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.
Isotherms of the EtBr adsorption on native and denatured poly(dA)poly(dT) in the temperature interval 20–70°C were obtained. The EtBr binding constants and the number of binding sites were determined. The thermodynamic parameters of the EtBr intercalation complex upon changes of solution temperature 20–48°C were calculated: 1.0·106 M−1K≤1.4·106 M−1, free energy ΔG o=−8.7±0.3 kcal/mol, enthalpy ΔH o≅0, and entropy ΔS o=28±0.5 cal/(mol deg). UV melting has shown that the melting temperature (T m) of EtBr-poly(dA)poly(dT) complexes (μ=0.022,4.16·10−5 M EtBr) increased by 17°C as compared with the ΔT m of free homopolymer, whereas the half-width of the transition (T m) is not changed. It was shown for the first time that EtBr forms complexes of two types on single-stranded regions of poly(dA)poly(dT) denatured at 70°C: strong (K 1=1.7·105 M−1; ΔG o=−8.10±0.03 kcal/mol) and weak (K 2=2.9·103 M−1; ΔG o=−6.0±0.3 kcal/mol).The ΔG o of the strong and weak complexes was independent of the solution ionic strength, 0.0022≤μ≤0.022. A model of EtBr binding with single-stranded regions of poly(dA)poly(dT) is discussed.  相似文献   

2.
Coralyne is a small crescent-shaped molecule known to intercalate duplex and triplex DNA. We report that coralyne can cause the complete and irreversible disproportionation of duplex poly(dT)·poly(dA). That is, coralyne causes the strands of duplex poly(dT)·poly(dA) to repartition into equal molar equivalents of triplex poly(dT)·poly(dA)·poly(dT) and poly(dA). Poly(dT)·poly(dA) will remain as a duplex for months after the addition of coralyne, if the sample is maintained at 4°C. However, disproportionation readily occurs upon heating above 35°C and is not reversed by subsequent cooling. A titration of poly(dT)·poly(dA) with coralyne reveals that disproportionation is favored by as little as one molar equivalent of coralyne per eight base pairs of initial duplex. We have also found that poly(dA) forms a self-structure in the presence of coralyne with a melting temperature of 47°C, for the conditions of our study. This poly(dA) self-structure binds coralyne with an affinity that is comparable with that of triplex poly(dT)·poly(dA)·poly(dT). A Job plot analysis reveals that the maximum level of poly(dA) self-structure intercalation is 0.25 coralyne molecules per adenine base. This conforms to the nearest neighbor exclusion principle for a poly(dA) duplex structure with A·A base pairs. We propose that duplex disproportionation by coralyne is promoted by both the triplex and the poly(dA) self-structure having binding constants for coralyne that are greater than that of duplex poly(dT)·poly(dA).  相似文献   

3.
Deoxyribonucleic acid is the site of storage and retrieval of genetic information through interaction with proteins and other small molecules. In the present study, the interaction of two natural cytotoxic protoberberine plant alkaloids, berberine and palmatine, and a synthetic derivative, coralyne, with mammalian herring testis DNA was investigated using a combination of isothermal titration calorimetry, differential scanning calorimetry, and optical melting experiments to characterize the energetics of their binding. The binding constants of these alkaloids to DNA under identical conditions were evaluated from the UV melting data, and the enthalpy of binding was elucidated from isothermal titration studies. The binding constants of berberine, palmatine, and coralyne to DNA were found to be 1.15 x 10(4), 2.84 x 10(4), and 3.5 x 10(6) M(-1) at 20 degrees C in buffer of 20 mM [Na+]. Parsing of the free energy change of the interaction observed into polyelectrolytic and nonpolyelectrolytic components suggested that although these alkaloids are charged, the major contributor of about 75% of the binding free energy arises from the nonpolyelectrolytic forces. The binding in case of palmatine and coralyne was predominantly enthalpy driven with favoring smaller entropy terms, while that of berberine was favored by both negative enthalpy and positive entropy changes. Temperature dependence of the binding enthalpies determined from ITC studies in the range 20-40 degrees C was used to calculate the binding-induced change in heat capacity (DeltaC(o)(p)) values as -117, -135, and -157 cal/mol K, respectively, for berberine, palmatine, and coralyne. Taken together, the results suggest that the DNA binding of the planar synthetic coralyne is stronger and thermodynamically more favored compared to the buckled natural berberine and palmatine.  相似文献   

4.
The interaction of berberine, palmatine, and coralyne with the B, Z, and HL form of poly[d(G‐C)] was studied. Berberine and palmatine showed moderate binding to the B form, while coralyne showed higher binding, as revealed from spectroscopic and thermodynamic data. Berberine and coralyne binding to the B form was exothermic and enthalpy‐driven, while palmatine showed exothermic binding which was favored by both negative enthalpy and negative entropy changes. Berberine and palmatine neither bind nor converted the Z‐form structure to B form. Coralyne, on the other hand, exhibited a strong binding affinity to Z DNA structure that was enthalpy‐driven. Berberine binding to the HL form was cooperative, exothermic, and favored by both negative enthalpy and negative entropy changes with the formation of an induced CD band. Palmatine showed weak binding, while coralyne showed a strong binding with the HL form. The structural differences in the isoquinoline alkaloids appear to influence the affinity and mode of interactions with these polymorphic DNA structures.  相似文献   

5.
6.
The binding of propidium to poly(dA).poly(dT) [poly(dA.dT)] and to poly[d(A-T)].poly[d(A-T)] [poly[d(A-T)2]] has been compared under a variety of solution conditions by viscometric titrations, binding studies, and kinetic experiments. The binding of propidium to poly[d(A-T)2] is quite similar to its binding to calf thymus deoxyribonucleic acid (DNA). The interaction with poly(dA.dT), however, is quite unusual. The viscosity of a poly(dA.dT) solution first decreases and then increases in a titration with propidium at 18 degrees C. The viscosity of poly[d(A-T)2] shows no decrease in a similar titration. Scatchard plots for the interaction of propidium with poly(dA.dT) show the classical upward curvature for positive cooperativity. The curvature decreases as the temperature is increased in binding experiments. A van't Hoff plot of the observed binding constants yields an apparent positive enthalpy of approximately +6 kcal/mol for the propidium-poly(dA.dT) interaction. Propidium binding to poly[d(A-T)2] shows no evidence for positive cooperativity, and the enthalpy change for the reaction is approximately -9 kcal/mol. Both the magnitude of the dissociation constants and the effects of ionic strength are quite similar for the dissociation of propidium from poly(dA-T)2] and from poly[d(A-T)2], suggesting that the intercalated states are similar for the two complexes. The observed association reactions, under pseudo-first-order conditions, are quite different. Plots of the observed pseudo-first-order association rate constant vs. polymer concentration have much larger slopes for propidium binding to poly[d(A-T)2] than to poly(dA.dT).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
The effect of hydrostatic pressure on the helix-coil transition temperature (Tm) was measured for the DNA oligomers (dA)n(dT)n, where n = 11, 15, and 19, in 50 mM NaCl. The data were analyzed in light of previously published data for the polymer, poly(dA)·poly(dT) under the same conditions. As has been observed for DNA polymers, increasing the hydrostatic pressure led to an increase in the Tm of the oligomers; however, the effect of pressure diminished with decreasing chain length. The value of dTm/dP decreased linearly with the inverse of the chain length varying from 3.15 × 10−2°C MPa−1 for the polymer to 0.7 × 10−2°C MPa−1 for the 11-mer. The two-state or van't Hoff enthalpy (ΔHvH) of the helix-coil transition was obtained by analysis of the half-width of the thermal transition. As expected, ΔHvH decreases with decreasing chain length. In contrast to the behavior of the polymer, poly(dA)·poly(dT), and (dA)19(dT)19, the ΔHvH of the two shorter duplex oligonucleotides displayed a small pressure dependence dΔHvH/dP≃−0.4 kJ MPa−1 in both cases. The changes observed in the Tm and ΔHvH were not sufficient to explain the magnitude of the chain-length dependence of the pressure effect. To interpret the large chain-length dependence of dTm/dP, we propose that the terminal base pairs contribute a negative volume change to the helix-coil transition. Base pairs distant from the ends exhibit behavior characterized by the polymer where end effects are assumed to be negligible, i.e., a positive volume change for the helix-coil transition. The negative volume change of separating terminal bases may originate from the imperfect interactions these base pairs form with water due to the existence of several energetically equivalent conformations. This is reminiscent of one of the mechanisms proposed to be important in the pressure-induced dissociation of multimeric proteins into their constituent subunits. © 1996 John Wiley & Sons, Inc.  相似文献   

8.
R M Wartell 《Biopolymers》1972,11(4):745-759
Helix–coil transition curves are calculated for poly (dA) poly(dT) and poly (dA-dT) poly (dA-dT) using the integral equation approach of Goel and Montroll.5 The transitions are described by the loop entropy model with the exponent of the loop entropy factor, k, remaining an arbitrary constant. The theoretical calculations are compared with experimental transition curves of the two polymers. Results indicate that the stacking energies for these two polymers differ by about 1 kcal/mole of base pairs. Also, a fit between theory and experiment was not possible for k > 1.70.  相似文献   

9.
CD spectra and melting curves were collected for a 28 base-pair DNA fragment in the form of a DNA dumbbell (linked on both ends by T4 single-strand loops) and the same DNA sequence in the linear form (without end loops). The central 16 base pairs (bp) of the 28-bp duplex region is the poly(pu) sequence: 5′-AGGAAGGAGGAAAGAG-3′. Mixtures of the dumbbell and linear DNAs with the 16-base single-strand sequence 5′-TCCTTCCTCCTTTCTC-3′ were also prepared and studied. At 22°C, CD measurements of the mixtures in 950 mM NaCl, 10 mM sodium acetate, 1 mM EDTA, pH 5.5, at a duplex concentration of 1.8 μM, provided evidence for triplex formation. Spectroscopic features of the triplexes formed with either a dumbbell or linear substrate were quite similar. Melting curves of the duplex molecules alone and in mixtures with the third strand were collected as a function of duplex concentration from 0.16 to 2.15 μM. Melting curves of the dumbbell alone and mixtures with the third strand were entirely independent of DNA concentration. In contrast, melting curves of the linear duplex alone or mixed with the third strand were concentration dependent. At identical duplex concentrations, the dumbbell alone melts ~20°C higher than the linear duplex. The curve of the linear duplex displayed a significant pretransition probably due to end fraying. On melting curves of mixtures of the dumbbell or linear duplex with the third strand, a low temperature transition with much lower relative hyperchromicity change (~ 5%) was observed. This transition was attributed to the melting of a new molecular species, e.g., the triplex formed between the duplex and single-strand DNA molecules. In the case of the dumbbell/single-strand mixture, these melting transitions of the triplex and the dumbbell were entirely resolvable. In contrast, the melting transitions of the linear duplex and the triplex overlapped, thereby preventing their clear distinction. To analyze the data, a three-state equilibrium model is presented. The analysis utilizes differences in relative absorbance vs temperature curves of dumbbells (or linear molecules) alone and in mixtures with the third strand. From the model analysis a straightforward derivation of fT(T), the fraction of triplex as a function of temperature, was obtained. Analysis of fT vs temperature curves, in effect melting curves of the triplexes, provided evaluation of thermodynamic parameters of the melting transition. For the triplex formed with the dumbbell substrate, the total transition enthalpy is ΔHT = 118.4 ± 12.8 kcal/mol (7.4 ± 0.8 kcal/mol per triplet unit) and the total transition entropy is ΔST = 344 ± 36.8 cal/K · mol (eu) (21.5 ± 2.3 eu per triple unit). The transition curves of the triplex formed with the linear duplex substrate displayed two distinct regions. A broad pretransition region from fT = 0 to 0.55 and a higher, sharper transition above fT = 0.55. The transition parameters derived from the lower temperature region of the curve are ΔHT = 44.8 ± 9.6 kcal/mol and ΔST = 112 ± 33.6 eu (or ΔH′ = 2.8 ± 0.6 kcal/mol and ΔS′ = 7.0 ± 2.1 eu per triplet). These values are probably too small to correspond to actual melting of the triplex but instead likely reveal effects of end fraying of the duplex substrate on triplex stability. Transition parameters of the upper transition are ΔHT = 128.0 ± 2.3 kcal/mol and ΔST = 379.2 ± 6.4 eu (ΔH′ = 8.0 ± 0.2 kcal/mol and ΔS′ = 23.7 ± 0.4 eu per triplet) in good agreement (within experimental error) with the transition parameters of the triplex formed with the dumbbell substrate. Supposing this upper transition reflects actual dissociation of the third strand from the linear duplex substrate this triplex is comparable in thermodynamic stability to the triplex formed with a dumbbell substrate. Even so, the biphasic melting character of the linear triplex obscures the whole analysis, casting doubt on its absolute reliability. Apparently triplexes formed with a dumbbell substrate offer technical advantages over triplexes formed from linear or hairpin duplex substrates for studies of DNA triplex stability. © 1993 John Wiley & Sons, Inc.  相似文献   

10.
Interaction of the 9-ON-aryl/arylalkyl amino carbonyl methyl substituted analogs of the anticancer isoquinoline alkaloid berberine with RNA triplex, poly(U)-poly(A)·poly(U) has been studied in comparison to the duplex poly(A)-poly(U), using multiple biophysical techniques. Spectrophotometric and spectrofluorimetric studies established the non-cooperative binding mode of all the analogs with both the duplex and the triplex. However, berberine exhibited cooperative binding with poly(A)-poly(U) and non-cooperative binding with poly(U)-poly(A)·poly(U). Analog BER1 showed the highest affinity to both the duplex and the triplex followed by BER2 and BER3. The overall binding affinity varied as BER1 > BER2 > BER3 > BER. The magnitude of the quantum efficiency values (Q > 1) revealed that energy was transferred from the bases of the triplex and the duplex to the analogs. Comparative ferrocyanide quenching and viscosity studies unambiguously established a stronger intercalative geometry of the analogs to both the triplex and the duplex in comparison to berberine. Circular dichroism studies revealed that the alkaloids perturbed the conformation of both RNA helices. The binding of all the alkaloids was found to be exothermic from isothermal titration studies. Binding of the analogs was highly entropy driven while that of berberine was enthalpy dominated. The results presented here reveal strong and specific binding of these new berberine analogs to the RNA triplex and duplex and highlight the remarkable influence of the 9-substitution on the interaction profile.  相似文献   

11.
The equilibrium binding of the cytotoxic plant alkaloid berberine to various DNAs and energetics of the interaction have been studied. At low ratios of bound alkaloid to base pair, the binding exhibited cooperativity to natural DNAs having almost equal proportions of AT and GC sequences. In contrast, the binding was non-cooperative to DNAs with predominantly high AT or GC sequences. Among the synthetic DNAs, cooperative binding was observed with poly(dA).poly(dT) and poly(dG).poly(dC) while non-cooperative binding was seen with poly(dA–dT).poly(dA–dT) and poly(dG–dC).poly(dG–dC). Both cooperative and non-cooperative bindings were remarkably dependent on the salt concentration of the media. Linear plots of ln Ka versus [Na+] for poly(dA).poly(dT) and poly(dA–dT).poly(dA–dT) showed the release of 0.56 and 0.75 sodium ions respectively per bound alkaloid. Isothermal titration calorimetry results revealed the binding to be exothermic and favoured by both enthalpy and entropy changes in all DNAs except the two AT polymers and AT rich DNA, where the same was predominantly entropy driven. Heat capacity values (ΔCpo) of berberine binding to poly(dA).poly(dT), poly(dA–dT).poly(dA–dT), Clostridium perfringens and calf thymus DNA were − 98, − 140, − 120 and − 110 cal/mol K respectively. This study presents new insights into the binding dependent base pair heterogeneity in DNA conformation and the first complete thermodynamic profile of berberine binding to DNAs.  相似文献   

12.
The modes of binding of 5′‐[4‐(aminoiminomethyl)phenyl]‐[2,2′‐Bifuran]‐5‐carboximidamide (DB832) to multi‐stranded DNAs: human telomere quadruplex, monomolecular R‐triplex, pyr/pur/pyr triplex consisting of 12 T*(T·A) triplets, and DNA double helical hairpin were studied. The optical adsorption of the ligand was used for monitoring the binding and for determination of the association constants and the numbers of binding sites. CD spectra of DB832 complexes with the oligonucleotides and the data on the energy transfer from DNA bases to the bound DB832 assisted in elucidating the binding modes. The affinity of DB832 to the studied multi‐stranded DNAs was found to be greater (Kass ≈ 107M?1) than to the duplex DNA (Kass ≈ 2 × 105M?1). A considerable stabilizing effect of DB832 binding on R‐triplex conformation was detected. The nature of the ligand tight binding differed for the studied multi‐stranded DNA depending on their specific conformational features: recombination‐type R‐triplex demonstrated the highest affinity for DB832 groove binding, while pyr/pur/pyr TTA triplex favored DB832 intercalation at the end stacking contacts and the human telomere quadruplex d[AG3(T2AG3)3] accommodated the ligand in a capping mode. Additionally, the pyr/pur/pyr TTA triplex and d[AG3(T2AG3)3] quadruplex bound DB832 into their grooves, though with a markedly lesser affinity. DB832 may be useful for discrimination of the multi‐sranded DNA conformations and for R‐triplex stabilization. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 8–20, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

13.
K Zieba  T M Chu  D W Kupke  L A Marky 《Biochemistry》1991,30(32):8018-8026
The role of water in the formation of stable duplexes of nucleic acids is being studied by determining the concurrent volume change, heats, and counterion uptake that accompany the duplexation process. The variability of the volume contraction that we have observed in the formation of a variety of homoduplexes suggests that sequence and conformation acutely affect the degree of hydration. We have used a combination of densimetric and calorimetric techniques to measure the change in volume and enthalpy resulting from the mixing of two complementary strands to form (a) fully paired duplexes with 10 or 11 base pairs and (b) bulged decameric duplexes with an extra dA or dT unmatched residue. We also monitored absorbance vs temperature profiles as a function of strand and salt concentration for all four duplexes. Relative to the decamer duplex, insertion of an extra dA.dT base pair to form an undecamer duplex results in a favorable enthalpy of -5.6 kcal/mol that is nearly compensated by an unfavorable entropy term of -5.1 kcal/mol. This enthalpy difference correlates with a differential uptake of water molecules, corresponding to an additional hydration of 16 mol of water molecules/mol of base pair. Relative to the fully paired duplexes, both bulged duplexes are 12-16 degrees C less stable and exhibit marginally larger counterion uptake on forming the duplex. The enthalpy change is slightly lower for the T-bulge duplex and less still for the A-bulge duplex. The volume change results indicate that an unmatched residue increases the amount of coulombic and/or structural hydration. The combined results strongly suggest that the destabilizing forces in bulged duplexes are partially compensated by an increase in hydration levels.  相似文献   

14.
Effects of Ni2+ and Cd2+ ions on thermally induced conformational transitions in the poly(dA)·poly(dT) polynucleotide duplex and poly(dA)·2poly(dT) triplex under near physiological ionic conditions were studied by measurement of UV absorption melting curves and static light scattering intensity. The diagrams of conformational transitions in poly(dA)-poly(dT)-Me2+ systems were plotted. An aggregation in these polynucleotide systems arises at certain values of the metal ions concentration and the temperature after the polymer dissociation into single strands. The phenomenon is conditioned by the aggregation of poly(dA) via the interstrand cross-linking by the dication bridges. Unlike Ni2+, Cd2+ induces formation of very stable aggregates which did not disintegrate even upon cooling up to room temperature.  相似文献   

15.
Y C Fu  H V Wart  H A Scheraga 《Biopolymers》1976,15(9):1795-1813
The enthalpy change associated with the isothermal pH-induced uncharged coil-to-helix transition ΔHh° in poly(L -ornithine) in 0.1 N KCl has been determnined calorimetrically to be ?1530 ± 210 and ?1270 ± 530 cal/mol at 10° and 25°C, respectively. Titration data provided information about the state of charge of the polymer in the calorimetric experiments, and optical rotatory dispersion data about its conformation. In order to compute ΔHh°, the observed calorimetric heat was corrected for the heat of breaking the sample cell, the heat of dilution of HCl, the heat of neutralization of the OH? ion, and the heat of ionization of the δ-amino group in the random coil. The latter was obtained from similar calorimetric measurements on poly(D ,L -ornithine). Since it was discovered that poly(L -ornithine) undergoes chain cleavage at high pH, the calorimetric measurements were carried out under conditions where no degradation occurred. From the thermally induced uncharged helix–coil transition curve for poly(L -ornithine) at pH 11.68 in 0.1 N KCl in the 0°–40°C region, the transition temperature Ttr and the quantity (?θh/?T)Ttr have been obtained. From these values, together with the measured values of ΔHh°, the changes in the standard free energy ΔGh° and entropy ΔGh°, associated with the uncharged coil-to-helix transition at 10°C have been calculated to be ?33 cal/mol and ?5.3 cal/mol deg, respectively. The value of the Zimm–Bragg helix–coil stability constant σ has been calculated to be 1.4 × 10?2 and the value of s calculated to be 1.06 at 10°C, and between 0.60 and 0.92 at 25°C.  相似文献   

16.
Ross PD  Howard FB 《Biopolymers》2003,68(2):210-222
To assess the thermodynamic contribution of the 5-methyl group of thymine, we have studied the two-stranded helical complexes poly(dA).poly(dU) and poly(dA).poly(dT) and the three-stranded complexes--poly(dA).2poly(dU), poly(dA).poly(dT).poly(dU) and poly(dA).2poly(dT)--by differential scanning calorimetry, and uv optical melting experiments. The thermodynamic quantities associated with the 3 --> 2, 2 --> 1, and 3 --> 1 melting transitions are found to vary with salt concentration and temperature in a more complex manner than commonly believed. The transition temperatures, T(m), are generally not linear in the logarithm of concentration or activity of NaCl. The change in enthalpy and in entropy upon melting varies with salt concentration and temperature, and a change in heat capacity accompanies each transition. The poly(dA).2poly(dU) triple helix is markedly different from poly(dA).2poly(dT) in both its CD spectrum and thermodynamic behavior, while the poly(dA).poly(dT).poly(dU) triple helix resembles poly(dA).2poly(dT) in these properties. In comparing poly(dA).2poly(dT) with either the poly(dA).poly(dT).poly(dU) or the poly(dA).2poly(dU) triplexes, the substitution of thymine for uracil in the third strand results in an enhancement of stability against the 3 --> 2 dissociation of deltadeltaG degrees = -135 +/- 85 cal (mol A)(-1) at 37 degrees C. This represents a doubling of the absolute stability toward dissociation compared to the triplexes with poly(dU) as the third strand. The poly (dA).poly (dT) duplex is more stable than poly(dA).poly(dU) by deltadeltaG degrees = -350 +/- 60 cal (mol base pair)(-1) at 37 degrees C. Poly(dA).poly(dT) has 50% greater stability than poly(dA).poly(dU) as a result of the dT for dU substitution in the duplex.  相似文献   

17.
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   

18.
H J Hinz  W Haar  T Ackermann 《Biopolymers》1970,9(8):923-936
The enthalpies of the helix-coil transitions of the ordered polynucleotide systems of poly(inosinic acid)–poly(cytidylic acid) [poly(I + C)], (helical duplex), and of poly (inosinic acid) [poly(I + I + I)], (proposed secondary structure: a triple-stranded helical complex), were determined by using an adiabatic twin-vessel differential calorimeter. Measuring the temperature course of the heat capacity of the aqueous polymer solutions, the enthalpy values for the dissociation of the helical duplex poly (I + C) and the three-stranded helical complex poly(I + 1 + 1), respectively, were obtained by evaluating the additional heat capacity involved in the conformational change of the polynucleotide system in the transition range. The ΔH values of the helix-coil transition of poly (I + C) resulting from the analysis of the calorimetric measurements vary between the limits 6.5 ± 0.4 kcal/mole (I + C) and 8.4 ± 0.4 kcal/mole (I + C). depending on the variation of the cation concentration ranging from 0.063 mole cations kg H2O to 1.003 mole cations/kg H2O. The calorimetric investigation of an aqueous poly I solution (cation concentration 1.0 mole/kg H2O) yielded the enthalpy value ΔH = 1.9 ± 0.4 kcal/mole (I), a result which has been interpreted qualitatively following current models of inter- and intramolecular forces of biologically significant macromolecules. Additional information on the transition behavior of poly(I+ C)Was obtained by ultraviolet and infrared absorption measurements.  相似文献   

19.
The existence of a soluble complex formed by polyuridylic acid (poly (U)) and 3′,5′-cyclic AMP (cAMP) is demonstrated by u.v. extinction vs. temperature curves, optical rotation, equilibrium dialysis, and reaction calorimetry. The complex hasthe stoichiometry of 2 poly (U)-cAMP and its formation is accompanied by an enthalpy change of ?13.0 kcal/mole of base triplet. The introuction of an empirical factor α in the equations given by Damle2 and Crothers2 leads to the evolution of a ΔH value of ?13.4 keal/mole. The parameter α is considered as a correction factor for the concentration dependence of the binding process. There is no relation between α and the reduction of monomer activity due to self-association of monomers. The study of the binding process at several temperatures showed that the cooperativity parameter, σ, is independent of temperature and its value of 6.5 × 10?3 is in good agreement with σ = 5 × 10?3 for the poly (U)·poly(A) system.3  相似文献   

20.
Abstract

Thermodynamic parameters of melting process (δHm, Tm, δTm) of calf thymus DNA, poly(dA)poly(dT) and poly(d(A-C))·poly(d(G-T)) were determined in the presence of various concentrations of TOEPyP(4) and its Zn complex. The investigated porphyrins caused serious stabilization of calf thymus DNA and poly poly(dA)poly(dT), but not poly(d(A-C))poly(d(G-T)). It was shown that TOEpyp(4) revealed GC specificity, it increased Tm of satellite fraction by 24°C, but ZnTOEpyp(4), on the contrary, predominately bound with AT-rich sites and increased DNA main stage Tm by 18°C, and Tm of poly(dA)poly(dT) increased by 40 °C, in comparison with the same polymers without porphyrin. ZnTOEpyp(4) binds with DNA and poly(dA)poly(dT) in two modes—strong and weak ones. In the range of r from 0.005 to 0.08 both modes were fulfilled, and in the range of r from 0.165 to 0.25 only one mode—strong binding—took place. The weak binding is characterized with shifting of Tm by some grades, and for the strong binding Tm shifts by ~ 30–40°C. Invariability of ΔHm of DNA and poly(dA)poly(dT), and sharp increase of Tm in the range of r from 0.08 to 0.25 for thymus DNA and 0.01–0.2 for poly(dA)poly(dT) we interpret as entropic character of these complexes melting. It was suggested that this entropic character of melting is connected with forcing out of H2O molecules from AT sites by ZnTOEpyp(4) and with formation of outside stacking at the sites of binding. Four-fold decrease of calf thymus DNA melting range width ΔTm caused by increase of added ZnTO- Epyp(4) concentration is explained by rapprochement of AT and GC pairs thermal stability, and it is in agreement with a well-known dependence, according to which ΔT~TGC-TAT for DNA obtained from higher organisms (L. V. Berestetskaya, M. D. Frank-Kamenetskii, and Yu. S. Lazurkin. Biopolymers 13, 193–205 (1974)). Poly (d(A-C))poly(d(G-T)) in the presence of ZnTOEpyp(4) gives only one mode of weak binding. The conclusion is that binding of ZnTOEpyp(4) with DNA depends on its nucleotide sequence.  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