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 15 毫秒
1.
270 MHz 1H NMR and theoretical studies indicate that the drug lucanthone forms intercalated complexes with the synthetic DNA poly(dA-dT)·poly(dA-dT). In the intercalated complex the long axis of the drug is perpendicular to the helix axis and parallel to the base pair axis, i.e., the long axis is perpendicular to the dyad axis.  相似文献   

2.
Core histones, (H2A,H2B,H3,H4)2, were reconstituted with the synthethic polynucleotides poly(dA-dT)·poly(dA-dT) and poly(dG-dC)·poly(dG-dC) to yield synthetic chromatins containing 200 basepairs per octamer. These synthetic chromatins displayed a 36% decrease in the circular dichroism (CD) peak ellipticity from the value of the polynucleotide free in solution; the poly(dA-dT)·poly(dA-dT)/chromatin showed an increase in the complexity of the thermal denaturation profile compared to that of the polynucleotide. Both the temperature of maximum dhdT for each transition (Tm) and the relative amount of poly(dA-dT)·poly(dA-dT) in the synthetic chromatin melting in each of the four thermal transitions is a function of the ionic strength over the 0–5 mM sodium phosphate range (0.25 mM EDTA, pH 7.0); a shift of material toward higher melting transitions was observed with increasing ionic strength. The CD peak ellipticity value for both synthetic chromatins was ionic strength-independent over the 0–5 mM sodium phosphate range. These results are in contrast to those observed with H1H5 stripped chicken erythrocyte chromatin (Fulmer, A. and Fasman, G.D. (1979) Biopolymers 18, 2875–2891), where an ionic strength dependence was found. Differences in the CD spectra between poly(dA-dT)·poly(dA-dT)/chromatin, poly(dG-dC)·poly(dG-dC)/chromatin and H1H5 stripped chicken erythrocyte chromatin suggest subtle differences in assembly. Finally, the temperature dependence of the CD spectra of poly(dA-dT)·poly(dA-dT)-containing synthetic chromatin, which is similar to that for the polynucleotide, suggests the core histone bound polynucleotide has a large degree of conformational flexibility allowing it to undergo the premelt transition.  相似文献   

3.
Abstract

Antibiotic netropsin is known to bind specifically to A and T regions in DNA; the mode of binding being non-intercalative. Obviously, H-bonding between the proton donors of netropsin and acceptors N3 of A and 02 of T comes as a strong possibility which might render this specificity. In netropsin there could be 8 proton donors: four terminal amino groups and four internal imino groups. However, methylation of the terminal amino groups does not alter the binding affinity of netropsin to DNA—but the modification of the internal imino groups significantly lowers the binding affinity. Hence, the logical conclusion is that netropsin may specifically interact with A and T through H-bonding and in order to do so, it should approach the helix from the minor groove. The present paper provides experimental data which verify the conclusion mentioned above.

Using poly(dA-dT)? poly(dA-dT) as a model system it was observed following a thorough theoretical stereochemical analysis that netropsin could bind to -(T-A-T) sequence of the polymer in the B-form through the minor groove by forming specific B-bonding. Models could be either right or left-handed B-DNA with a mono or dinucleotide repeat.

By monitoring the 31P signals of free poly(dA-dT) ? poly(dA-dT) and netropsin-poly(dA-dT)? poly(dA-dT) complex we show that the drug changes the DNA structure from essentially a mononucleotide repeat to that of very dominant dinucleotide repeat; however the base- pairing in the DNA-drug complex remain to be Watson-Crick. Whether H-bonding is the specific mode of interaction was judged by monitoring the imino protons of netropsin in the presence of poly(dA-dT) ? poly(dA-dT). This experiment was conducted in 90% H2O + 10% D2O Using the time-shared long pulse. It was found that exchangeable imino protons of netropsin appear in the drug-DNA complex and disappear upon increasing the D2O content; thus confirming that H-bonding is indeed the specific mode of interaction. From these and several NOE measurements, we propose a structure for poly(dA-dT)? poly(dA-dT(-netropsin complex.

In summary, experimental data indicate that netropsin binds to poly(dA-dT)? poly(dA-dT) by forming specific hydrogen bonds and that the binding interaction causes the structure to adopt a Watson-Crick paired dinucleotide repeat motif. The proposed hydrogen bonds can form only if the drug approaches the DNA from the minor groove. Within the NMR time scale the interaction between the ligand and DNA is a fast one. From the NOE experimental data, it appears that poly(dA-dT)? poly(dA-dT) in presence of netropsin exists as an equilibrium mixture of right- and left-handed B-DNA duplexes with a dinucleotide repeat—with a predominance of the left-handed form. The last conclusion is a soft one because it was very difficult to make sure the absence of spin diffusion. In a 400 base pairs long DNA duplex- drug complex (as used in this study), equilibrium between right and left-handed helices can also mean the existence of both helical domains in the same molecule with fast interchange between these domains or/and unhindered motion/propagation of these domains along the helix axis.  相似文献   

4.
Abstract

The article reviews data indicating that poly(dA-dT)?poly (dA-dT) is able of adopting three distinct double helical structures in solution, of which only the A form conforms to classical notions. The other two structures have dinucleotides as double helical repeats. At low salt concentrations poly(dA-dT)?poly(dA-dT) adopts a B-type alternating conformation which is exceptionally variable. Its architecture can gradually move in the limits demarcated by the CD spectra with inverted long wavelength CD bands and the 31P NMR spectra with a very low and a 0.6 ppm separation of two resonances. Contrary to Z-DNA, the 31P NMR spectrum of the limiting alternating B conformation of poly(dA-dT)?poly(dA-dT) is characterized by an upfield shift of one resonance. We attribute the exceptional conformational flexibility of the alternating B conformation to the unequal tendency of bases in the dA-dT and dT-dA steps to stack.

However, by assuming the limiting alternating B conformation, the variability of the synthetic DNA is not exhausted. Specific agents make it isomerize into another conformation by a fast, two-state mechanism, which is reflected by a further deepening of the negative long wavelength CD band and a downfield shift of the 31P NMR resonance of poly (dA-dT)?poly(dA-dT) that was constant in the course of the gradual alterations of the alternating B conformation. These changes are, however, qualitatively different from the way poly(dG-dC)?poly(dG-dC) behaves in the course of the B-Z isomerization. Poly(dG-dC) ?poly(dG-dC) displays purine-pyrimidine (dGpdC) resonance in the characteristic downfield position, while the downfield resonance of poly(dA-dT)?poly(dA-dT) belongs to the pyrimidine-purine (dTpdA) phosphodiester linkages. Consequently, phosphodiester linkages in the purine-pyrimidine steps play a similar role in the appearance of the Z form to the pyrimidine-purine phosphodiesters in the course of the isomerization of poly(dA-dT)?poly(dA-dT). This excludes that the high-salt structures of poly(dA-dT)?poly(dA-dT) and poly(dG-dC)?poly(dG-dC) are members of the same conformational family. We call the high-salt conformation of poly(dA-dT)?poly(dA-dT) X-DNA.

It furthermore follows from the review that synthetic molecules of DNA with alternating purine-pyrimidine sequences of bases can adopt either the Z form or the X form, or even both, depending on the environmental conditions. This introduces a new dimension into the DNA double helix conformational variability. The possible biological relevance of the X form is suggested by experiments with linear molecules of natural DNA. These indicate that Arich regions in natural DNAs can isomerize into the X form while the bulk of the molecule remains in the B form. The coexistence of both structures in a single DNA molecule may be understood in view of the favourable kinetic and thermodynamic properties with which the X form appears.  相似文献   

5.
Abstract

Poly(dA-dT) poly(dA-dT) structures in aqueous solutions with high NaCl concentrations and in the presence of Ni2+ ions have been studied with resonance Raman spectroscopy (RRS). In low water activity the effects of added 95 mM NiCl2 in solution stabilize the syn geometry of the purines and reorganize the water distribution via local interactions of Ni-water charged complexes with the adenine N7 position. It is shown that RRS provides good marker bands for a left-handed helix: i) a purine ring breathing mode around 630 cm″?1coupled to the deoxyribose vibration in the syn geometry, ii) a 1300-1340 cm?1 region characterizing local chemical interactions of the Ni2+ ions with the adenien N7 position, iii) lines at about 1483-and 1582 cm?1 correlated to the anti/syn reorientation of the adenine residues on B-Z structure transition, iv) marker bands of the thymidine carbonyl group couplings at 1680-and 1733 cm?1 due to the disposition of the thymidine residues in the Z helix specific geometry. Hence poly(dA-dT) poly(dA-dT) can adopt a Z form in solution. The Z form observed in alternate purine-pyrimidine sequences does not require G-C base pairs.  相似文献   

6.
X-ray diffraction analyses of fibers of polydeoxyadenylic acid · polydeoxythymidylic acid show that this molecule exists as a 10-fold double-helix with axial rise per nucleotide h = 3.24 to 3.29 A?. The structure is very similar to B-DNA (h = 3.37 A?) in having C3-exo furanose rings and base-pairs positioned centrally on the helix axis, but distinctive enough to have two packing modes, neither of which has been observed for B-DNA. Although the triple-stranded poly(dT) · poly(dA) · poly(dT) also has a large value of h(3.26 Å), each of the chains is a 12-fold helix of the A-genus with C3-endo furanose rings and bases displaced several Angstrom units from the helix axis.  相似文献   

7.
Abstract

Using CD measurements we show that the interaction of netropsin to poly(dA-dT)·poly(dA-dT) involves two binding modes at low ionic strength. The first and second binding modes are distinguished by a defined shift of the CD maximum and the presence of characteristic isodichroic points in the long wavelength range from 313 nm to 325 nm. The first binding mode is independent of ionic strength and is primarily determined by specific interaction to dA·dT base pairs. Employing a netropsin derivative and different salt conditions it is demonstrated that ionic contacts are essential for the second binding mode. Other alternating duplexes and natural DNA also exhibit more or less a second step in the interaction with netropsin observable at high ratio of ligand per nucleotide. The second binding mode is absent for poly(dA)·poly(dT). The presence of a two-step binding mechanism is also demonstrated in the complex formation of poly(dA-dT)·poly(dA-dT) with the distamycin analog consisting of pentamethylpyrrolecarboxamide. While the binding mode I of netropsin is identical with its localization in the minor groove, for binding mode II we consider two alternative interpretations.  相似文献   

8.
Isoquinoline alkaloids and their analogs represent an important class of molecules for their broad range of clinical and pharmacological utility. These compounds are of current interest owing to their low toxicity and excellent chemo preventive properties. These alkaloids can play important role in stabilising the nucleic acid triple helices. The present study has focused on the interaction of five 9-O-(ω-amino) alkyl ether berberine analogs with the DNA triplex poly(dT)·poly(dA)*poly(dT) and the parent duplex poly(dA)·poly(dT) studied using various biophysical techniques. Scatchard analysis of the spectral data indicated that the analogs bind both to the duplex and triplex in a non-cooperative manner in contrast to the cooperative binding of berberine to the DNA triplex. Strong intercalative binding to the DNA triplex structure was revealed from ferrocyanide quenching, fluorescence polarization and viscosity results. Thermal melting studies demonstrated higher stabilization of the Hoogsteen base paired third strand of the DNA triplex compared to the Watson–Crick strand. Circular dichroism studies suggested a stronger perturbation of the DNA triplex conformation by the alkaloid analogs compared to the duplex. The binding was entropy-driven in each case and the entropy contribution to free energy increased as the length of the alkyl side chain increased. The analogs exhibited stronger binding affinity to the triple helical structure compared to the parent double helical structure.  相似文献   

9.
When the model polynucleotide poly(dG-dC)?poly(dG-dC) [polyGC] is titrated with a strong acid (HCl) in unbuffered aqueous solutions containing the chlorides of the alkali metals in the concentration range 0.010?M-0.600?M, two transitions in the absorbance vs. pH plots are evidenced, characterized by the constants pK(a(?)) and pK(a(?)). The limiting values at infinite saline concentrations of these two constants, namely pK(∞)(a(?)) and pK(∞)(a(?)) obtained making use of the "one site saturation constant" equation or, in turn, of the double logarithmic plot: pK(a) vs. log([salt]?1), exhibit a clear dependence on the nature of the cations. The effects of the different alkali cations on the pK(∞)(a) values follow the Hofmeister series. In fact, the pK(∞)(a(?)) and the pK(∞)(a(?)) values are smaller for Li+ and Na+ than for Rb+ and Cs+, with K+ at the border between the two, showing that the transitions require higher concentrations of protons to occur in the presence of high concentrations of the cosmotropic ions.  相似文献   

10.
Abstract

CD spectra of poly(dA-dT)· poly(dA-dT) in low salt (10–100 mM NaCl) and high salt (4–6 M CsF) are different i.e. 275 nm band gets inverted in going from low to high salt (Vorhickova et. al.MarJ. Mol. Biol. 166, 85, 1983). However, from CD spectra alone it is not possible to decipher any structural differences that might exist between the low and high salt forms of poly(dA-dT)? poly(dA-dT). Hence, we took recourse to high resolution NMR spectroscopy to understand the structural properties of poly(dA-dT)? poly(dA-dT) in low and high salt. A detailed analysis of shielding constants and extensive use of NOE studies under minimum spin diffusion conditions using C(8)-deuterated poly(dA-dT)? poly(dA-dT) enabled us to come up with the following conclusions (i) base-pairing is Watson-Crick under low and high salt conditions, (ii) under both the conditions of salt the experimental data can be explained in terms of an equilibrium blend of right and left-handed B-DNA duplexes with the left-handed form 70% and the right-handed 30%. In a 400 base pairs long poly(dA-dT)? polyidA-dT) (as used in this study), equilibrium between right and left-handed helices can also mean the existence of both helical domains in the same molecule with fast interchange between these domains or/and unhindered motion/propagation of these domains along the helix axis, (iii) However, there are other structural differences between the low and high salt forms of poly(dA-dT) ? poly(dA-dT); under the low salt condition, right-and left-handed B-DNA duplexes have mononucleotide as a structural repeat while under the high salt conditions, right-and left-handed B-DNA duplexes have dinucleotide as a structural repeat. In the text we provide the listing of torsion angles for the low and high salt structural forms, (iv) Salt (CsF) induced structural transition in poly(dA-dT)? poly(dA-dT) occurs without any breakage of Watson- Crick pairing, (v) The high salt form of poly(dA-dT)? poly(dA-dT) is not the left-handed Z-helix.

Although the results above from NMR data are quite unambiguous, a question still remains i.e. what does the salt (CsF) induced change in the CD spectra of poly(dA-dT)? poly(dA-dT) really indicate? Interestingly, we could show that the salt (CsF) induced change in poly(dA-dT)? poly(dA-dT) is quite similar to that caused by a basic polypeptide viz. poly-L(Lys2-Ala)n i.e. both the agents induced a ψ-structure in DNA. And it was also demonstrated that the changes in poly(dA-dT)? poly(dA-dT) as caused by CsF and poly-L-(Lys2-Ala)n could be reverted back by ethidium bromide-a relaxing agent.

To minimize complications from spin diffusion in this study we have used very small presaturation pulse lengths and C(8)-deuterated poly(dA-dT)? poly(dA-dT) of 400 ± 150 bp long. Even though deuteration of a primary site of diffusion such as C(8)H substantially decreases diffusion, in order to make sure that our conclusions are not compromised by possible diffusion in such a long fragment under small presaturation times, we have repeated our experiments using the six base pair long duplex of d(A-T-A-T-A-T) and found the results to be strikingly similar to that from the polymer.  相似文献   

11.
The influence of Cd2+ ions on the conformational equilibrium of single-stranded (poly(U), poly(A), poly(I)) and triple-stranded polyribonucleotides (A2I, A2U) in aqueous solutions (0.1 M Na+ pH 7) has been investigated using difference UV spectroscopy and thermal denaturation. Analysis of the shape and intensity of the DUV spectra of poly(A), poly(I), and A2I has revealed the presence of two types of complex formed as a result of (i) interaction between Cd2+ and the N7 atoms of purines, producing macrochelates; and (ii) binding of Cd2+ to the N1 atoms of poly(A) and poly(I). Since Cd2+ ions are not bound to heteroatoms of the bases in A2U, the conformation of the structure remains stable up to 0.02 M Cd2+. There is a critical Cd2+ concentration (~1.5?10?4 M) above which A2I assumes a new helical conformation with lower thermal stability. It is supposed that, upon the formation of the “metallized” A2I triplex, the Cd2+ ions are located inside the triple helix and form bridges between the hypoxanthine and adenine of the homopolynucleotide strands.  相似文献   

12.
13.
The binding properties of [RuL2(mip)]2+ {where L is 1,10-phenanthroline (phen) or 4,7-dimethyl-1,10-phenanthrollne (4,7-dmp) and mip is 2′-(3″,4″-methylenedioxyphenyl)imidazo[4′,5′-f][1,10]phenanthroline} with regard to the triplex RNA poly(U)·poly(A)*poly(U) were investigated using various biophysical techniques and quantum chemistry calculations. In comparison with [Ru(4,7-dmp)2(mip)]2+, remarkably higher binding affinity of [Ru(phen)2(mip)]2+ for the triplex RNA poly(U)·poly(A)*poly(U) was achieved by changing the ancillary ligands. The stabilization of the Hoogsteen-base-paired third strand was improved by about 10.9 °C by [Ru(phen)2(mip)]2+ against 6.6 °C by [Ru(4,7-dmp)2(mip)]2+. To the best of our knowledge, [Ru(phen)2(mip)]2+ is the first metal complex able to raise the third-strand stabilization of poly(U)·poly(A)*poly(U) from 37.5 to 48.4 °C. The results reveal that the ancillary ligands have an important effect on third-strand stabilization of the triplex RNA poly(U)·poly(A)*poly(U) when metal complexes contain the same intercalative ligands.  相似文献   

14.
Abstract

Poly(dA-dT)?poly(dA-dT) can adopt the B- and D- forms in the fibrous state. Theoretical energy calculations and fiber diffraction analyses suggest that there can be three structural models of poly(dA-dT)?poly(dA-dT) in each of these two forms viz right and left-handed Watson Crick models and left-handed Hoogsteen—a total of six possible models. Fiber data for the polymer in the B- or the D-form or energy calculations cannot distinguish any one model from the other. However, a comparison of observed proton chemical shifts with the theoretically computed ones and the NOE studies on exchangeable and nonexchangeable protons suggest that poly(dA-dT)?poly(dA-dT) in low salt solution exists predominantly in the left-handed B-conformation.  相似文献   

15.
Homogeneously purified poly(ADP-ribose) polymerase (PARP) specifically stimulated the activity of immunoaffinity-purified calf or human DNA polymerase by about 6 to 60-fold. Apparently, poly(ADP-ribosyl)ation of DNA polymerase was not necessary for the stimulation. The effects of PARP on DNA polymerase were biphasic: at very low concentrations of DNA, it rather inhibited its activity, whereas, at higher DNA concentrations, PARP greatly stimulated it. The autopoly(ADP-ribosyl)ation of PARP suppressed both its stimulatory and inhibitory effects. By immunoprecipitation with an anti-DNA polymerase antibody, it was clearly shown that PARP may be physically associated with DNA polymerase . Stimulation of DNA polymerase may be attributed to the physical association between the two, rather than to the DNA-binding capacity of PARP, since the PARP fragment containing only the DNA binding domain showed little stimulatory activity. The existence of PARP-DNA polymerase complexes were also detected in crude extracts of calf thymus.  相似文献   

16.
Streptomyces albulus PD-1 can co-produce antimicrobial homo-polymers poly(ε-lysine) (ε-PL) and poly(l-diaminopropionic acid) (PDAP). In this study, a novel feeding strategy of citric acid coupled with glucose-(NH4)2SO4 feeding was employed to S. albulus PD-1. When the pH of the culture broth dropped to 4.0, the feeding solution was added continuously to maintain the concentrations of glucose and citric acid at 10 and 4 g L?1, respectively. As a result, the final concentration of ε-PL increased from 21.7 to 29.7 g L?1 and the final concentration of PDAP decreased from 4.8 to 3.2 g L?1. Assays on intracellular nucleotide levels and key enzyme activities were performed to elucidate the underlying regulation mechanism. The addition of citric acid increased NADH/NAD+ ratio and decreased intracellular ATP level; meanwhile, the activities of pyruvate kinase, citrate synthase and isocitrate dehydrogenase decreased while aspartate aminotransferase activity increased. Therefore, we deduced that citric acid feeding resulted in metabolic flux redistribution at the node of phosphoenolpyruvate; the metabolic pathway from phosphoenolpyruvate directed into tricarboxylic acid cycle was weakened and thus PDAP production was inhibited. On the other hand, the metabolic pathway from phosphoenolpyruvate directed into oxaloacetate and l-aspartate was enhanced, thereby improving ε-PL production. This fermentation strategy may be potentially useful in ε-PL production because it can effectively inhibit the formation of by-products, such as PDAP.  相似文献   

17.
Poly(ethylene glycol)-b-poly(γ-benzyl L-glutamate)s bearing the disulfide bond (PEG-SS-PBLGs), which is specifically cleavable in intracellular compartments, were prepared via a facile synthetic route as a potential carrier of camptothecin (CPT). Diblock copolymers with different lengths of PBLG were synthesized by ring-opening polymerization of benzyl glutamate N-carboxy anhydride in the presence of a PEG macroinitiator (PEG-SS-NH(2)). Owing to their amphiphilic nature, the copolymers formed spherical micelles in an aqueous condition, and their particle sizes (20-125 nm in diameter) were dependent on the block length of PBLG. Critical micelle concentrations of the copolymers were in the range 0.005-0.065 mg/mL, which decreased as the block length of PBLG increased. CPT, chosen as a model anticancer drug, was effectively encapsulated up to 12 wt % into the hydrophobic core of the micelles by the solvent casting method. It was demonstrated by the in vitro optical imaging technique that the fluorescence signal of doxorubicin, quenched in the PEG-SS-PBLG micelles, was highly recovered in the presence of glutathione (GSH), a tripeptide reducing disulfide bonds in the cytoplasm. The micelles released CPT completely within 20 h under 10 mM GSH, whereas only 40% of CPT was released from the micelles in the absence of GSH. From the in vitro cytotoxicity test, it was found that CPT-loaded PEG-SS-PBLG micelles showed higher toxicity to SCC7 cancer cells than CPT-loaded PEG-b-PBLG micelles without the disulfide bond. Microscopic observation demonstrated that the disulfide-containing micelle could effectively deliver the drug into nuclei of SCC7 cells. These results suggest that PEG-SS-PBLG diblock copolymer is a promising carrier for intracellular delivery of CPT.  相似文献   

18.
Poly(lactide-co-glycolide) (PLGA) is extensively used in pharmaceutical applications, for example, in targeted drug delivery, because of biocompatibility and degradation rate, which is easily tuned by the copolymer composition. Nevertheless, synthesis of sugar-labeled amphiphilic copolymers with a PLGA backbone is quite a challenge because of high sensitivity to hydrolytic degradation. This Article reports on the synthesis of a new amphiphilic copolymer of PLGA grafted by mannosylated poly(ethylene oxide) (PEO). A novel building block, that is, α-methoxy-ω-alkyne PEO-clip-N-hydroxysuccinimide (NHS) ester, was prepared on purpose by photoreaction of a diazirine containing molecular clip. This PEO block was mannosylated by reaction of the NHS ester groups with an aminated sugar, that is, 2-aminoethyl-α-d-mannopyroside. Then, the alkyne ω-end-group of PEO was involved in a copper alkyne- azide coupling (CuAAC) with the pendent azides of the aliphatic copolyester. The targeted mannose-labeled poly(lactide-co-glycolide-co-ε-caprolactone)-graft-poly(ethylene oxide) copolymer was accordingly formed. Copolymerization of d,l-lactide and glycolide with α-chloro-ε-caprolactone, followed by substitution of chlorides by azides provided the azido-functional PLGA backbone. Finally, micelles of the amphiphilic mannosylated graft copolymer were prepared in water, and their interaction with Concanavalin A (ConA), a glyco-receptor protein, was studied by quartz crystal microbalance. This study concluded to the prospect of using this novel bioconjugate in targeted drug delivery.  相似文献   

19.
poly(1)·poly(C)-滤纸是一种亲和材料,可以用来吸附与双链核酸有亲和力的酶或蛋白。本文介绍用对-β硫酸酯乙砜基苯胺为活化剂制备poly(I)·poly(C)-滤纸的方法。poly(I)·poly(C)的结合容量为10—35μg/cm~2,用来吸附兔网织红细胞裂解液中2’-5’A合成酶效果良好。在一定范围内,酶活与被吸附裂解液量呈线性关系,说明可以用来定量检测未知样品中与poly(I)·poly(C)有亲和力的酶。poly(I)·poly(C)-滤纸在-20℃保存四个月亲和能力不变。本方法与文献报道的方法相比,操作简便试剂易得。  相似文献   

20.
Monomeric bis-benzimidazole (MB) is an analog of the Hoechst 33258 dye. The enthalpy and entropy of MB binding were evaluated by analyzing the calorimetric data on MB reverse titration with poly(dA) · poly(dT). A mathematical model was developed to estimate the thermodynamic parameters of binding on the basis of calorimetric data. The results agree well with spectrophotometric data on the binding of analogous compounds. The model was used to estimate the parameters of binding with poly(dA) · poly(dT) for dimeric bis-benzimidazole (DB), which consists of two bis-benzimidazole monomers linked via a flexible chain. The ligand was assumed to produce different types of complexes with the polymer.  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