of restriction endonucleases is crucial for their proper function. The X-ray structures of the specific and non-cognate complexes of the restriction nuclease EcoRV are considerably different suggesting significant differences in the hydration and binding free energies. Nonetheless, the majority of studies performed at pH 7.5, optimal for enzymatic activity, have found a < 10-fold difference between EcoRV binding constants to the specific and nonspecific sequences in the absence of divalent ions. We used a recently developed self-cleavage assay to measure EcoRV-DNA competitive binding and to evaluate the influence of water activity, pH and salt concentration on the binding stringency of the enzyme in the absence of divalent ions. We find the enzyme can readily distinguish specific and nonspecific sequences. The relative specific-nonspecific binding constant increases strongly with increasing neutral solute concentration and with decreasing pH. The difference in number of associated waters between specific and nonspecific DNA-EcoRV complexes is consistent with the differences in the crystal structures. Despite the large pH dependence of the sequence specificity, the osmotic pressure dependence indicates little change in structure with pH. The large osmotic pressure dependence means that measurement of protein-DNA specificity in dilute solution cannot be directly applied to binding in the crowded environment of the cell. In addition to divalent ions, water activity and pH are key parameters that strongly modulate binding specificity of EcoRV.  相似文献   

3.
The molecular code of specific DNA recognition by proteins as a paradigm in molecular biology remains an unsolved puzzle primarily because of the subtle interplay between direct protein-DNA interaction and the indirect contribution from water and ions. Transformation of the nonspecific, low affinity complex to a specific, high affinity complex is accompanied by the release of interfacial water molecules. To provide insight into the conversion from the loose to the tight form, we characterized the structure and energetics of water at the protein-DNA interface of the BamHI complex with a noncognate sequence and in the specific complex. The fully hydrated models were produced with Grand Canonical Monte Carlo simulations. Proximity analysis shows that water distributions exhibit sequence dependent variations in both complexes and, in particular, in the noncognate complex they discriminate between the correct and the star site. Variations in water distributions control the number of water molecules released from a given sequence upon transformation from the loose to the tight complex as well as the local entropy contribution to the binding free energy. We propose that interfacial waters can serve as a "hydration fingerprint" of a given DNA sequence.  相似文献   

4.
The binding of a netropsin analogue to random sequence DNA, monitored by CD, is seen dependent on the concentration of neutral solutes. The binding free energy decreases linearly with solute osmolal concentration and the magnitude of the effect is insensitive to the chemical identity of the solute fur betaine, sorbitol, and triethylene glycol. These solutes appear to modulate binding through their effect on water activity and changes in the hydration of the drug and DNA in the complex reaction, not through a direct interaction with the reactants or the product. The dependence of binding constant on solute concentration can be interpreted as an additional binding of some 50–60 extra solute excluding water molecules by the complex. A water sensitivity of drug binding is further seen from the dependence of binding constants on the type of anion in solution. Anions in the Hofmeister series strongly affect bulk water free energies and entropies. The differences in netropsin analogue binding to DNA with Cl, F, and CIO are consistent with the effect observed with neutral solutes. The ability to measure changes in water binding associated with a specific DNA interaction is a first step toward correlating changes in hydration with the strength and specificity of binding. © 1995 John Wiley & Sons, Inc.  相似文献   

5.
Water release associated with specific binding of gal repressor.   总被引:7,自引:1,他引:6       下载免费PDF全文
M M Garner  D C Rau 《The EMBO journal》1995,14(6):1257-1263
Water release coupled to the association of gal repressor with DNA is measured from the sensitivity of the binding constant to the solution osmotic pressure, using neutral solutes that are typically excluded from polar protein and DNA surfaces. Differences in water release for binding of repressor to different sequences are linked with differences in specificity and binding energies. With sucrose, the specific binding of repressor to operator sequences is accompanied by the release of 130 water molecules. No water release is seen for the weak, non-specific binding of repressor to poly(dI-dC).(dI-dC). A difference in the release of six water molecules is seen even for the binding of gal repressor to two different operator sequences that differ in affinity by only a factor of two.  相似文献   

6.
In this report, the effects of osmotic pressure on BamHI cognate binding and catalysis were investigated and compared with a previous study on EcoRI (Robinson, C. R. and Sligar, S. G. (1998) Proc. Natl. Acad. Sci. U.S.A. 95, 2186-2191). Our observation of the dependence of binding and catalytic parameters on osmotic pressure has allowed for the comparison of hydration changes associated with site-specific DNA recognition for both endonucleases. Over a large range of osmotic pressures (pi), the dependence of BamHI on osmotic stress during cognate binding and catalysis was very different from that of the related endonuclease EcoRI. The binding of EcoRI to cognate DNA was dominated by a dehydration of the endonuclease-DNA complex, whereas binding by BamHI to its cognate sequence was accompanied by a solvent release corresponding to some 125 fewer waters. Catalytic analysis at elevated osmotic pressures indicated that both endonucleases had undergone a net hydration of the complex with BamHI displaying a much greater dependence on osmotic stress than EcoRI. Although the enzymes shared core structural motifs, comparisons of high resolution x-ray structures revealed many different secondary structural features of the complexed endonucleases. The large difference in hydration changes by both BamHI and EcoRI could be attributed to these dissimilar secondary structural features, as well as the functional differences of the two endonucleases during site-specific DNA recognition.  相似文献   

7.
The osmotic stress technique was used to measure changes in macromolecular hydration that accompany binding of wild-type Escherichia coli lactose (lac) repressor to its regulatory site (operator O1) in the lac promoter and its transfer from site O1 to nonspecific DNA. Binding at O1 is accompanied by the net release of 260 +/- 32 water molecules. If all are released from macromolecular surfaces, this result is consistent with a net reduction of solvent-accessible surface area of 2370 +/- 550 A. This area is only slightly smaller than the macromolecular interface calculated for a crystalline repressor dimer-O1 complex but is significantly smaller than that for the corresponding complex with the symmetrical optimized O(sym) operator. The transfer of repressor from site O1 to nonspecific DNA is accompanied by the net uptake of 93 +/- 10 water molecules. Together these results imply that formation of a nonspecific complex is accompanied by the net release of 165 +/- 43 water molecules. The enhanced stabilities of repressor-DNA complexes with increasing osmolality may contribute to the ability of Escherichia coli cells to tolerate dehydration and/or high external salt concentrations.  相似文献   

8.
We use the osmotic pressure dependence of dissociation rates and relative binding constants to infer differences in sequestered water among complexes of lambda Cro repressor with varied DNA recognition sequences. For over a 1000-fold change in association constant, the number of water molecules sequestered by non-cognate complexes varies linearly with binding free energy. One extra bound water molecule is coupled with the loss of approximately 150 cal/mol complex in binding free energy. Equivalently, every tenfold decrease in binding constant at constant salt and temperature is associated with eight to nine additional water molecules sequestered in the non-cognate complex. The relative insensitivity of the difference in water molecules to the nature of the osmolyte used to probe the reaction suggests that the water is sterically sequestered. If the previously measured changes in heat capacity for lambda Cro binding to different non-cognate sequences are attributed solely to this change in water, then the heat capacity change per incorporated water is almost the same as the difference between ice and water. The associated changes in enthalpies and entropies, however, indicate that the change in complex structure involves more than a simple incorporation of fixed water molecules that act as adaptors between non-complementary surfaces.  相似文献   

9.
It is well established that salt enhances the interaction between solutes (e.g., proteins, displacers) and the weak hydrophobic ligands in hydrophobic interaction chromatography (HIC) and that various salts (e.g., kosmotropes, chaotropes, and neutral) have different effects on protein retention. In this article, the solute affinity in kosmotropic, chaotropic, and neutral mobile phases are compared and the selectivity of solutes in the presence of these salts is examined. Since solute binding in HIC systems is driven by the release of water molecules, the total number of released water molecules in the presence of various types of salts was calculated using the preferential interaction theory. Chromatographic retention times and selectivity reversals of both proteins and displacers were found to be consistent with the total number of released water molecules. Finally, the solute surface hydrophobicity was also found to have a significant effect on its retention in HIC systems.  相似文献   

10.
Turgor maintenance, solute content and recovery from water stress were examined in the drought-tolerant shrub Artemisia tridentata. Predawn water potentials of shrubs receiving supplemental water remained above ?2 MPa throughout summer, while predawn water potentials of untreated shrubs decreased to ?5 MPa. Osmotic potentials decreased in conjunction with water potentials maintaining turgor pressures above 0 MPa. The decreases in osmotic potentials were not the result of osmotic adjustment (i.e. solute accumulation). Leaf solute contents decreased during drought, but leaf water volumes decreased more than 75% from spring to summer, thereby passively concentrating solutes within the leaves. The maintenance of positive turgor pressures despite decreases in leaf water volumes is consistent with other studies of species with elastic cell walls. Inorganic ion, organic acid, and carbohydrate contents of leaves declined during drought. The only solutes accumulating in leaves of A. tridentata with water stress were proline and a cyclitol, both considered compatible solutes. Total and osmotic potentials recovered rapidly following rewatering of shrubs; solute contents did not change except for a decrease in proline. Maintaining turgor through the passive concentration of solutes may be advantageous compared to synthesis of new solutes for osmotic adjustment in arid environments.  相似文献   

11.
Phospholipid vesicles prepared by the freeze-thaw extrusion method contain internal solute concentrations which are much higher than the external values (entrapment ratios much greater than 1). This concentrating effect is a complex function of the total impermeant solute concentration in the medium used to prepare vesicles, the presence or absence of permeant solutes in the medium and the apparent competitive binding interactions between solutes and phospholipid. Increases in water phase solute concentration during freezing are thought to underlie the concentrating phenomenon, while osmotic pressure driven lysis of vesicles during thawing appears to limit its magnitude. By judicious selection of solute concentration and physical properties, further increases in the entrapment ratio should be obtainable, improving the usefulness of these vesicles as drug delivery vesicles and experimental systems.  相似文献   

12.
The Kedem-Katchalsky (KK) equations are often used to obtain information about the osmotic properties and conductance of channels to water. Using human red cell membranes, in which the osmotic flow is dominated by Aquaporin-1, we show here that compared to NaCl the reflexion coefficient of the channel for methylurea, when corrected for solute volume exchange and for the water permeability of the lipid membrane, is 0.54. The channels are impermeable to these two solutes which would seem to rule out flow interaction and require a reflexion coefficient close to 1.0 for both. Thus, two solutes can give very different osmotic flow rates through a semi-permeable pore, a result at variance with both classical theory and the KK formulation. The use of KK equations to analyze osmotic volume changes, which results in a single hybrid reflexion coefficient for each solute, may explain the discrepancy in the literature between such results and those where the equations have not been employed. Osmotic reflexion coefficients substantially different from 1.0 cannot be ascribed to the participation of other 'hidden' parallel aqueous channels consistently with known properties of the membrane. Furthermore, we show that this difference cannot be due to second-order effects, such as a solute-specific interaction with water in only part of the channel, because the osmosis is linear with driving force down to zero solute concentration, a finding which also rules out the involvement of unstirred-layer effects. Reflexion coefficients smaller than 1.0 do not necessitate water-solute flow interaction in permeable aqueous channels; rather, the osmotic behaviour of impermeable molecular-sized pores can be explained by differences in the fundamental nature of water flow in regions either accessible or inaccessible to solute, created by a varying cross-section of the channel.  相似文献   

13.
R. F. Meyer  J. S. Boyer 《Planta》1981,151(5):482-489
Soybean (Glycine max (L.) Merr.) seedlings osmoregulate when the supply of water is limited around the roots. The osmoregulation involves solute accumulation (osmotic adjustment) by the elongating region of the hypocotyls. We investigated the relationship between growth, solute accumulation, and the partitioning of solutes during osmoregulation. Darkgrown seedlings were transplanted to vermiculite containing 1/8 (0.13 x) the water of the controls. Within 12–15 h, the osmotic potential of the elongating region had decreased to-12 bar, but it was-7 bar in the controls. This osmoregulation involved a true solute accumulation by the hypocotyls, since cell volume and turgor were virtually the same regardless of the water regime. The hypocotyls having low water potentials elongated slowly but, when deprived of their cotyledons, did not elongate or accumulate solute. This result indicated a cotyledonary origin for the solutes and a dependence of slow growth on osmotic adjustment. The translocation of nonrespired dry matter from the cotyledons to the seedling axis was unaffected by the availability of water, but partitioning was altered. In the first 12 h, dry matter accumulated in the elongating region of the 0.13 x hypocotyls, and osmotic adjustment occurred. The solutes involved were mostly free amino acids, glucose, fructose, and sucrose, and these accounted for most of the increased dry weight. After osmotic adjustment was complete, dry matter ceased to accumulate in the hypocotyls and bypassed them to accumulate in the roots, which grew faster than the control roots. The proliferation of the roots resulted in an increased root/shoot ratio, a common response of plants to dry conditions.Osmotic adjustment occurred in the elongating region of the hypocotyls because solute utilization for growth decreased while solute uptake continued. Adjustment was completed when solute uptake subsequently decreased, and uptake then balanced utilization. The control of osmotic adjustment was therefore the rate of solute utilization and, secondarily, the rate of solute uptake. Elongation was inhibited by unknown factors(s) despite the turgor and substrates associated with osmotic adjustment. The remaining slow elongation depended on osmotic adjustment and represented some optimum between the necessary inhibition for solute accumulation and the necessary growth for seedling establishment.  相似文献   

14.
Eggers DK 《Biochemistry》2011,50(12):2004-2012
A new phenomenological model for interpreting the effects of solutes on biological equilibria is presented. The model attributes changes in equilibria to differences in the desolvation energy of the reacting species that, in turn, reflect changes in the free energy of the bulk water upon addition of secondary solutes. The desolvation approach differs notably from that of other solute models by treating the free energy of bulk water as a variable and by not ascribing the observed shifts in reaction equilibria to accumulation or depletion of solutes next to the surfaces of the reacting species. On the contrary, the partitioning of solutes is viewed as a manifestation of the different subpopulations of water that arise in response to the surface boundary conditions. A thermodynamic framework consistent with the proposed model is used to derive a relationship for a specific reaction, an aqueous solubility equilibrium, in two or more solutions. The resulting equation reconciles some potential issues with the transfer free energy model of Tanford. Application of the desolvation energy model to the analysis of a two-state protein folding equilibrium is discussed and contrasted to the application of two other solute models developed by Timasheff and by Parsegian. Future tabulation of solvation energies and bulk water energies may allow biophysical chemists to confirm the mechanism by which secondary solutes influence binding and conformational equilibria and may provide a common ground on which experimentalists and theoreticians can compare and evaluate their results.  相似文献   

15.
Guard cell solution osmotic potential changes resulting in the opening and closing of stomata apertures follow an initial influx of potassium ions, their substitution with sucrose molecules and the subsequent reduction of the latter. To provide an insight into the osmotic mechanism of the changes, the new equation for calculating osmotic pressure, which equates the difference between the energy of pure water across a semi-permeable membrane interface with that of solution water, was used to compare the osmotic properties of KCl and sucrose. For sucrose solutions, the effect of the sucrose molecules in increasing the spacing of the solution water was mainly responsible for osmotic potential; this contrasted with K+ + Cl? ions where their spacing effect was only a little higher to that of water held to those ions. At solute concentrations giving an osmotic potential level of ?3.0 MPa near that of turgid guard cells, the spacing effect on the potential of the unattached solution water molecules caused by sucrose, but in its theoretical absence, was estimated as ?2.203 MPa compared with ?1.431 MPa for KCl. In contrast, the potential attributed to water molecules firmly held to the K+ + Cl? ions was ?1.212 MPa versus zero for sucrose. The potential to keep the sucrose molecules in solution was ?0.797 MPa compared with ?0.357 MPa for KCl. The findings illustrate that the way KCl effects osmotic pressure is very different to that of sucrose. It is concluded that stomata aperture modulation is closely linked to the osmotic properties of its guard cell solution solutes.  相似文献   

16.
The molecular mechanism by which DNA-binding proteins find their specific binding sites is still unclear. To gain insights into structural and energetic elements of this mechanism, we used the crystal structure of the nonspecific BamHI-DNA complex as a template to study the dominant electrostatic interaction in the nonspecific association of protein with DNA, and the possible sliding pathways that could be sustained by such an interaction. Based on calculations using the nonlinear Poisson-Boltzmann method and Brownian dynamics, a model is proposed for the initial nonspecific binding of BamHI to B-form DNA that differs from that seen in the crystal structure of the nonspecific complex. The model is electrostatically favorable, and the salt dependence as well as other thermodynamic parameters calculated for this model are in good agreement with experimental results. Several residues in BamHI are identified for their important contribution to the energy in the nonspecific binding model, and specific mutagenesis experiments are proposed to test the model on this basis. We show that a favorable sliding pathway of the protein along DNA is helical.  相似文献   

17.
Fluid secretion and reabsorption by a variety of plant and animal tissues appear to be accomplished by osmotic coupling between solute transport and water movement. The local osmosis model suggests that active accumulation of solutes within narrow folds at the cell surface may produce the local gradients that generate water flow. Both micropuncture techniques and electron-probe X-ray microanalysis have established that local osmotic gradients occur in absorptive epithelia, but they have not as yet been detected in secretory tissues.Hormonal control of secretion involves stimulation of solute pumps and adjustments of permeability to non-transported solutes. Since hormone receptors and pumps are often located on opposite surfaces of the cell, intracellular second messengers convey the secretory signal through cytoplasm. Much has been learned by study of insect tissues that are anatomically simple and that function for long periods in vitro. Aspects of hormone-receptor interaction have been explored, including the action of halluninogenic molecules. In insect salivary glands cyclic AMP appears to stimulate cation transport, while calcium increases anion permeability. The various second messengers probably interact with each other in complex feedback loops that stabilize the system and make it quickly responsive to hormone. Cyclic AMP may stimulate release of calcium from mitochondria. Unresolved is the way second messengers alter properties of the cell surface.  相似文献   

18.
We recently showed that a nonspecific complex of the restriction nuclease EcoRI with poly (dI-dC) sequesters significantly more water at the protein-DNA interface than the complex with the specific recognition sequence. The nonspecific complex seems to retain almost a full hydration layer at the interface. We now find that at low osmotic pressures a complex of the restriction nuclease EcoRI with a DNA sequence that differs by only one base pair from the recognition site (a 'star' sequence) sequesters about 70 waters more than the specific one, a value virtually indistinguishable from nonspecific DNA. Unlike complexes with oligo (dI-dC) or with a sequence that differs by two base pairs from the recognition sequence, however, much of the water in the 'star' sequence complex is removed at high osmotic pressures. The energy of removing this water can be calculated simply from the osmotic pressure work done on the complex. The ability to measure not only the changes in water sequestered by DNA-protein complexes for different sequences, but also the work necessary to remove this water is a potentially powerful new tool for coupling inferred structural changes and thermodynamics.  相似文献   

19.
Thellungiella salsuginea, a Brassicaceae species closely related to Arabidopsis thaliana, is tolerant to high salinity. The two species were compared under conditions of osmotic stress to assess the relationships between stress tolerance, the metabolome, water homeostasis and growth performance. A broad range of metabolites were analysed by metabolic fingerprinting and profiling, and the results showed that, despite a few notable differences in raffinose and secondary metabolites, the same metabolic pathways were regulated by salt stress in both species. The main difference was quantitative: Thellungiella had much higher levels of most metabolites than Arabidopsis whatever the treatment. Comprehensive quantification of organic and mineral solutes showed a relative stability of the total solute content regardless of the species or treatment, meaning that little or no osmotic adjustment occurred under stress. The reduction in osmotic potential observed in plants under stress was found to result from a passive loss of water. Thellungiella shoots contain less water than Arabidopsis shoots, and have the ability to lose more water, which could contribute to maintain a water potential gradient between soil and plant. Significant differences between Thellungiella and Arabidopsis were also observed in terms of the physicochemical properties of their metabolomes, such as water solubility and polarity. On the whole, the Thellungiella metabolome appears to be more compatible with dehydration. Osmotic stress was also found to impact the metabolome properties in both species, increasing the overall polarity. Together, the results suggest that Thellungiella copes with osmotic stress by tolerating dehydration, with its metabolic configuration lending itself to osmoprotective strategies rather than osmo-adjustment.  相似文献   

20.
Escherichia coli and Lactobacillus plantarum were subjected to final water potentials of −5.6 MPa and −11.5 MPa with three solutes: glycerol, sorbitol and NaCl. The water potential decrease was realized either rapidly (osmotic shock) or slowly (20 min) and a difference in cell viability between these conditions was only observed when the solute was NaCl. The cell mortality during osmotic shocks induced by NaCl cannot be explained by a critical volume decrease or by the intensity of the water flow across the cell membrane. When the osmotic stress is realized with NaCl as the solute, in a medium in which osmoregulation cannot take place, the application of a slow decrease in water potential resulted in the significant maintenance of cell viability (about 70–90%) with regard to the corresponding viability observed after a sudden step change to same final water potential (14–40%). This viability difference can be explained by the existence of a critical internal free Na+ concentration. Received: 20 May 1998 / Received revision: 31 July 1998 / Accepted: 31 July 1998  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