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.
Complex formation between poly(U) and adenosine in solutions of salts that stabilize (Na2SO4), destabilize (NaClO4), or have little effect on the water structure (NaCl), as well as the poly(U)·poly(A) interaction in NaClO4, was studied by equilibrium dialysis and uv spectroscopy. At 3°C and neutral pH, Ado·2 poly(U) is formed in 1M NaCl and 0.33M Na2SO4. In NaClO4 solutions under the same conditions, an Ado·poly(U) was found over the whole range of salt concentration investigated (10 mM?1M), which has not been previously observed under any conditions. The Ado-poly(U) was also found in a NaCl/NaClO4 mixture, the transition from the triple- to the double-helical complex occurring within a narrow range of concentration of added NaClO4. In the presence of 1M NaCl this transition is observed on adding as little as 10 mM NaClO4, i.e., at a [ClO]/[Cl?] ratio of about 1:100. However, when NaClO4 is added to a 1M solution of the stabilizing salt Na2SO4, no transition occurs even at a [ClO]/[SO] ratio of 1:4. Investigation of melting curves and uv spectra has shown that in an equimolar mixture of the polynucleotides, only a double-helical poly(U)·poly(A) exists in 1M NaClO4 at low temperatures; this also holds for 1M NaCl. This changes to a triple-helical 2 poly(U)·poly(A) and then dissociates as the temperature increases. At low temperatures and the poly(U)/poly(A) concentration ratio of 2:1, a mixture of 2 poly(U)·poly(A) and poly(U)·poly(A) was observed in 1M NaClO4, in contrast to the case of 1M NaCl. Thus, sodium perchlorate, a strong destabilizer of water structure, promotes formation of double-helical complexes both in the polynucleotide–monomer and the polynucleotide–polynucleotide systems. Beginning with a sufficiently high ionic strength (μ ? 0.9), a further increase in the salt molarity results in an increase of the poly(U)·adenosine melting temperature in both stabilizing and neutral salts and a decrease in the destabilizing salt. In Na2SO4 concentrations higher than 1.2M Ado·2 poly(U) precipitates at room temperature. Analysis of the binding isotherms and melting profiles of the complexes between poly(U) and adenosine according to Hill's model shows that the cooperativity of binding, due to adenosine stacking on poly(U), increases in the order NaClO4 < NaCl < Na2SO4. The free energy of adenosine stacking on the template is similar to that of hydrogen bonding between adenosine and poly(U) and ranges from ?1 to ?2 kcal/mol. The values of ΔHt [the effective enthalpy of adenosine binding to poly(U) next to an occupied site, obtained from the relationship between complex melting temperature and free monomer concentration at the midpoint of the transition] are ?14.2, ?18.3, and ?16.8 kcal/mol for 1M solutions of NaClO4, NaCl, and Na2SO4, respectively. The results indicate that the effects of anions of the salts studied are related to water structure alterations rather than to their direct interaction with the complexes between poly(U) and adenosine.  相似文献   

2.
Complexes of poly(BU) with various adenine derivatives were investigated by circular dichroism (CD) and absorption spectroscopy. A 1:2 stoichiometry was indicated on CD mixing curves for typical complexes of 9-substituted adenine and 2-aminoadenine derivatives with poly(BU). The CD spectrum of adenosine·2poly(BU) is characterized by well-resolved bands in the range of 210–350 nm. Other adenine derivative–poly(BU) complexes also afford similar CD spectra, while 2-aminoadenine derivative–poly(BU) complexes give quite different spectra. Attempts to assign representative CD spectra were made using the transition of helical poly(BU) and the respective purine polynucleotides. The similarity of the CD spectra suggests that poly(A)·2poly(BU) and adenine derivative–poly(BU) complexes are nearly identical in structure except for the ribose–phosphate linkage. The fact that the uv isosbestic point of adenosine·2poly(BU) falls in close proximity to that of the corresponding polymer complex also supports this conclusion. In the formation of stable helices, the ribose moiety is dispensable in the “strand” of purine. The Tm of 9-methyladenine·2poly(BU) is somewhat higher than that of adenosine·2poly(BU) under equivalent conditions. The Tm difference with the monomer–poly(U) system was found to be about 20°C in 0.4M NaCl–0.02M Na–cacodylate–5 × 10?4M EDTA (pH 7.0). Further, it was noted that the monomer–poly(BU) complexes are formed even when the Tm is lower than that of self-folded poly(BU).  相似文献   

3.
Double-helical poly(dG-dC) and poly(dA-dT) are DNA analogs in which the interactions between the two strands of the helix are, respectively, either the stronger G/C type or the weaker A/T type along the entire length of macromolecules. Thus, these synthetic polynucleotides can be considered as representatives of the most stable and the least stable DNA. In the investigations presented here, potentiometric titrations and stopped-flow kinetic experiments were carried out in order to compare the pH-induced helix–coil conformations (10°C and 150mM [Na+]) the pH of the helix–coil transition (pHm) is 12.81 for poly(dG-dC) and 11.76 for poly(dA-dT). The unwinding of double-helical poly(dG-dC) initiated by a sudden change in pH was found to be a simple exponential process with rate constants in the range of 200–600 sec?1, depending on the final value of the pH jump. The intramolecular double-helix formation of poly(dG-dC) was studied by lowering the pH of the solutions from a value above pHm to that below pHm in dilute solutions (15.5 ug/ml [polymer]). Under these conditions, the observed rewinding reactions displayed a major and two exponential phases, all of which were independent of polymer concentration. From the comparison of the results of poly(dA-dT) and poly(dG-dT) would unwind faster than poly(dG-dC). However, if the pH jumps are such that they present the same perturbation of these polymers relative to their pHm values, no significant differences exist between the rates of helix–coil conformation changes of poly(dA-dT) and poly(dG-dC).  相似文献   

4.
T E Gunter  K K Gunter 《Biopolymers》1972,11(3):667-678
Thermal denaturation of DNA's and the corresponding helix–coil transformation of artificial polyribonucleic and polydeoxyribonucleic acids have been studied extensively both theoretically1–13 and experimentally. 14–30 Much less work has been carried out on the properties of these polynucleic acids at high pressure, and in particular, on the presure dependence of the helix–coil transition temperature.31–33 Light-scattering techniques have been used in this study to measure the pressure dependence of the helix–coil transition temperature of the two- and three-stranded helices of polyriboadenylic and polyribouridilic acids and of calf thymus DNA. From the slopes of the transition temperature vs. pressure curves and heats of transition obtained from the literature,20,34 the following volume changes from these helix–coil transitions have been obtained: (a) ?0.96 cc/mole of nucleotide base pairs for the poly (A + U) transition, (b) +0.35 cc/mole of nucleotide base trios for the poly (A + 2U) transition, and (c) +2.7 cc/mole of nucleotide base pairs for the DNA transition. The relative magnitudes and signs of these volume changes which show that poly (A + U) is destabilized by increased pressure, whereas poly (A + 2U) and calf thymus DNA are stabilized by increased pressure, indicates that further development of the helix–coil transition theory for polynucleotides is needed.  相似文献   

5.
Iwao Satake  Jen Tsi Yang 《Biopolymers》1975,14(9):1841-1846
The conformational phase diagram of poly(L -lysine) (4.6 × 10?4 M, residue) in sodium dodecyl sulfate (1.6 × 10?2 M) solution was constructed from circular dichroism results at various temperatures and pH's. Poly(L -lysine)–sodium dodecyl sulfate complexes undergo a β–helix transition upon raising the pH of the solution. The transition pH tends to shift downward at elevated temperatures. No helix–β transition can be detected for poly(L -lysine) in sodium dodecyl sulfate solution (pH > 11) even after 1-hr heating at 70°C. This is in marked contrast with uncharged poly(L -lysine) solution without sodium dodecyl sulfate, which is converted into the β-form upon mild heating of the solution above 50°C.  相似文献   

6.
The interaction between poly (G) and poly (C) was investigated in neutral and acid medium by optical methods. Three main points arise from this investigation. (1) The formation of poly (G)·poly (C) was complete only above an ionic strength of about 0.6M [Na+]. Lowering the ionic strength increased the amounts of free poly (G) and free poly (C) that could be detected. (2) When titrating towards acid pH values a transition took place which was characterized by potentiometry, mixing curves, and circular dichroism: a three-stranded poly (G)·poly (C)·poly (C+) complex was formed analogous to the transition observed for the acid titration of poly (I)·poly (C). (3) Even when the poly (G)·poly (C) complex was incompletely formed (at low ionic strength) in neutral medium all poly (C) entered the triple-stranded complex.  相似文献   

7.
Thermodynamics of the B to Z transition in poly(dGdC)   总被引:1,自引:0,他引:1  
The thermodynamics of the B to Z transition in poly(dGdC) was examined by differential scanning calorimetry, temperature-dependent absorbance spectroscopy, and CD spectroscopy. In a buffer containing 1 mM Na cacodylate, 1 mM MgCl2, pH 6.3, the B to Z transition is centered at 76.4°C, and is characterized by ΔHcal = 2.02 kcal (mol base pair)?1 and a cooperative unit of 150 base pairs (bp). The tm of this transition is independent of both polynucleotide and Mg2+ concentrations. A second transition, with ΔHcal = 2.90 cal (mol bp)?1, follows the B to Z conversion, the tm of which is dependent upon both the polynucleotide and the Mg2+ concentrations. Turbidity changes are concomitant with the second transition, indicative of DNA aggregation. CD spectra recorded at a temperature above the second transition are similar to those reported for ψ(–)-DNA. Both the B to Z transition and the aggregation reaction are fully and rapidly reversible in calorimetric experiments. The helix to coil transition under these solution conditions is centered at 126°C, and is characterized by ΔHcal = 12.4 kcal (mol bp)?1 and a cooperative unit of 290 bp. In 5 mM MgCl2, a single transition is seen centered at 75.5°C, characterized by ΔHcal = 2.82 kcal (mol bp)?1 and a cooperative unit of 430 bp. This transition is not readily reversible in calorimetric experiments. Changes in turbidity are coincident with the transition, and CD spectra at a temperature just above the transition are characteristic of ψ(–)-DNA. A transition at 124.9°C is seen under these solution conditions, with ΔHcal = 10.0 kcal (mol bp)?1 and which requires a complex three-step reaction mechanism to approximate the experimental excess heat capacity curve. Our results provide a direct measure of the thermodynamics of the B to Z transition, and indicate that Z-DNA is an intermediate in the formation of the ψ-(–) aggregate under these solution conditions.  相似文献   

8.
H J Li  B Brand  A Rotter  C Chang  M Weiskopf 《Biopolymers》1974,13(8):1681-1697
Thermal denaturation of direct-mixed and reconstituted polylysine–DNA complexes in 2.5 × 10?4 M EDTA, pH 8.0 and various concentrations of NaCl has been studied. For both complexes, increasing ionic strength of the solution raises Tm, the melting temperature of free base pairs. The linear dependence of Tm on log Na+ indicates that the concept of electrostatic shielding on phosphate lattice of an infinitely long pure DNA by Na+ can be applied to short free DNA segments in a nucleoprotein. For a direct-mixed polylysine–DNA complex, the melting temperature of bound base pairs Tm′ remains constant at various ionic strengths. On the other hand, the Tm′ in a reconstituted polylysine–DNA complex is shifted to lower temperature at higher ionic strength. This phenomenon occurs for reconstituted complex with long polylysine of one thousand residues or short polylysine of one hundred residues. It is shown that such a decrease of Tm′ is not due to a reduction of coupling melting between free and bound regions in a complex when the ionic strength is raised. It is also not due to intermolecular or intramolecular change from a reconstituted to a direct-mixed complex. It is suggested that this phenomenon is due to structural change on polylysine-bound regions by ionic strength. It is suggested further that Na+ may replace water molecules and bind polylysine-bound regions in a reconstituted complex. Such a dehydration effect destabilizes these regions and lowers Tm′. This explanation is supported by circular dichroism (CD) results.  相似文献   

9.
Sharon S. Yu  Hsueh Jei Li 《Biopolymers》1973,12(12):2777-2788
Protamine–DNA complexes prepared by the method of direct and slow mixing in 2.5 × 10?4M EDTA, pH 8.0, have been studied by thermal denaturation and circular dichroism. The complexes show biphasic melting with Tm at about 50 °C corresponding to the melting of free DNA regions and Tm′ at about 92 °C corresponding to the melting of protamine-bound regions. In protamine-bound regions there are 1.38 amino acid residues per nucleotide, indicating a nearly completely charge neutralization. Tm is increased but Tm′ is not when the ionic strength of the buffer is raised. This also supports a full charge neutralization in protamine-bound regions. The circular dichroism of the complexes can be decomposed into two components, Δε0 of free DNA regions in B-form conformation and Δεb of protamine-bound regions in a characteristic conformation neither that of B- nor C-form but somewhere between them.  相似文献   

10.
Melting parameters of 2U:1A complexes formed by polyuridylic acid [poly(U)] and three adenine dinucleotides, diribonucleoside monophosphonate ApA and diastereoisomers of dideoxyribonucleoside methyl phosphonate [(dApA)1 and (dApA)2], in 1M NaCl and at a number of dinucleotide concentrations were obtained from differential scanning microcalorimetric data and interpreted in terms of the theory of helix–coil equilibrium in oligonucleotide–polynucleotide systems. The apparent binding constant, 1/cm, at 39°C and melting temperatures, Tm, at 1 × 10?3 M dinucleotide concentration indicate the following order of thermodynamic stability of the complexes: 2 poly(U) · (dApA)2 (2.27 × 103M?1, 44.2°C) > 2 poly(U) · (dApA)1 (9.9 × 102M1, 39.2°C) > 2 poly(U) · (ApA) (5.9 × 102M?1, 35.8°C). Corresponding calorimetric enthalpies of melting, ΔHm: 13.5, 12.7, and 12.8 kcal/mol (UUA base triplets) were found to be considerably lower than the van't Hoff enthalpies, ΔHapp: 29.4, 16.2, and 16.2 kcal/mol, respectively, evaluated from the dependence of the melting temperatures on dinucleotide concentration. Self-association of dinucleotides and their simultaneous binding as monomers, dimers, and higher-order associated species is suggested as the most probable cause of the differences between ΔHm and ΔHapp values. The differences in thermodynamic properties of the complexes formed by (dApA)1 and (dApA)2 diastereoisomers are discussed in connection with their known conformational properties. The higher and essentially enthalpic stability of the 2 poly(U) · (dApA)2 complex correlates with a lower degree of intramolecular stacking of the (dApA)2 isomer. The hydrophobically enhanced strong self-association of the latter greatly influences the thermodynamics of its complex formation with poly(U) and results in ΔHappHm = 2.3.  相似文献   

11.
The DNA helix–coil transition has been studied in the presence of high concentrations of manganese ions (about 10?3M), which corresponds to the conditions close to equal stability of the A+T and G+C pairs, at the ionic strengths of 10?1, 10?2, and 1.6 × 10?3M Na+. With the Mn2+ ion effect, the transition range is significantly reduced to not more than 0.2°C at 1.2 × 10?3M Mn2+ and 1.6 × 10?3M Na+. The melting curves display a sharp kink at the end of the helix–coil transition, which is interpreted as an indication of the second-order phase transition. It is shown that the melting curves obtained can be approximated by a simple analytical expression 1 – θ = exp[–a(tc - t)], where θ is the DNA helix fraction, tc is the phase transition temperature, and a is an empirical parameter characterizing the breadth of the melting range and responsible for the magnitude of a jump of the helicity derivative with respect to the temperature at the phase transition point.  相似文献   

12.
The ability of oligodesoxyribonucleotides of various chain lengths to form complexes has been compared with that of oligoribonucleotides. Four series of oligonucleotidcs were prepared and investigated, i.e., dCn at acid pH versus rCn, dAn and dTn versus. rAn and rUn at neutral pH. The results indicate that in dilute solution, the formation of complexes is greatly facilitated in the case of desoxyoligomers and occurs for shorter oligomere than in the corresponding ribooligomers. The spectrophotometric titration of deoxyribooligo C indicates the appearance of two pK values in the 4–5 pH region characteristic of the double-stranded form, which occurs for much shorter dCn than rCn. The circular dichroism (CD.) spectra of deoxycytidylies in dilute solution starting from the trimer are conservative, characteristic of the double-stranded helical form of poly C at acid pH. In contrast, the CD spectra of a series of corresponding ribo Cn, under identical conditions is of nonconservative character similar to that of the single-stranded form of poly C at neutral pH, but differs in the band position. This spectrum is called intermediate. Only at higher concentrations of oligonucleotidcs (i.e., 10?3Minstead of 10?4M) does the circular dichroism spectrum of longer ribocytidylics assume conservative character. Thermal denaturation of deoxycytidylces at acid pH are strongly dependent on chain length and concentration, its one would expect for a cooperative helix-coil transition. The circular dichroism spectra measured at different temperatures shows one isosbestic point. In dilute solution, the standard-state enthalpy change found was 5–6 kcal/mole for higher oligomers (dC7). These properties are all in agreement with a structural transition from the d-Cn double-stranded form to a coil for n > 3. Studies of dAn and dTn in solutions of high ionic strength at low temperature indicate that complex formation occurs already at the level of trimer and for high oligomers. Under identical conditions a complex between rAn and rUn is detected only for oligomers longer than the hexamer. The nature of the “intermediate” form of oligoribo C at acid pH and low temperature was investigated by sedimentation and circular dichroism. A model of rCn is proposed of linear molecules which are partially double-stranded and partially single-stranded, which probably are slowly rearranged by “slippage” into a regular-double-stranded helical form.  相似文献   

13.
M F Hacques  C Marion 《Biopolymers》1986,25(12):2281-2293
CD and uv spectroscopy reveal that the synthetic polynucleotides poly(dG–dC) · poly(dG–dC) and poly(dG–m5dC) · poly(dG–m5dC) undergo a transition induced by small amounts of Ni++ cation from a right-handed B-form to left-handed Z-type conformations. We describe the application of steady-state and transient electric birefringence to the characterization of the transition observed at very low ionic strength (10 mM Tris HCl, pH 7.4). Dialysis experiments show that the changes in spectroscopic and electro-optic properties upon addition of Ni++ are completely reversible. The differences in shape of the inverted CD spectra suggest the existence of a family of left-handed conformations, depending on the polymer used and on the amounts of cation added. The stoichiometry required for inducing the Z-conformation of poly(dG–m5dC) is 1 cation/4 nucleotide phosphates. The transition is accompanied by a decrease in birefringence, an increase in length, and the more important contribution of a permanent or slowly induced dipole moment to the orientation mechanism. High concentrations of Ni++ promote the Z → Z* transition. Upon increasing the Ni++ concentration, poly(dG–dC) undergoes a biphasic transition, first to one intermediate conformation that is neither B- nor Z-like and then to a left-handed form that is probably different from Z*. These conversions are accompanied by regular decreases both in birefringence and in chain length, but no transient appears in the field-reversal experiments.  相似文献   

14.
Quasielastic and static light-scattering measurements were made on DNA isolated from chicken erythrocyte mononucleosomes as a function of ionic strength between 6 × 10?4 and 1.0M. A transition from single-exponential autocorrelation functions to markedly non-single-exponential decays was observed around 10?2M ionic strength and was accompanied by a large decrease in the excess light-scattering intensity. Autocorrelation functions recorded below 10?2M salt were well fit by the sum of two exponential relaxation which differed by as much as 100-fold in time constants. Apparent diffusion coefficients for the fast and slow processes plateaued around 10?3M with numerical values approximately 10-fold and 1/10, respectively, of the translational diffusion coefficient for mononucleosome DNA at high ionic strength. This behavior is similar to that observed with poly(L -lysine), for which the slow decay has been associated with a transition to an extraordinary phase. The strong and complex salt dependence observed here illustrates potential difficulties in deriving structural information from scattering by polyions at low ionic strength.  相似文献   

15.
Studies on poly(L-lysine50, L-tyrosine50)-DNA interaction   总被引:3,自引:0,他引:3  
R M Santella  H J Li 《Biopolymers》1974,13(9):1909-1926
Interaction between poly(Lys50, Tyr50) and DNA has been studied by absorption, circular dichroism (CD), and fluorescence spectroscopy and thermal denaturation in 0.001M Tris, pH 6.8. The binding of this copolypeptide to DNA results in an absorbance enhancement and fluorescence quenching on tyrosine. There is also an increase in the tyrosine CD at 230 nm. The CD of DNA above 250 nm is slightly shifted to the longer wavelength which is qualitatively similar to, but quantitatively much smaller than, that induced by polylysine binding. At physiological pH the poly(Lys50, Tyr50)–DNA complex is soluble until there is one lysine and one tyrosine per nucleotide in the complex. The same ratio of amino acid residues to nucleotide has also been observed in copolypeptide-bound regions of the complex. The addition of more poly(Lys50, Tyr50) to DNA yields a constant melting temperature, Tm′, for bound base pairs at 90°C which is close to that of polylysine-bound DNA under the same condition. The melting temperature, Tm, of free base pairs at about 60°C on the other hand, is increased by 10°C as more copolypeptide is bound to DNA. As the temperature is raised, both absorption and CD spectra of the complexes with high coverage are changed, suggesting structural alteration, perhaps deprotonation, on bound tyrosine. The results in this report also suggest that intercalation of tyrosine in DNA is unlikely to be the mode of binding.  相似文献   

16.
The kinetics of double-helix formation by poly U and the complementary monomer N-6,9-dimethyladenine (m6m9A) has been measured using a new fast temperature-jump apparatus. The cooperative binding kinetics are complicated by the extensive self-association of the monomers, but a satisfactory analysis using average relaxation times was possible in terms of three different models. Application of a model which considers only monomer binding yields the upper limit for the binding rate constant of an m6m9A monomer next to an already bound monomer on a poly U strand: (2 ± 0.4) × 108 M?1sec?1. A lower limit is found by using a model which allows for binding of all m6m9A stacks to poly U with equal rate constants: (3 ± 0.3) × 107 M?1sec?1. A third model with “weighted” rate constants consistent with the data: (7.5 ± 1.0) × 107 M?1sec?1. The rate of cooperative binding of m6m9A to the trimer UpUpU has also been measured. The rate constants obtained with the trimer agree with those obtained with the polymer for each of the three models within experimental error.  相似文献   

17.
T Ito  M Zama  J Amagasa 《Biopolymers》1972,11(8):1583-1592
Circular dichroism spectra of acridine orange bound to E. coli tRNA were studied at varying tRNA phosphate-to-dye (P/D) ratios for both unfractionated and purified materials in the absence of Mg++. From the rather discrete features exhibited in the circular dichroism spectra three types of interactions were observed: (1) A high P/D ratio such as 75.2 or 49.8 indicates the interaction between the nucleotide base and dye molecule. The spectra with a large positive peak at 515 mμ are, however, quite different from that of DNA–AO complex under similar conditions. (2) With an intermediate P/D ratio (26.5 to 9.6) dye molecules bound strongly to the polynucleotide chain. (3) With low P/D ratios (≤7.5) the interaction appears to be due to the stacked dye molecules in the single-stranded part of tRNA. The spectra of the third group have an isobestic point at 477 mμ. Below a P/D ratio of 4 the spectrum shows one positive and two negative bands which may be the characteristics of circular dichroism of stacked dyes in polynucleotide chain. Although no drastic change in the conformation of tRNA itself was detectable in the presence of Mg++ in the ultraviolet region, a dramatic change was observed in the circular dichroism of tRNA–acridine orange complex when Mg++ concentration was increased to 10?3M. It was inferred that certain conformational changes other than simple hydrogen bond formation occured in tRNA molecules at this high Mg++ concentration, so that the amount of bound dye in the stacking condition was increased through the transition.  相似文献   

18.
G Govil  I C Smith 《Biopolymers》1973,12(11):2589-2598
The temperature-dependent conformations of poly(U) in 0.5M CsC1 have been studied by carbon-13 nuclear magnetic resonance. The transition from random coil to an ordered structure results in broadening of lines in the 13C spectra, due to intramolecular 1H–13C dipolar interactions and restricted motions in the ordered state. Changes in the chemical shifts suggest that the bases are interacting below the transition temperature. The random coil form shows conformation preferences for internal rotation about C4′–C5′, C5′–O5′, and C3′–O 3′ bonds. The statistical randomness of the coil arises mainly because of flexibility about O–P bonds. The results are analyzed in conjunction with theoretical calculations and light-scattering data.  相似文献   

19.
The three-dimensional Poisson–Boltzmann equation for the distribution of counterion charge density around double-helical DNA has been solved for solutions of .01M, .10M, and .20M monovalent salt. The polymers, poly[d(CpGp)] and poly[d(m5CpGp)], were studied in the B- and the Z-conformations. The effect of methylation on the relative stabilities of these conformers in solutions of different ionic strengths is known to favor the Z-form. Accumulation of charge density around the B- and the Z-conformers is compared in detail. The relative electrostatic stabilities of the B- and Z-conformers in .01M, .10M, and .20M solutions are compared and discussed in terms of the ion–DNA interactions and the self-energy of the structured ionic environment. The ion–DNA interaction energies, termed “phosphate screening,” monotonically decrease with ionic strength and are consistent with a B-to-Z conformation change induced in either polymer by increased electrolyte concentration. However, these calculated energies alone do not account for the fact that the ionic strength at the midpoint of the transition of the methylated polymer is substantially lower than that of its unmethylated analogues. The phosphate screening effect is counterbalanced by changes in the self-energy required for the creation of the structured counterion environment. This self-energy of the electrolyte environment monotonically increases with ionic strength. Methylation-induced shifts in the overall conformational equilibria depend on the relative changes of these competing effects. Increasing salt concentration is calcualted to favor the Z-conformer. The effect of methylation, lowering the ionic strength of the transition midpoint, is proposed to originate in minor structural changes in the Z-form of the polymer, making the groove more accessible to counterions in the G(3′ – 5′)C region. This allows a redistribution of counterion density and a lowering of the self-energy of the ionic environment, conferring added stability to the Z-conformation, as indicated by calculations of relative entropies. The experimentally observed temperature dependence of the B-to-Z transition, however, cannot be explained without assuming the release of bound water. Maps of the calculated three-dimensional structure at the counterion distribution near the surface of these molecules in both the B- and the Z-forms are also presented.  相似文献   

20.
The interaction between B-form DNA and twelve cationic triaryl-methane dyes was studied with respect lo optical properties and stabilities, using linear dichroism (LD) and aqueous two-phase partition techniques. Monovalent dyes derived from crystal violet as a rule form a single strong complex (K1 ca 105 M?1; site density per nucleotide base n1 ca 0.1 at 0.1M ionic strength) in which the plane of the dye is at an angle of less than 50° to the local DNA helix axis. The complex with fuchsin is weaker (104M?1) but can be explained by a similar orientation. For some of the dyes (those with pseudo-C2v symmetry) XXXre angular orientations of two molecule-fixed axes can be obtained. For the divalent methyl green a second complex appears to be formed at low ionic strength. Methyl green (and to some extent 2-thiophene green and malachite green) show exciton splitting in the LD spectrum and circular dichroism assignable to exciton coupling between transition dipoles roughly parallel to the helical strands, indicating a dye-dye interaction. Tne optical data, supported by fitting experiments with space-filling models, suggests a general structure for the binding site. The dye is not intercalated but is bound to exposed hydrophobic regions in the major groove. The ligand is in part (the charged amino groups) in contact with the phosphoribose chain but its main surface lies against the hydrophobic base-pair stack. For a diphenylmethane dye, Michler's hydrol blue, a perpendicular orientation was observed, possibly due to intercaiation.  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