首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The deoxyribonucleic acid (DNA) binding characteristics of adriamycin and several new anthracycline glycosides, including marcellomycin, aclacinomycin, rudolfomycin, musettamycin, and pyrromycin, have been studied. The fluorescence spectra were determined for all six anthracyclines, and the fluorescence quenching effects caused by interactions with the natural DNAs poly(dAdT)--poly(dAdT) and poly(dGdC) were characterized. Binding parameters were determined by Scatchard analyses of results obtained by spectrofluorometric titrations of anthracyclines with DNA. Consistent with earlier structure--activity relationship studies of nucleic acid synthesis inhibitory effects, the results demonstrate a correlation between the length of the glycosidic side chain and DNA binding affinity. In addition, the sugar residue 2-deoxyfucose appears to confer greater DNA binding ability than do the sugars rednosamine and cinerulose when present in the terminal position of the glycosidic side chain, also in agreement with earlier studies. The sequence preference of anthracycline--DNA interaction has been examined by using DNAs of varying GC content, including the naturally occurring calf thymus DNA (43% GC), Clostridium perfringens DNA (28% GC), and Micrococcus luteus DNA (72% GC) and the synthetic double-stranded copolymers poly(dGdC)--poly(dGdC) and poly(dAdT)--POLY(DAdT). The results demonstrate that although adriamycin shows an absolute requirement for GC sequences for DNA binding, marcellomycin and its analogues showed no such sequence requirement. Furthermore, an AT preference for DNA binding was demonstrated with marcellomycin and its analogues.  相似文献   

2.
Equilibrium binding is believed to play an important role in directing the subsequent covalent attachment of many carcinogens to DNA. We have utilized UV spectroscopy to examine the non-covalent interactions of aflatoxin B1 and B2 with calf thymus DNA, poly(dAdT):poly(dAdT), and poly(dGdC):poly(dGdC), and have utilized NMR spectroscopy to examine non-covalent interactions of aflatoxin B2 with the oligodeoxynucleotide d(ATGCAT)2. UV-VIS binding isotherms suggest a greater binding affinity for calf thymus DNA and poly(dAdT):poly(dAdT) than for poly(dGdC):poly(dGdC). Scatchard analysis of aflatoxin B1 binding to calf thymus DNA in 0.1 M NaCl buffer indicates that binding of the carcinogen at levels of bound aflatoxin less than 1 carcinogen per 200 base pairs occurs with positive cooperativity. The cooperative binding effect is dependent on the ionic strength of the medium; when the NaCl concentration is reduced to 0.01 M, positive cooperativity is observed at carcinogen levels less than 1 carcinogen per 500 base pairs. The Scatchard data may be fit using a "two-site" binding model [L.S. Rosenberg, M.J. Carvlin, and T.R. Krugh, Biochemistry 25, 1002-1008 (1986)]. This model assumes two independent sets of binding sites on the DNA lattice, one a high affinity site which binds the carcinogen with positive cooperativity, the second consisting of lower affinity binding sites to which non-specific binding occurs. NMR analysis of aflatoxin B2 binding to d(ATGCAT)2 indicates that the aflatoxin B2/oligodeoxynucleotide complex is in fast exchange on the NMR time scale. Upfield chemical shifts of 0.1-0.5 ppm are observed for the aflatoxin B2 4-OCH3, H5, and H6a protons. Much smaller chemical shift changes (less than or equal to 0.06 ppm) are observed for the oligodeoxynucleotide protons. The greatest effect for the oligodeoxynucleotide protons is observed for the adenine H2 protons, located in the minor groove. Nonselective T1 experiments demonstrate a 15-25% decrease in the relaxation time for the adenine H2 protons when aflatoxin B2 is added to the solution. This result suggests that aflatoxin B2 protons in the bound state may be in close proximity to these protons, providing a source of dipolar relaxation. Further experiments are in progress to probe the nature of the aflatoxin B1 and B2 complexes with polymeric DNA and oligodeoxynucleotides, and to establish the relationship between the non-covalent DNA-carcinogen complexes observed in these experiments, and covalent aflatoxin B1-guanine N7 DNA adducts.  相似文献   

3.
The high-mobility-group (HMG) chromosomal protein wheat HMGa was purified to homogeneity and tested for its binding characteristics to double-stranded DNA. Wheat HMGa was able to bind to P268, an A/T-rich fragment derived from the pea plastocyanin gene promoter, producing a small mobility shift in gel retardation assays where the bound complex was sensitive to addition of proteinase K but resistant to heat treatment of the protein, consistent with the identity of wheat HMGa as a putative HMG-I/Y protein. Gel retardation assays and southwestern hybridization analysis revealed that wheat HMGa could selectively interact with the DNA polynucleotides poly(dA).poly(dT), poly(dAdT).poly(dAdT), and poly(dG).poly(dC), but not with poly(dGdC).poly(dGdC). Surface plasmon resonance analysis determined the kinetic and affinity constants of sensor chip-immobilized wheat HMGa for double-stranded DNA 10-mers, revealing a good affinity of the protein for various dinucleotide combinations, except that of alternating GC sequence. Thus contrary to prior reports of a selectivity of wheat HMGa for A/T-rich DNA, the protein appears to be able to interact with sequences containing guanine and cytosine residues as well, except where G/C residues alternate directly in the primary sequence.  相似文献   

4.
D E Graves  T R Krugh 《Biochemistry》1983,22(16):3941-3947
Phase partition techniques have been used to measure the binding of the antitumor drugs adriamycin (NSC-123127) and daunorubicin (NSC-82151) to various DNAs. These methods provide reliable equilibrium binding data at the low levels of drug binding that may be expected in vivo. Both adriamycin and daunorubicin exhibit positive cooperativity (and/or allosterism) in their equilibrium binding to DNA as indicated by the positive slope in the initial region of the binding isotherms (Scatchard plots) under conditions simulating physiological ionic strengths. The cooperative binding (i.e., the appearance of initial positive curvature in the binding isotherms) is dependent upon the ionic strength, which suggests a role for DNA flexibility in the cooperative binding process. An analysis of the slope of the initial portion of the binding isotherms for the interaction of adriamycin with synthetic deoxypolynucleotides shows that the degree of cooperative binding decreases in the order poly(dGdT) X poly(dAdC) greater than or equal to poly(dAdT) X poly(dAdT) greater than poly(dGdC) X poly(dGdC). Marky and Breslauer [Marky, L.A., & Breslauer, K. J. (1982) Biopolymers 21, 2185-2194] found that the average base stacking enthalpies of these synthetic poly-nucleotides were in the same order, which also suggests that the properties of the DNA influence the cooperative binding (and/or allosteric effects). Adriamycin binds with a higher degree of cooperativity than daunorubicin (0.1 M NaCl); although this correlates with the effectiveness of the drugs as antitumor agents, the exact relationship between the observation of cooperative binding and pharmacological activity is yet to be determined.  相似文献   

5.
The regulatory roles fulfilled by polyamines by governance of chromatin structure are made possible by their strong association with cellular DNA, and hence by their ability to modulate DNA structure and function. Towards this end, it is crucial to understand the manifestation of sequence-dependent polyamine binding at the secondary and tertiary structural levels of DNA. This study utilizes circular dichroism (CD) and isothermal titration calorimetry (ITC) to address this relationship by using 20bp oligonucleotides with sequences-poly(dA):poly(dT), poly(dAdT):poly(dAdT), poly(dG):poly(dC), poly(dGdC):poly(dGdC)-that yield physiologically relevant structures, and poly(dIdC):poly(dIdC). CD studies show that at physiological ionic strength (150mM NaCl), spermine preferentially stabilizes A-tracts, and increases flexibility of the G-tract oligomer; the latter is also suggested by the larger change in entropy (DeltaS) of spermine binding to G-tracts. Given the chromatin destabilizing property of these sequences, these findings suggest a role for spermine in stabilization of non-nucleosomal A-tracts, and a compensating mechanism for incorporation of G-tracts in the chromatin structure. Other implications of these findings in sequence dependent DNA packaging are discussed.  相似文献   

6.
Induced CD of DNA intercalators: electric dipole allowed transitions   总被引:3,自引:0,他引:3  
The induced CD of an electric dipole allowed transition of a DNA intercalator has been calculated using the “matrix method” and a set of DNA ππ* transitions recently adopted for calculating the CD of DNA by Rizzo and Schellman [(1984) Biopolymers 23 , 435–470]. The induced CD is strongly dependent on the angular orientation of the intercalator and only moderately dependent on its location within the intercalation pocket. The dependence of the CD on the orientation and location of the intercalator was studied for some representative conformations of di- and tetranucleotide duplexes of (dGdC) and (dAdT). The effect of alternative DNA transition moment directions was also tested. The orientation dependence compares well with the previously predicted 1-2 cos2 γ dependence [B. Nordén and F. Tjerneld (1982) Biopolymers 21 , 1713–1734]. Measured induced CD spectra of methylene blue (MB) intercalated in double-stranded poly(dAdT), poly(dGdC), and calf-thymus DNA are discussed on the basis of the results of the calculations. Major differences between the induced CD spectra are likely to reflect different modes of intercalation for the different nucleotide sequences. In particular, the concluded geometry in solution for MB intercalated in poly(dAdT) differs significantly from the corresponding geometry found in dinucleotide–intercalator crystals.  相似文献   

7.
Abstract

Equilibrium binding is believed to play an important role in directing the subsequent covalent attachment of many carcinogens to DNA. We have utilized UV spectroscopy to examine the non-covalent interactions of aflatoxin B1 and B2 with calf thymus DNA, poly(dAdT):poly(dAdT), and poly(dGdC):poly(dGdC), and have utilized NMR spectroscopy to examine non-covalent interactions of aflatoxin B2 with the oligodeoxynucleotide d(ATGCAT)2. UV-VIS binding isotherms suggest a greater binding affinity for calf thymus DNA and poly(dAdT):poly(dAdT) than for poly(dGdC):poly(dGdC). Scatchard analysis of aflatoxin B1 binding to calf thymus DNA in 0.1 M NaCl buffer indicates that binding of the carcinogen at levels of bound aflatoxin ? 1 carcinogen per 200 base pairs occurs with positive cooperativity. The cooperative binding effect is dependent on the ionic strength of the medium; when the NaCl concentration is reduced to 0.01 M, positive cooperativity is observed at carcinogen levels ? 1 carcinogen per 500 base pairs. The Scatchard data may be fit using a “two-site” binding model [L.S. Rosenberg, M J. Carvlin, and T.R. Krugh, Biochemistry 25, 1002–1008 (1986)]. This model assumes two independent sets of binding sites on the DNA lattice, one a high affinity site which binds the carcinogen with positive cooperativity, the second consisting of lower affinity binding sites to which non-specific binding occurs. NMR analysis of aflatoxin B2 binding to d(ATGCAT)2 indicates that the aflatoxin B2/oligodeoxynucleotide complex is in fast exchange on the NMR time scale. Upfield chemical shifts of 0.1–0.5 ppm are observed for the aflatoxin B2 4-OCH3, H5, and H6a protons. Much smaller chemical shift changes ? 0.06 ppm) are observed for the oligodeoxynucleotide protons. The greatest effect for the oligodeoxynucleotide protons is observed for the adenine H2 protons, located in the minor groove. Nonselective T1 experiments demonstrate a 15–25 % decrease in the relaxation time for the adenine H2 protons when aflatoxin B2 is added to the solution. This result suggests that aflatoxin B2 protons in the bound state may be in close proximity to these protons, providing a source of dipolar relaxation. Further experiments are in progress to probe the nature of the aflatoxin B1 and B2 complexes with polymeric DNA and oligodeoxynucleotides, and to establish the relationship between the non-covalent DNA-carcinogen complexes observed in these experiments, and covalent aflatoxin B1,-guanine N7 DNA adducts.  相似文献   

8.
A series of thio- and selenopyrylium analogues of 2,4-di(4-dimethylaminophen-yl)-6-methylthiopyrylium iodide were prepared in five steps from 4-dimethylaminophenyl-propargyl aldehyde and the corresponding lithium acetylide. When bound to DNA, all of the dyes absorb at wavelengths >600nm, which avoids the hemoglobin band I maximum at 575nm. The binding of the series of dyes to double-stranded DNA was examined spectrophotometrically and by isothermal titration calorimetry to determine binding constants, by a topoisomerase I DNA unwinding assay, by competition dialysis with [poly(dGdC)](2) and [poly(dAdT)](2), and by ethidium bromide displacement studies to examine propensities for intercalation, and by circular dichroism studies. The dyes were found to show mixed binding modes.  相似文献   

9.
Chicken erythrocyte inner histones (H2A, H2B, H3 and H4) were associated with the two complementary homopolymeric polydeoxyribonucleotides and the two alternating copolymeric polydeoxyribonucleotides. No evidence for formation of chromatin-like structures was obtained for the complexes with poly(dG) . poly(dC) or poly(dA) . poly(dT). Both poly (dGdC) . poly(dGdC) and poly(dAdT) . poly(dAdT) could be folded by histones to yield material digested by DNAase I to multiples of about 10 and by staphylococcal nuclease to 146 bp core particles. Due to the lack of sequence heterogeniety in the complex of histones with poly(dAdT) . poly(dAdT), core particles with remarkable fine structural detail are obtained. The internal organization of DNA in the AT-containing and GC-containing core particles appears not to be identical.  相似文献   

10.
Thermodynamics of HMGB1 interaction with duplex DNA   总被引:4,自引:0,他引:4  
Müller S  Bianchi ME  Knapp S 《Biochemistry》2001,40(34):10254-10261
The high mobility group protein HMGB1 is a small, highly abundant protein that binds to DNA in a non-sequence-specific manner. HMGB1 consists of 2 DNA binding domains, the HMG boxes A and B, followed by a short basic region and a continuous stretch of 30 glutamate or aspartate residues. Isothermal titration calorimetry was used to characterize the binding of HMGB1 to the double-stranded model DNAs poly(dAdT).(dTdA) and poly(dGdC).(dCdG). To elucidate the contribution of the different structural motifs to DNA binding, calorimetric measurements were performed comparing the single boxes A and B, the two boxes plus or minus the basic sequence stretch (AB(bt) and AB), and the full-length HMGB1 protein. Thermodynamically, binding of HMGB1 and all truncated constructs to duplex DNA was characterized by a positive enthalpy change at 15 degrees C. From the slopes of the temperature dependence of the binding enthalpies, heat capacity changes of -0.129 +/- 0.02 and -0.105 +/- 0.05 kcal mol(-1) K(-1) were determined for box A and full-length HMGB1, respectively. Significant differences in the binding characteristics were observed using full-length HMGB1, suggesting an important role for the acid tail in modulating DNA binding. Moreover, full-length HMGB1 binds differently these two DNA templates: binding to poly(dAdT).(dTdA) was cooperative, had a larger apparent binding site size, and proceeded with a much larger unfavorable binding enthalpy than binding to poly(dGdC).(dCdG).  相似文献   

11.
At submicromolar concentrations two novel phenanthridine biguanides exhibit distinctly different spectroscopic signals for dGdC and dAdT sequences, respectively, by opposite fluorimetric changes (quenching for dGdC and increase for dAdT) and especially the bis-biguanide derivative gives an opposite ICD response (negative ICD for dGC and strong positive for dAdT). This specific signalling was explained by the ability of compounds to switch the binding mode from intercalation into dGdC to minor groove binding into dAdT sequences. Both compounds bind to rArU by intercalation, yielding different fluorimetric and CD response in comparison to any of aforementioned ds-DNA. Moreover, both compounds revealed significantly higher affinity toward ds-polynucleotides in comparison to previously studied alkylamine- and urea-analogues. Furthermore, DNA/RNA binding properties of novel compounds could be controlled by pH, due to the protonation of heterocyclic nitrogen. Low in vitro cytotoxicity of both compounds against human cell lines makes them interesting spectrophotometric probes.  相似文献   

12.
Resonance Raman, NMR, and visible spectroscopies, as well as viscosity and equilibrium dialysis studies were used to assess the effect of the N-alkyl substituent of meso-tetrakis(4-N-alkylpyridinium-4-yl)porphyrin cations on DNA binding. The DNAs studied include the native DNA, calf thymus DNA (CT DNA), the synthetic polynucleotides [poly(dGdC)]2 and [poly(dAdT)]2, and the oligonucleotide d(TATACGTATA)2. Both the porphyrins and the metalloporphyrins containing Ni(II) were examined with the N-alkyl = propyl (TPrpyP(4) and NiTPrpyP(4)) and 2-hydroxyethyl (TEtOHpyP(4) and NiTEtOHpyP(4)). The results were compared to those from the parent porphyrins with the N-methyl substituent (TMpyP(4) and NiTMpyP(4)). For almost all the comparisons made, the new porphyrin cations gave results very similar to those for the TMpyP(4) species. The resonance Raman study indicated that for the three DNA polymers all the Ni species were in the four-coordinate form when bound to all three polymers. It is suggested that both TPrpyP(4) and TEtOHpyP(4) bind to GC regions of DNA in the same intercalative manner as TMpyP(4) with the N-alkyl substituent extended into the solvent. For AT regions of DNA, the binding of TPrpyP(4) and TEtOHpyP(4) is nonintercalative, as found previously for TMpyP(4). The NiPrpy(4) and NiTEtOHpyP(4) cations bind to these polymers in a similar manner to the apo-porphyrins. The similar Raman spectral changes for the three Ni porphyrins upon addition of [poly(dAdT)]2 suggest that partial intercalation is not occurring because models indicate that it would be difficult to accommodate the bulkier N-alkyl substituents.  相似文献   

13.
We report the relative molar sound velocity increments, [U], partial molar volumes, V(o), and partial molar adiabatic compressibilities, K(S)(o), of the Li(+), Na(+), K(+), Rb(+), Cs(+), NH(4)(+), and N(CH(3))(4)(+) salts of poly(dAdT)poly(dAdT), poly(dGdC)poly(dGdC), poly(dIdC)poly(dIdC), poly(rA)poly(rU), poly(rG)poly(rC), poly(rI)poly(rC), and poly(rU) at 25 degrees C. When analyzing these data, we take into account the Donnan membrane equilibrium effect. Comparison between the values of [U], V(o), and K(S)(o) exhibited by the nucleic acid salts and respective chlorides (LiCl, NaCl, KCl, RbCl, CsCl, NH(4)Cl, and N(CH(3))(4)Cl) yields information about the state of counterion hydration in the vicinity of each nucleic acid structure studied here. Our analysis reveals that the poly(dGdC)poly(dGdC), poly(dIdC)poly(dIdC), and poly(rI)poly(rC) duplexes and single-stranded poly(rU) do not significantly influence the hydration properties of their condensed counterions. In the vicinity of these polymers, counterions retain their full hydration shells (within +/-15%). By contrast, counterions condensed around the poly(dAdT)poly(dAdT), poly(rA)poly(rU), and poly(rG)poly(rC) duplexes are significantly dehydrated and retain, respectively, only 65(+/-18)%, 34(+/-21)%, and 33(+/-9)% of their original hydration shells. Taken together, the volumetric data reported here provide important new information that ultimately may help us understand the central role that hydration and counterions play in modulating the conformational preferences of nucleic acids and the energetics of DNA recognition events.  相似文献   

14.
Formation of psi (+) and psi (-) DNA   总被引:4,自引:0,他引:4  
Y A Shin  G L Eichhorn 《Biopolymers》1984,23(2):325-335
DNA molecules can be organized into ordered aggregates of opposite handedness by complexation with polylysine and other polypeptides; we have investigated the conditions under which ψ(+) and ψ(?) structures are produced with double-helical synthetic polynucleotides. Both poly(dGdC)·poly(dGdC) and poly(dAdT)·poly(dAdT) readily form ψ(?) structures with polylysine, although the method of preparation can alter the CD spectra. The GC copolymer, which is more susceptible to conversion into A or Z conformers, forms ψ(+) structures with lysine–alanine copolypeptides more readily than the AT copolymer. Nucleotide base modifications that favor the Z structure, such as bromination and methylation, also favor ψ(+) formation, and the Co(NH3)6Cl3 reagent that readily induces the Z structure also leads to ψ(+). Thus, the production of the ψ(+) structure seems to be frequently correlated with susceptibility to A or Z formation, although there are some cases in which the B conformer also leads to ψ(+). Polyethylene glycol generally produces a ψ(?) structure; the differentiation between ψ(+) and ψ(?) structures seems to require electrically charged polymers.  相似文献   

15.
Left-handed Z-DNA is radically different from the most common right-handed B-DNA and can be stabilized by interactions with the Zα domain, which is found in a group of proteins, such as human ADAR1 and viral E3L proteins. It is well-known that most Zα domains bind to Z-DNA in a conformation-specific manner and induce rapid B–Z transition in physiological conditions. Although many structural and biochemical studies have identified the detailed interactions between the Zα domain and Z-DNA, little is known about the molecular basis of the B–Z transition process. In this study, we successfully converted the B–Z transition-defective Zα domain, vvZαE3L, into a B–Z converter by improving B-DNA binding ability, suggesting that B-DNA binding is involved in the B–Z transition. In addition, we engineered the canonical B-DNA binding protein GH5 into a Zα-like protein having both Z-DNA binding and B–Z transition activities by introducing Z-DNA interacting residues. Crystal structures of these mutants of vvZαE3L and GH5 complexed with Z-DNA confirmed the significance of conserved Z-DNA binding interactions. Altogether, our results provide molecular insight into how Zα domains obtain unusual conformational specificity and induce the B–Z transition.  相似文献   

16.
We examined three sonicated, specific-seqiemce polydeoxynucleotides in solution over a wide range of concentrations of several salts by 13P-nmr spectroscopy, and we found that the alternating copolymer poly(dAdT)·poly(dAdT) exhibits a dinucleotide repeat unit in all five salts and at all concentrations studied, as indicated by the presence of a doubled in its 31P-nmr spectra. The two components of the doublet show selective shift effects. The upfield component is assigned to dApdT in the gauche?-gauche? conformation and shifts upfield in all four monovalent salts used, relative to a single-stranded oligonucleotide control. The downfield component is assigned to dTpdA in the trans-gauche? conformation and shifts downfield with increasing CsF concentration but remains essentially constant in LiCl, NaCl, and CsCl. These changes indicate a fast noncooperative transition for poly(dAdT)·poly-(dAdT) from a presumed right-handed dinucleotide-repeat B-form to another conformation with a dinucleotide-repeat structure, via a continuum of structures that may differ in the extent of the winding of the double helix. Ethanol causes the upfield component to collapse into the other component, indicating conversion to a structure with a mononucleotide repeat unit and a trans-gauche? conformation. Up to 1M Mg2+ appears to have no significant effect on the phosphodiester conformations of poly(dAdT)·poly(dAdT). By contrast, poly-(dGdC)·poly(dGdC) gives a slow cooperative transition from what is considered to be a right-handed regular B-form to a left-handed Z-form on increasing MgCl2 and NaCl concentrations, although we observed no changes in chemical shifts below the transition points. The homopolymer poly(dA)·poly(dT) exhibits no unusual shift effects or transitions upon the addition of salts when compared to the oligonucleotide control and is considered to be a regular B-form with a gauche?-gauche? phosphodiester backbone conformation. These differences emphasize the distinct secondary structures of DNAs of different sequences and their selective responses to changes in solution conditions.  相似文献   

17.
The DNA binding efficacy and preferred mode of binding of a series of rhodamine-related chalcogenoxanthylium dyes was investigated by isothermal titration calorimetry (ITC) using ctDNA, [poly(dCdG)](2) and [poly(dAdT)](2), and by a topoisomerase I DNA unwinding (Topo I) assay. The dyes of this study showed tight binding to ctDNA with binding constants, K(b), on the order of 10(6)-10(7)M(-1). The ITC and Topo I assay studies suggested that the 9-substituent has a strong impact on binding modes ranging from an apparent preference for intercalation with a 9-2-thienyl substituent (similar binding to [poly(dCdG)](2) and [poly(dAdT)](2), re-supercoiling of DNA in the Topo I assay at <10(-5)M dye), to mixed binding modes with 9-phenyl derivatives (2- to 3-fold preference for binding to [poly(dAdT)](2), re-supercoiling of DNA in the Topo I assay at approximately 2 x 10(-5)M dye), to minor groove binding in a 9-(2-thienyl-5-diethylcarboxamide) derivative (strong preference for binding to [poly(dAdT)](2), did not show complete re-supercoiling in the Topo I assay). No binding to ctDNA was observed in one derivative with a 9-(3-thienyl-2-diethylcarboxamide) substituent, which cannot be co-planar with the xanthylium core. In series of dyes where the chalcogen atom was varied, the selenoxanthylium derivatives had 2- to 3-fold higher values of K(b) than the corresponding xanthylium, thioxanthylium, or telluroxanthylium derivatives, which all showed comparable values of K(b). The chalcogen atom appeared to have little influence on binding mode.  相似文献   

18.
Zinc ion-DNA polymer interactions   总被引:1,自引:0,他引:1  
X Jia  L G Marzilli 《Biopolymers》1991,31(1):23-44
The adjacent GN7-M-GN7 cross-linking and adjacent G-M-G sandwich-complex models for DNA metal ion binding were evaluated both with native DNAs differing in GC content as well as with the synthetic polymers poly [(dGdC)]2, poly[(dAdT)]2, and poly[(dAdC)(dGdT)]. The effect of Zn2+ was studied in depth, and limited studies were also performed with Co2+ and Mg2+. The results were compared to the extensive information available on Cu2+ binding to native DNAs and poly[(dAdT)]2. At high ratios of metal/base (R), Zn2+ caused all native DNAs to denature with the same melting temperature Tm, approximately 61 degrees C. A similar pattern was reported previously for Cu2+, but the typical Tm was approximately 35 degrees C. The extent of renaturation on cooling DNAs denatured in the presence of Zn2+ increased with GC content, as reported previously for Cu2+. These results, together with previously reported similarities, strongly indicate that the DNA binding characteristics of the two cations are similar. By comparison of the Tm values and hyperchromicity changes monitored at 260 and 282 nm, it is clear that, during thermal denaturation in the presence of Zn2+, both AT and GC regions were denatured, even at high R. The Tm vs R profile for the native DNAs was typical. The rise at low R and subsequent decrease at high R were inversely and directly related, respectively, to GC content. Except for poly[(dAdT)]2, where Tm increased with R, the other synthetic polymers exhibited the increase/decrease pattern. Poly[(dAdC)(dGdT)] gave a Tm value at high R of 54 degrees C. In the absence of Zn2+, this polymer exhibited little hypochromicity on cooling of denatured polymer. However, in the presence of Zn2+, nearly complete hypochromicity was observed, although the midpoint of the cooling curve was lower than the Tm value by approximately 15 degrees C at R = 10. These characteristics were similar to those with native DNAs, although viscosity and CD studies suggested that the "renatured" polymer was not identical to the unheated polymer. Furthermore, addition of Zn2+ after denaturation nearly completely reversed the absorption increase. This finding contrasts with those for native DNAs, where the Zn2+ must be present during denaturation in order to reverse the absorption increase nearly completely on cooling. With some caveats, poly[(dAdC)(dGdT)] appears to be a good model for native DNAs since its properties, including CD and uv changes on addition of Zn2+ to premelted and melted polymer, parallel those of the native polymers.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Abstract

The DNA binding of BMS 181176, an antitumor antibiotic derivative of rebeccamycin was characterized by DNA unwinding assays, as well as by fluorescence emission and polarization spectroscopic techniques. Unwinding and rewinding of supercoiled DNA was interpreted in terms of intercalation of BMS 181176 into DNA BMS 181176 shows an enhanced fluorescence emission upon binding to the AT sequence and no enhancement upon binding to the GC sequence. BMS 181176 appears to be a weaker binder to poly(dAdT).poly(dAdT) compared to doxorubicin and ethidium bromide. When bound to DNA, the rotational motion of BMS 181176 is substantially decreased as evident from the increase in fluorescence polarization. BMS 181176 exhibits a range of binding strengths depending on the DNA This is demonstrated by the Acridine Orange displacement assay using fluorescence polarization.  相似文献   

20.
Using high precision densimetric and ultrasonic measurements, we have determined, at 25°C, the apparent molar volumes ΦV and the apparent molar compressibilities ΦKS of four nucleic acid duplexes—namely, the DNA duplex, poly(dIdC)poly(dIdC); the RNA duplex, poly(rA)poly(rU); and the two DNA/RNA hybrid duplexes, poly(rA)poly(dT) and poly(dA)poly(rU). Using available fiber diffraction data on these duplexes, we have calculated the molecular volumes as well as the solvent‐accessible surface areas of the constituent charged, polar, and nonpolar atomic groups. We found that the hydration properties of these nucleic acid duplexes do not correlate with the extent and the chemical nature of the solvent‐exposed surfaces, thereby suggesting a more specific set of duplex–water interactions beyond general solvation effects. A comparative analysis of our volumetric data on the four duplexes, in conjunction with available structural information, suggests the following features of duplex hydration: (a) The four duplexes exhibit different degrees of hydration, in the order poly(dIdC)poly(dIdC) > poly(dGdC)poly(dGdC) > poly(dAdT)poly(dAdT) ≈ poly(dA)poly(dT). (b) Repetitive AT and IC sequences within a duplex are solvated beyond general effects by a spine of hydration in the minor groove, with this sequence‐specific water network involving about 8 additional water molecules from the second and, perhaps, even the third hydration layers. (c) Repetitive GC and IC sequences within a duplex are solvated beyond general effects by a “patch of hydration” in the major groove, with this water network involving about 13 additional water molecules from the second and, perhaps, even the third hydration layers. (d) Random sequence, polymeric DNA duplexes, which statistically lack extended regions of repetitive AT, GC, or IC sequences, do not experience such specific enhancements of hydration. Consequently, consistent with our previous observations (T. V. Chalikian, A. P. Sarvazyan, G. E. Plum, and K. J. Breslauer, Biochemistry, 1994, Vol. 33, pp. 2394–2401), duplexes with approximately 50% AT content exhibit the weakest hydration, while an increase or decrease from this AT content causes enhancement of hydration, either due to stronger hydration of the minor groove (an increase in AT content) or due to stronger hydration of the major groove (an increase in GC content). (e) In dilute aqueous solutions, a B‐DNA duplex is more hydrated than an A‐DNA duplex, a volumetric‐based conclusion that is in agreement with previous results obtained on crystals, fibers, and DNA solutions in organic solvent–water mixtures. (f) the A‐like, RNA duplex poly(rA)poly(rU) and the structurally similar A‐like, hybrid duplex poly(rA)poly(dT), exhibit similar hydration properties, while the structurally distinct A‐like, hybrid duplex poly(rA)poly(dT) and non‐A‐like, hybrid duplex poly(dA)poly(rU) exhibit differential hydration properties, consistent with structural features dictating hydration characteristics. We discuss how volumetric characterizations, in conjunction with structural studies, can be used to describe, define, and resolve the general and sequence/conformation‐specific hydration properties of nucleic acid duplexes. © 1999 John Wiley & Sons, Inc. Biopoly 50: 459–471, 1999  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