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.
2.
L A Marky  R B Macgregor 《Biochemistry》1990,29(20):4805-4811
We report differences in the interaction of two structurally similar phenanthroline intercalators, ethidium and propidium, with poly(dA).poly(dT) and poly[d(A-T)] as a function of ionic strength based on titration microcalorimetry, fluorescence titration, and hydrostatic pressure measurements. Both ethidium and propidium bind more strongly to poly[d(A-T)].poly[d(A-T)] than to poly(dA).poly(dT). Ethidium intercalation into the latter polymer displays titrations with positive cooperativity; this is not found with propidium. The enthalpy of intercalation (delta H degrees) is exothermic for both dyes with poly[d(A-T)].poly[d(A-T)]; however, the value of this parameter is nearly zero in the case of poly(dA).poly(dT). The molar volume change (delta V degrees) accompanying dye intercalation is negative under all conditions for poly[d(A-T)].poly[d(A-T)] whereas it is positive for poly(dA).poly(dT). The changes observed in delta V degrees correlate well with the entropy changes derived from the titration and calorimetric data for this reaction. The results, interpreted in terms of the relative hydration of these two polymers, are consistent with a higher extent of hydration of poly(dA).poly(dT) relative to poly[d(A-T)].poly[d(A-T)].  相似文献   

3.
We report the temperature and salt dependence of the volume change (DeltaVb) associated with the binding of ethidium bromide and netropsin with poly(dA).poly(dT) and poly[d(A-T)].poly[d(A-T)]. The DeltaV(b) of binding of ethidium with poly(dA).poly(dT) was much more negative at temperatures approximately 70 degrees C than at 25 degrees C, whereas the difference is much smaller in the case of binding with poly[d(A-T)].poly[d(A-T)]. We also determined the volume change of DNA-drug interaction by comparing the volume change of melting of DNA duplex and DNA-drug complex. The DNA-drug complexes display helix-coil transition temperatures (Tm several degrees above those of the unbound polymers, e.g., the Tm of the netropsin complex with poly(dA)poly(dT) is 106 degrees C. The results for the binding of ethidium with poly[d(A-T)].poly[d(A-T)] were accurately described by scaled particle theory. However, this analysis did not yield results consistent with our data for ethidium binding with poly(dA).poly(dT). We hypothesize that heat-induced changes in conformation and hydration of this polymer are responsible for this behavior. The volumetric properties of poly(dA).poly(dT) become similar to those of poly[d(A-T)].poly[d(A-T)] at higher temperatures.  相似文献   

4.
L A Marky  D W Kupke 《Biochemistry》1989,28(26):9982-9988
The minor-groove ligand netropsin provides a sensitive probe of the hydration difference between poly(dA).poly(dT) and poly[d(AT)].poly[d(AT)]. We have measured the volume change delta V accompanying binding of netropsin to these polymers, using an improved magnetic suspension densimeter. For poly(dA).poly(dT) we find delta V = +97 mL/mol of bound netropsin at pH 7.0 and 10 mM sodium phosphate buffer. For poly[d(AT)].poly[d(AT)] we find delta V = -16 mL/mol of bound netropsin. This striking differential effect suggests that the poly(dA).poly(dT) duplex compresses more water (or is more extensively hydrated). From our enthalpy and entropy results we estimate the approximately 10 water molecules, immobilized in the minor groove of this system, are displaced by each netropsin bound. The volume increase, however, is substantially larger than can be explained by a simple melting of these immobilized water molecules in the minor groove. A decompression of at least 40 water molecules must attend the complexation to the poly(dA).poly(dT) duplex. This suggests that the conformation change attending the binding of the drug to this polymer duplex causes a further dehydration, whereas no such change in dehydration and configuration for the heteropolymer system is indicated.  相似文献   

5.
The kinetics of the hydrogen-deuterium exchange reactions of double-helical poly[d(A-T)]·poly[d(A-T)], poly(dA)·poly(dT), and constituent nucleosides (deoxyadenosine and thymidine) have been examined at various temperatures by stopped-flow ultraviolet spectrophotometry, in the spectral region 240–300 nm. The results were interpreted on the basis of a mechanism of the hydrogen exchange reaction of a helical polynucleotide, proposed by Englander and colleagues as well as by the Tsuboi and Nakanishi group. It was concluded that the rates of the base-pair opening reactions are nearly equal to one another in double-helical DNAs, irrespective of the base sequence. On the other hand, the free energy required for bringing the open segment at a particular base-pair was found to be much greater for poly(dA)·poly(dT) than for poly[d(A-T)]· poly[d(A-T)].  相似文献   

6.
H Takashima  M Nakanishi  M Tsuboi 《Biochemistry》1985,24(18):4823-4825
The kinetics of the hydrogen-deuterium exchange reactions of poly(dA).poly(rU) and poly(rA).poly(dT) has been examined, at pH 7.0 and at various temperatures in the 15-35 degrees C range, by stopped-flow ultraviolet spectrophotometry. For comparison, the deuteration kinetics of poly[d(A-T)].poly[d(A-T)] and poly(rA).poly(rU) has been reexamined. At 20 degrees C, the imino deuteration (NH----ND) rates of the two hybrid duplexes were found to be 1.5 and 1.8 s-1, respectively. These are nearly equal to the imino deuteration rates of poly[d(A-T)].poly[d(A-T)] (1.1 s-1) and poly(rA).poly(rU) (1.5 s-1) but appreciably higher than that of poly(dA).poly(dT) (0.35 s-1). It has been suggested that a DNA.RNA hybrid, an RNA duplex, and the AT-alternating DNA duplex have in general higher base-pair-opening reaction rates than the ordinary DNA duplex. The amino deuteration (NH2----ND2) rates, on the other hand, have been found to be 0.25, 0.28, and 0.33 s-1, respectively, for poly(dA).poly(rU), poly(rA).poly(dT), and poly[d(A-T)].poly[d(A-T)], at 20 degrees C. These are appreciably higher than that for poly(rA).poly(rU) (0.10 s-1). In general, the equilibrium constants (K) of the base-pair opening are considered to be greatest for the DNA.RNA hybrid duplex (0.05 at 20 degrees C), second greatest for the RNA duplex (0.02 at 20 degrees C), and smallest for the DNA duplex (0.005 at 20 degrees C), although the AT-alternating DNA duplex has an exceptionally great K (0.07 at 20 degrees C). From the temperature effect on the K value, the enthalpy of the base-pair opening was estimated to be 3.0 kcal/mol for the DNA.RNA hybrid duplex.  相似文献   

7.
B C Sang  D M Gray 《Biochemistry》1987,26(23):7210-7214
Circular dichroism (CD) data indicated that fd gene 5 protein (G5P) formed complexes with double-stranded poly(dA.dT) and poly[d(A-T).d(A-T)]. CD spectra of both polymers at wavelengths above 255 nm were altered upon protein binding. These spectral changes differed from those caused by strand separation. In addition, the tyrosyl 228-nm CD band of G5P decreased more than 65% upon binding of the protein to these double-stranded polymers. This reduction was significantly greater than that observed for binding to single-stranded poly(dA), poly(dT), and poly[d(A-T)] but was similar to that observed for binding of the protein to double-stranded RNA [Gray, C.W., Page, G.A., & Gray, D.M. (1984) J. Mol. Biol. 175, 553-559]. The decrease in melting temperature caused by the protein was twice as great for poly[d(A-T).d(A-T)] as for poly(dA.dT) in 5 mM tris(hydroxymethyl)aminomethane hydrochloride (Tris-HCl), pH 7. Upon heat denaturation of the poly(dA.dT)-G5P complex, CD spectra showed that single-stranded poly(dA) and poly(dT) formed complexes with the protein. The binding of gene 5 protein lowered the melting temperature of poly(dA.dT) by 10 degrees C in 5 mM Tris-HCl, pH 7, but after reducing the binding to the double-stranded form of the polymer by the addition of 0.1 M Na+, the melting temperature was lowered by approximately 30 degrees C. Since increasing the salt concentration decreases the affinity of G5P for the poly(dA) and poly(dT) single strands and increases the stability of the double-stranded polymer, the ability of the gene 5 protein to destabilize poly(dA.dT) appeared to be significantly affected by its binding to the double-stranded form of the polymer.  相似文献   

8.
We have employed a variety of physical methods to study the equilibrium melting and temperature-dependent conformational dynamics of dA.dT tracts in fractionated synthetic DNA polymers and in well-defined fragments of kinetoplast DNA (kDNA). Using circular dichroism (CD), we have detected a temperature-dependent, "premelting" event in poly(dA).poly(dT) which exhibits a midpoint near 37 degrees C. Significantly, we also detect this CD "premelting" behavior in a fragment of kDNA. By contrast, we do not observe this "premelting" behavior in the temperature-dependent CD spectra of poly[d(AT)].poly[d(AT)], poly(dG).poly(dC), poly[d(GC)].poly[d(GC)], or calf thymus DNA. Thus, poly(dA).poly(dT) and kDNA exhibit a common CD-detected "premelting" event which is absent in the other duplex systems studied in this work. Furthermore, we find that the anomalous electrophoretic retardation of the kDNA fragments we have investigated disappears at temperatures above approximately 37 degrees C. We also observe that the rotational dynamics of poly(dA).poly(dT) and kDNA as assessed by singlet depletion anisotropy decay (SDAD) and electric birefringence decay (EBD) also display a discontinuity near 37 degrees C, which is not observed for the other duplex systems studied. Thus, in the aggregate, our static and dynamic measurements suggest that the homo dA.dT sequence element [common to both poly(dA).poly(dT) and kDNA] is capable of a temperature-dependent equilibrium between at least two helical states in a temperature range well below that required to induce global melting of the host duplex. We suggest that this "preglobal" melting event may correspond to the thermally induced "disruption" of "bent" DNA.  相似文献   

9.
The binding modes of three benzopyrido [4,3-b]indole derivatives (and one benzo[-f]pyrido [4-3b] quinoxaline derivative) with respect to double helical poly(dA) · poly(dT) and poly[d(A-T)]2 and triple-helical poly(dA) · 2poly(dT) have been investigated using linear dichroism (LD) and CD: (I) 3-methoxy-11-amino-BePI where BePI = (7H-8-methyl-benzo[e]pyrido [4,3-b]indole), (II) 3-methoxy-11-[(3′-amino) propylamino]-BePI, (III) 3-methoxy-7-[(3′-diethylamino)propylamino] BgPI where BgPI = (benzo[g]pyrido[4,3-b]indole), and (IV) 3-methoxy-11-[(3′-amino)propylamino] B f P Q where B f P Q = {benzo[-f]pyrido[4-3b]quinoxaline}. The magnitudes of the reduced LD of the electronic transitions of the polynucleotide bases and of the bound ligands are generally very similar, suggesting an orientation of the plane of the ligands' fused-ring systems preferentially perpendicular to the helix axis. The LD results suggest that all of the ligands are intercalated for all three polynucleotides. The induced CD spectrum of the BePI chromophore in the (II-BePI)-poly[d(A-T)]2 complex is almost a mirror image of that for the (I-BePI)-poly(dA) · poly(dT) and (I-BePI)-poly(dA) · 2poly(dT) complexes, suggesting an antisymmetric orientation of the BePI moiety upon intercalation in poly[d(A-T)]2 compared to the other polynucleotides. The induced CD of I-BePI bound to poly(dA) · 2poly(dT) suggests a geometry that is intermediate between that of its other two complexes. The concluded intercalative binding as well as the conformational variations between the different BePI complexes are of interest in relation to the fact that BePI derivatives are triplex stabilizers. © 1997 John Wiley & Sons, Inc. Biopoly 42: 101–111, 1997  相似文献   

10.
The pressure dependence of the helix–coil transition of poly(dA)∙poly(dT) and poly[d(A-T)]·poly[d(A-T)] in aqueous solutions of NaCl and CsCl at concentrations between 10 and 200 mM is reported and used to calculate the accompanying volume change. We also investigated the binding parameters and volume change of ethidium bromide binding with poly(dA)∙poly(dT) and poly[d(A-T)]·poly[d(A-T)] in aqueous solutions of these two salts. The volume change of helix–coil transition of poly(dA)∙poly(dT) in Cs+-containing solutions differs by less than 1 cm3 mol− 1 from the value measured when Na+ is the counter-ion. We propose that this insensitivity towards salt type arises if the counter-ions are essentially fully hydrated around DNA and the DNA conformation is not significantly altered by salt types. Circular dichroism spectroscopy showed that the previously observed large volumetric disparity for the helix–coil transition of poly[d(A-T)]·poly[d(A-T)] in solutions containing Na+ and Cs+ is likely result of a Cs+-induced conformation change that is specific for poly[d(A-T)]·poly[d(A-T)]. This cation-specific conformation difference is mostly absent for poly(dA)∙poly(dT) and EB bound poly[d(A-T)]·poly[d(A-T)].  相似文献   

11.
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   

12.
Summary The effect of Aclacinomycin B (ACM-B), an anthracycline antitumor antibiotic, on the DNA-dependent RNA synthesis using single- and double-stranded DNAs of known base content and sequence is studied. The data show that ACM-B effectively inhibits the double-stranded DNA-directed RNA synthesis with a preference of poly[d(A-T)] > poly[d(G-C)] > poly[d(I-C)]. In contrast, it has no inhibitory effect on the template function of single-stranded DNA (e.g. poly dA, poly dT, and poly dC). These results suggest that the mechanism of ACM-13 inhibition, like other anthracycline antibiotics, is by intercalation. In addition to the base specificity, there are also dramatic differences in inhibition depending on the base sequence in the DNA template. Thus, ACM-13 preferentially inhibits the alternating double-stranded copolymers over the double-stranded homopolymers; e.g. poly [d(A-T)] is inhibited to a greater extent than poly dA · poly dT and poly [d(G-C)] is inhibited more than poly dG · poly dC. Since the inhibition by ACM-13 can be totally abolished when assayed in excess amount of DNA, this result suggests that ACM-B inhibition of RNA synthesis is solely on the DNA template (which is in support of the intercalation model), and has ruled out the possibility that ACM-B may also exert an inhibitory effect on the activity of RNA polymerase per se.  相似文献   

13.
14.
Melting measurements of the mono-base-pair DNA polymers showed that the melting temperature Tm of the B-DNA homopolymer poly (dA ) · poly (dT) is higher than that of the copolymer poly [d(A-T)]. On the other hand, the Tmof the B-DNA homopolymer poly (dG) · poly (dC) is lower than that of the copolymer poly [d (G-C)]. From a structural point of view, the cross-strand base-stacking interaction in a DNA homopolymer is weaker than that in a DNA copolymer with the same base pair. One would then expect that all the DNA homopolymers are less stable than the copolymer with the same base pair. We find that the inversion of the melting order seen in the AT mono-base-pair DNA polymers is caused by the enhanced thermal stability of poly (dA) · poly (dT) from a well-defined spine of hydration attached to its minor groove. In this paper we employ the modified self-consistent phonon theory to calculate base-pair opening probabilities of four B-DNA polymers: poly(dA)-poly(dT), poly(dG) · poly(dC), poly[d(A-T)], and poly[d(G-C)] at temperatures from room temperature through the melting regions. Our calculations show that the spine of hydration can give the inverted melting order of the AT polymers as compared to the GC polymers in fair agreement with experimental measurements. Our calculated hydration spine disruption behavior in poly(dA) · poly(dT) at premelting temperatures is also in agreement with experimentally observed premelting transitions in poly (dA) · poly (dT). The work is in a sense a test of the validity of our models of nonbonded interactions and spine of hydration interactions. We find we have to develop the concept of a strained bond to fit observations in poly (dA) · poly(dT). The strained-bond concept also explains the otherwise anomalous stability of the hydration chain. © 1993 John Wiley & Sons, Inc.  相似文献   

15.
H P Vosberg  F Eckstein 《Biochemistry》1977,16(16):3633-3640
We have synthesized fd and phi X174DNA in the presence of 2'-deoxyadenosine 5'-O-(1-thiotriphosphate) (dATP alpha S) and the corresponding phosphorothioate derivatives of dCTP and dTTP using ether-permeabilized E. coli cells or crude cell extracts of E. coli DNA polymerase I. Reaction rates of enzymes involved in the formation or breakdown of DNA are decreased in the presence of phosphorothioates. The amount of label incorporated with [35S]dATP alpha S suggests that the dAMP has been completely substituted by 2'-deoxyadenosine 5'-0-phosphorothioate (dAMPS). The substituted DNAs have the same sedimentation coefficients, similar buoyant density, infectivity, and thermal stability as the unsubstituted DNAs. The procedure therefore allows specific modification at the 5' position of dA, dC, or dT in the DNA. In view of the recent demonstration of specific binding of Pt2+ complexes to the phosphorothioate analogue of poly[r(A-U)] (Strothkamp, K.G., and Lippard, S.J. (1976), Proc. Natl. Acad. Sci. U.S.A. 73, 2536), the synthesis of phosphorothioate containing DNA may be of use for DNA sequencing by electron microscopy.  相似文献   

16.
Binding to DNA of two synthetic peptides, Val-Thr-Thr-Val-Val-NH-NH-Dns and Thr-Val-Thr-Lys-Val-Gly-Thr-Lsy-Val-Gly-Thr-Val-Val-NH-NH-Dns (where Dns is a residue of 5-dimethylaminonaphthalene-1-sulfonic acid), has been studied by circular dichroism, electron microscopy and fluorescence methods. It has been found that these two peptides can self-associate in aqueous solution as follows from the fact that concentration-dependent changes are observed in the UV absorbance and fluorescence spectra. The two peptides can bind to DNA both in self-associated and monomeric forms. The pentapeptide in the beta-associated form binds more strongly to poly(dG).poly(dC) than to poly[d(A-C)].poly[d(G-T)] and poly(dA).poly(dT) whereas the tridecapeptide exhibits an opposite order of preferences binding more strongly to poly[d(A-C)].poly[d(G-T)] and poly(dA).poly(dT) than to poly(dG).poly(dC). Binding is a cooperative process which is accompanied by the DNA compaction at peptide/DNA base pair ratios greater than 1. At the initial stage of the compaction process, the coalescence of DNA segments covered by bound peptide molecules leads to the formation of DNA loops stabilized by the interaction between peptide molecules bound to different DNA segments. Further increase in the peptide/DNA ratio leads to the formation of rod-like structures each consisting of two or more double-stranded DNA segments. The final stage of the compaction process involves folding of fibrillar macromolecular complexes into a globular structure containing only one DNA molecule.  相似文献   

17.
The requirements of cofactor DNA for DNA-dependent ATPases B and C3 were analyzed in detail. ATPase B and C3 required the presence of a polynucleotide for their activities. Among the DNAs tested, ATPase B showed a preference for poly(dT) as its cofactor. The other deoxyhomopolymers, except poly(dG) and heat-denatured DNA also were effective. The alternating polydeoxyribonucleotide, poly[d(A-T)] had an efficiency 23% that of heat-denatured DNA. Unlike ATPase B, ATPase C3 showed almost no activity with deoxyhomopolymers. The most effective cofactor for ATPase C3 so far tested is poly[d(A-T)]. Relatively high activity was obtained with heat-denatured DNA. The high activity of ATPase B with poly(dT) was reduced by the addition of poly(dA). The addition of noncomplementary homopolymers did not affect enzyme activity. ATPase C3 activity in the presence of 10 microM poly(dT) increased gradually with concentrations of poly(dA) up to 20 microM, after which it decreased. Almost no increase in activity was observed when noncomplementary homopolymers were added. The relatively high activity of ATPase C3 with heat-denatured DNA was suggested by its high sensitivity to ethidium bromide to be due to the double-stranded region in the heat-denatured DNA formed by self-annealing.  相似文献   

18.
Four chromatographically distinct DNA-dependent ATPases, B, C1, C2, and C3, have been partially purified from mouse FM3A cell extracts. These ATPases are distinguished from each other by their physical and enzymological properties. DNA-dependent ATPases B, C1, C2, and C3 have sedimentation coefficients in 250 mM KCl of 5.5, 5.3, 7.3, and 3.4 S, respectively. ATPases B, C2, and C3 hydrolyze dATP as efficiently as ATP, whereas C1 does not. ATPase B hydrolyzes other ribonucleoside triphosphates with relatively high efficiency as compared to the other three enzymes. ATPase C3 prefers poly[d(A-T)] to poly(dT) as cofactor, whereas the other three enzymes prefer poly(dT) to poly[d(A-T)]. Among the four ATPases, ATPase C3 has been highly purified and characterized in detail. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the most purified fraction of ATPase C3 showed two major bands corresponding to molecular weights of 66 000 and 63 000. The Km values of the enzyme for ATP and dATP are 0.53 and 0.86 mM, respectively. As cofactor, poly[d(A-T)] is the most effective among the DNAs tested. Heat-denatured DNA and native DNA are also effective but used with less efficiency. Almost no or very little activity has been detected with ribohomopolymers and oligonucleotides. The activity attained with poly(dT) and poly(dA) is 11 and 6% of that with heat-denatured DNA, respectively. When both polymers were added at a molar ratio 1 to 1, very high activity was obtained with these polymers. On the other hand, little activity was observed by the combination of noncomplementary homopolymers such as poly(dT) and poly(dG).  相似文献   

19.
J E Herrera  J B Chaires 《Biochemistry》1989,28(5):1993-2000
Circular dichroism and UV absorbance spectroscopy were used to monitor and characterize a premelting conformational transition of poly(dA)-poly(dT) from one helical form to another. The transition was found to be broad, with a midpoint of tm = 29.9 degrees C and delta HVH = +19.9 kcal mol-1. The transition renders poly(dA)-poly(dT) more susceptible to digestion by DNase I and facilitates binding of the intercalator daunomycin. Dimethyl sulfoxide was found to perturb poly(dA)-poly(dT) structure in a manner similar to temperature. These combined results suggest that disruption of bound water might be linked to the observed transition. A thermodynamic analysis of daunomycin binding to poly(dA)-poly(dT) shows that antibiotic binding is coupled to the polynucleotide conformational transition. Daunomycin binding renders poly(dA)-poly(dT) more susceptible to DNase I digestion at low binding ratios, in contrast to the normal behavior of intercalators, indicating that antibiotic binding alters the conformation of the polynucleotide. The unusual thermodynamic profiles previously observed for the binding of many antibiotics to poly(dA)-poly(dT) can be explained by our results as arising from the coupling of ligand binding to the polynucleotide conformational transition. Our data further suggest a physical basis for the temperature dependence of DNA bending.  相似文献   

20.
We synthesized seven homologs of spermine (H2N(CH2)3NH(CH2)nNH(CH2)3NH2, where n = 2-9; n = 4 for spermine) and studied their effects on melting temperature (Tm), conformation, and precipitation of poly(dA).2poly(dT). The triplex DNA melting temperature, Tm1 was 34.4 degrees C in the presence of 150 mM KCl. Addition of spermine homologs increased Tm1 in a concentration-dependent and structure-dependent manner, with 3-6-3 (n = 6) exerting optimal stabilization. The dTm1/dlog[polyamine] values were 9-24 for these compounds. The duplex melting temperature, Tm2 was insensitive to homolog concentration and structure, suggesting their ability to stabilize triplex DNA without altering the stability of the underlying duplex. Circular dichroism spectral studies revealed psi-DNA formation in a concentration-dependent and structure-dependent manner. Phase diagrams were constructed showing the critical ionic/polyamine concentrations stabilizing different structures. These compounds also exerted structural specificity effects on precipitating triplex DNA. These data provide new insights into the ionic/structural determinants affecting triplex DNA stability and indicate that 3-6-3 is an excellent ligand to stabilize poly(dA).2poly(dT) triplex DNA under physiologic ionic conditions for antigene therapeutics.  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