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.
Influence of surface shape on DNA binding of bimetallo helicates   总被引:1,自引:0,他引:1  
In order to probe the DNA-helicate interactions responsible for the DNA binding and remarkable changes of the DNA secondary structure induced by a tetracationic bi-metallo helicate [Fe(2)(L(1))(3)](4+) (L(1)=C(25)H(20)N(4)), we have designed and synthesised derivatives with hydrophobic methyl groups at different positions on the ligand backbone. Two dimetallo helicates [Fe(2)(L(i))(3)](4+) were prepared using ligands L(3) and L(5) with the methyl substituent on, respectively, the 3 and 5 positions of the pyridyl ring thus producing a wider or slightly longer tetracationic DNA binder. UV/visible absorbance, circular and linear dichroism spectroscopies have been used to characterize the interactions of the cylinders with DNA with the aim of investigating any sequence preference or selectivity upon binding. Competitive binding studies using fluorescent dyes Hoechst 33258 (a minor groove binder), ethidium bromide (an intercalator) and a major groove binding cation (cobalt (III) hexammine) which induces the B-->Z transition have been employed to determine the binding geometries of the enantiomers of two methylated helicates (L(3) and L(5)) to DNA and compare with the data obtained previously for the unmethylated analogue (L(1)). The results demonstrate that the racemic mixtures and the resolved enantiomers of all helicates bind to DNA inducing structural changes. The overall conclusion from the effect of adding these groups to the surface of the parent helicate is that increasing the width (L(3)) reduces the DNA binding strength, the bending and coiling effect and the groove selectivity of the enantiomers compared with the parent compound. There is limited evidence to suggest a slight GC sequence preference. Lengthening the helicate (L(5)) results in DNA interactions similar to those of the parent compounds, with an increased preference of the P enantiomer for the minor groove indicating an enhancement of mode selectivity.  相似文献   

2.
Proudfoot EM  Mackay JP  Karuso P 《Biochemistry》2001,40(15):4867-4878
The molecular recognition of oligonucleotides by chiral ruthenium complexes has been probed by NMR spectroscopy using the template Delta-cis-alpha- and Delta-cis-beta-[Ru(RR-picchxnMe(2)) (bidentate)](2+), where the bidentate ligand is one of phen (1,10-phenanthroline), dpq (dipyrido[3,2-f:2',3'-h]quinoxaline), or phi (9,10-phenanthrenequinone diimine) and picchxnMe(2)() is N,N'-dimethyl-N,N'-di(2-picolyl)-1,2-diaminocyclohexane. By varying only the bidentate ligand in a series of complexes, it was shown that the bidentate alone can alter binding modes. DNA binding studies of the Delta-cis-alpha-[Ru(RR-picchxnMe(2))(phen)](2+) complex indicate fast exchange kinetics on the chemical shift time scale and a "partial intercalation" mode of binding. This complex binds to [d(CGCGATCGCG)](2) and [d(ATATCGATAT)](2) at AT, TA, and GA sites from the minor groove, as well as to the ends of the oligonucleotide at low temperature. Studies of the Delta-cis-beta-[Ru(RR-picchxnMe(2))(phen)](2+) complex with [d(CGCGATCGCG)](2) showed that the complex binds only weakly to the ends of the oligonucleotide. The interaction of Delta-cis-alpha-[Ru(RR-picchxnMe(2))(dpq)](2+) with [d(CGCGATCGCG)](2) showed intermediate exchange kinetics and evidence of minor groove intercalation at the GA base step. In contrast to the phen and dpq complexes, Delta-cis-alpha- and Delta-cis-beta-[Ru(RR-picchxnMe(2))(phi)](2+) showed evidence of major groove binding independent of the metal ion configuration. DNA stabilization induced by complex binding to [d(CGCGATCGCG)](2) (measured as DeltaT(m)) increases in the order phen < dpq and DNA affinity in the order phen < dpq < phi. The groove binding preferences exhibited by the different bidentate ligands is explained with the aid of molecular modeling experiments.  相似文献   

3.
Baruah H  Wright MW  Bierbach U 《Biochemistry》2005,44(16):6059-6070
[PtCl(en)(ACRAMTU-S)](NO(3))(2) (PT-ACRAMTU; en = ethane-1,2-diamine, ACRAMTU = 1-[2-(acridin-9-ylamino)ethyl]-1,3-dimethylthiourea) is a dual metalating/intercalating DNA binding drug conjugate that shows cytotoxicity at micromolar to nanomolar concentrations in a wide range of solid tumor cell lines. In approximately 80% of its adducts, PT-ACRAMTU binds to guanine-N7 in the major groove, selectively at 5'-CG sites [Budiman, M. E. et al. (2004) Biochemistry 43, 8560-8567]. Here, we report the synthesis, physical characterization, and NMR solution structure of a site-specifically modified octamer containing this adduct, 5'-CCTCGTCC-3'/3'-GGAGCAGG-5', where the asterisk indicates the [Pt(en)ACRAMTU)](3+) fragment. The structure was determined by a combination of high-resolution 2-D NMR spectroscopy and restrained molecular dynamics/molecular mechanics (rMD/MM) calculations using 179 NOE distance restraints and refined to an r(6) weighted residual (R(x)) of 9.2 x 10(-)(2) using the complete relaxation matrix approach. An average structure was calculated from the final ensemble of 19 rMD geometries showing pairwise root-mean-square deviations of <1.05 A. The dual binding increases the thermal stability of the octamer compared to the unmodified duplex (DeltaT(m) = 13.2 degrees ). The modified sequence shows structural features reminiscent of both B- and A-type DNA. Watson-Crick hydrogen bonding is intact at and beyond the adduct site. Platinum is bound to the N7 position of G5 in the major groove, and ACRAMTU intercalates into the central 5'-C4G5/C12G13 base-pair step on the 5'-face of the platinated nucleobase. The chromophore's long axis is aligned with the long axes of the adjacent base pairs, maximizing intermolecular pi-pi stacking interactions. PT-ACRAMTU lengthens (rise, 6.62 A) and unwinds (twist, 15.4 degrees ) the duplex at the central base-pair step but does not cause helical bending. No C3'-endo deoxyribose pucker and no significant roll are observed at the site of intercalation/platination, which clearly distinguishes the PT-ACRAMTU-induced damage from the 1,2-intrastrand cross-link formed by cisplatin. Overall, the DNA perturbations produced by PT-ACRAMTU do not appear to mimic those caused by the major cisplatin lesion. Instead, intriguing structural similarities are observed for PT-ACRAMTU's monoadduct and the N7 adducts of dual major-groove alkylating/intercalating antitumor agents, such as the pluramycins.  相似文献   

4.
The interaction of the enantiomeric complexes Lambda- and Delta-[Ru(bpy)(2)(pbmz)](PF(6))(2) (bpy=2,2'-bipyridine, pbmz=2-(2'-pyridyl)benzimidazole) with the DNA duplex d(CGCGAATTCGCG)(2) was investigated by means of 2D NMR techniques. The synthesis of the enantiomers was based on the optically pure complexes Lambda- and Delta-[Ru(bpy)(2)(py)(2)](2+) and were characterized by CD and NMR spectroscopy. NMR data indicate that both enantiomers bind weakly to the oligonucleotide, approaching from the minor groove at the centre of the helix. The perturbation of the B-DNA conformation is minor with an apparent absence of enantioselectivity. Molecular modelling calculations in conjunction with the NOE data support the suggestion that more than one binding modes are present. The imidazole amine group of the pbmz ligand is probably hydrogen bonded to the DNA phosphodiesteric backbone at the AATT step, and this may provide an explanation for the diminished enantioselectivity observed.  相似文献   

5.
Spectroscopic and electrochemical study of the [Fe(4)(mu(3)-S)(3)(NO)(7)](-) photochemical reaction and thermodynamic calculations of relevant systems demonstrate the redox character of this process. The photoinduced electron transfer between substrate clusters in excited and ground state (probably via exciplex formation) results in dismutation yielding unstable [Fe(4)(mu(3)-S)(3)(NO)(7)](2-) and [Fe(4)(mu(3)-S)(3)(NO)(7)](0). Back electron transfer between the primary products is responsible for fast reversibility of the photochemical reaction in deoxygenated solutions. In the presence of an electron acceptor (such as O(2), MV(2+) or NO) an oxidative quenching of the (*)[Fe(4)(mu(3)-S)(3)(NO)(7)](-) is anticipated, although NO seems to participate as well in the reductive quenching. The electron acceptors can also regenerate the substrate from its reduced form ([Fe(4)(mu(3)-S)(3)(NO)(7)](2-)), whereas the other primary product ([Fe(4)(mu(3)-S)(3)(NO)(7)](0)) decomposes to the final products. The suggested mechanism fits well to all experimental observations and shows the thermodynamically favored pathways and explains formation of all major (Fe(2+), S(2-), NO) and minor products (N(2)O, Fe(3+)). The photodissociation of nitrosyl ligands suggested earlier as the primary photochemical step cannot be, however, definitely excluded and may constitute a parallel pathway of [Fe(4)(mu(3)-S)(3)(NO)(7)](-) photolysis.  相似文献   

6.
Solution structures of DNA duplexes containing oxanine (Oxa, O) opposite a cytosine (O:C duplex) and opposite a thymine (O:T duplex) have been solved by the combined use of (1)H NMR and restrained molecular dynamics calculation. One mismatch pair was introduced into the center of the 11-mer duplex of [d(GTGACO(6)CACTG)/d(CAGTGX(17)GTCAC), X = C or T]. (1)H NMR chemical shifts and nuclear Overhauser enhancement (NOE) intensities indicate that both the duplexes adopt an overall right-handed B-type conformation. Exchangeable resonances of C(17) 4-amino proton of the O:C duplex and of T(17) imino proton of O:T duplex showed unusual chemical shifts, and disappeared with temperature increasing up to 30 °C, although the melting temperatures were >50 °C. The O:C mismatch takes a wobble geometry with positive shear parameter where the Oxa ring shifted toward the major groove and the paired C(17) toward the minor groove, while, in the O:T mismatch pair with the negative shear, the Oxa ring slightly shifted toward the minor groove and the paired T(17) toward the major groove. The Oxa mismatch pairs can be wobbled largely because of no hydrogen bond to the O1 position of the Oxa base, and may occupy positions in the strands that optimize the stacking with adjacent bases.  相似文献   

7.
The nitrogenase-catalyzed H(2) evolution and acetylene-reduction reactions using Ti(III) and dithionite (DT) as reductants were examined and compared under a variety of conditions. Ti(III) is known to make the all-ferrous Fe protein ([Fe(4)S(4)](0)) and lowers the amount of ATP hydrolyzed during nitrogenase catalysis by approximately 2-fold. Here we further investigate this behavior and present results consistent with the Fe protein in the [Fe(4)S(4)](0) redox state transferring two electrons ([Fe(4)S(4)](2+)/[Fe(4)S(4)](0)) per MoFe protein interaction using Ti(III) but transferring only one electron ([Fe(4)S(4)](2+)/[Fe(4)S(4)](1+)) using DT. MoFe protein specific activity was measured as a function of Fe:MoFe protein ratio for both a one- and a two-electron transfer reaction, and nearly identical curves were obtained. However, Fe protein specific activity curves as a function of MoFe:Fe protein ratio showed two distinct reactivity patterns. With DT as reductant, typical MoFe inhibition curves were obtained for operation of the [Fe(4)S(4)](2+)/[Fe(4)S(4)](1+) redox couple, but with Ti(III) as reductant the [Fe(4)S(4)](2+)/[Fe(4)S(4)](0) redox couple was functional and MoFe inhibition was not observed at high MoFe:Fe protein ratios. With Ti(III) as reductant, nitrogenase catalysis produced hyperbolic curves, yielding a V(max) for the Fe protein specific activity of about 3200 nmol of H(2) min(-1) mg(-1) Fe protein, significantly higher than for reactions conducted with DT as reductant. Lag phase experiments (Hageman, R. V., and Burris, R. H. (1978) Proc. Natl. Acad. Sci. U. S. A. 75, 2699-2702) were carried out at MoFe:Fe protein ratios of 100 and 300 using both DT and Ti(III). A lag phase was observed for DT but, with Ti(III) product formation, began immediately and remained linear for over 30 min. Activity measurements using Av-Cp heterologous crosses were examined using both DT and Ti(III) as reductants to compare the reactivity of the [Fe(4)S(4)](2+)/[Fe(4)S(4)](1+) and [Fe(4)S(4)](2+)/[Fe(4)S(4)](0) redox couples and both were inactive. The results are discussed in terms of the Fe protein transferring two electrons per MoFe protein encounter using the [Fe(4)S(4)](2+)/[Fe(4)S(4)](0) redox couple with Ti(III) as reductant.  相似文献   

8.
The structure of the 1:1 nogalamycin:d(ATGCAT)2 complex has been determined in solution from high-resolution NMR data and restrained molecular dynamics (rMD) simulations using an explicit solvation model. The antibiotic intercalates at the 5'-TpG step with the nogalose lying along the minor groove towards the centre of the duplex. Many drug-DNA nuclear Overhauser enhancements (NOEs) in the minor groove are indicative of hydrophobic interactions over the TGCA sequence. Steric occlusion prevents a second nogalamycin molecule from binding at the symmetry-related 5'-CpA site, leading to the conclusion that the observed binding orientation in this complex is the preferred orientation free of the complication of end-effects (drug molecules occupy terminal intercalation sites in all X-ray structures) or steric interactions between drug molecules (other NMR structures have two drug molecules bound in close proximity), as previously suggested. Fluctuations in key structural parameters such as rise, helical twist, slide, shift, buckle and sugar pucker have been examined from an analysis of the final 500 ps of a 1 ns rMD simulation, and reveal that many sequence-dependent structural features previously identified by comparison of different X-ray structures lie within the range of dynamic fluctuations observed in the MD simulations. Water density calculations on MD simulation data reveal a time-averaged pattern of hydration in both the major and minor groove, in good agreement with the extensive hydration observed in two related X-ray structures in which nogalamycin is bound at terminal 5'-TpG sites. However, the pattern of hydration determined from the sign and magnitude of NOE and ROE cross-peaks to water identified in 2D NOESY and ROESY experiments identifies only a few "bound" water molecules with long residence times. These solvate the charged bicycloaminoglucose sugar ring, suggesting an important role for water molecules in mediating drug-DNA electrostatic interactions within the major groove. The high density of water molecules found in the minor groove in X-ray structures and MD simulations is found to be associated with only weakly bound solvent in solution.  相似文献   

9.
To achieve cellular iron deprivation by chelation, it is important to develop chelators with selective metal-binding properties. Selectivity for iron has long been the province of certain oxygen-donor chelators such as desferrioxamine, which target Fe(III) and exploit the strength of a relatively ionic Fe(III)-O interaction. We have been studying novel chelators that possess mechanisms to selectively chelate +2 biometals, particularly tachpyr [N,N',N"-tris(2-pyridylmethyl)-1,3,5-cis,cis-triaminocyclohexane] and derivatives from N,N',N"-trialkylation and pyridine ring alkylation. Metal-exchange and metal-binding competition reactions have been conducted at pH 7.4, 37 degrees C and time periods until no further change was observed (generally 24-48 h). Under anaerobic conditions, tachpyr is strongly selective for iron, binding 95+/-5% Fe(II) versus 5+/-5% Zn(II) in the forms [Fe(tachpyr)](2+) and [Zn(tachpyr)](2+) respectively. Under aerobic conditions, tachpyr complexes Fe(II) more effectively than Fe(III), forming iminopyridyl complexes [Fe(tachpyr-ox-n)](2+) (n=2, 4) by O(2)-induced and iron-mediated oxidative dehydrogenation. Complexes [Fe(tachpyr-ox-n)](2+) are also strongly bound forms of iron that are unaffected by an excess of Zn(II) (75 mol zinc:1 mol iron complex). The preference of tachpyr for iron over zinc under aerobic conditions appears to be hindered by oxidation of Fe(II) to Fe(III), such that the proportions bound are 44+/-10% Fe(II) versus 56+/-10% Zn(II), in the respective forms [Fe(tachpyr-ox-n)](2+) and [Zn(tachpyr)](2+). However, upon addition of the reducing agent Na(2)S(2)O(4) that converts Fe(III) to Fe(II), the binding proportions shift to 76+/-10% Fe(II) versus 24+/-10% Zn(II), demonstrating a clear preference of tachpyr for Fe(II) over Zn(II). Iron(II) is in the low-spin state in [Fe(tachpyr)](2+) and [Fe(tachpyr-ox-n)](2+) (n=2, 4), which is a likely cause of the observed selectivity. N-methylation of tachpyr [giving (N-methyl)(3)tachpyr] results in the loss of selectivity for Fe(II), which is attributed to the steric effect of the methyl groups and a resulting high-spin state of Fe(II) in [Fe(N-methyl)(3)tachpyr)](2+). The relationship of chelator selectivity to cytotoxicity in the tach family will be discussed.  相似文献   

10.
The (dG)n.(dC)n-containing 34mer DNA duplex [d(A2G15C15T2)]2 can be effectively converted from the B-DNA to the A-DNA conformation by neomycin, spermine and Co(NH3)6(3+). Conversion is demonstrated by a characteristic red shift in the circular dichroism spectra and dramatic NMR spectral changes in chemical shifts. Additional support comes from the substantially stronger CH6/GH8-H3'NOE intensities of the ligand-DNA complexes than those from the native DNA duplex. Such changes are consistent with a deoxyribose pucker transition from the predominate C2'-endo (S-type) to the C3'-endo (N-type). The changes for all three ligand-DNA complexes are identical, suggesting that those three complex cations share common structural motifs for the B- to A-DNA conversion. The A-DNA structure of the 4:1 complex of Co(NH3)6(3+)/d(ACCCGCGGGT) has been analyzed by NOE-restrained refinement. The structural basis of the transition may be related to the closeness of the two negatively charged sugar-phosphate backbones along the major groove in A-DNA, which can be effectively neutralized by the multivalent positively charged amine functions of these ligands. In addition, ligands like spermine or Co(NH3)6(3+) can adhere to guanine bases in the deep major groove of the double helix, as is evident from the significant direct NOE cross-peaks from the protons of Co(NH3)6(3+) to GH8, GH1 (imino) and CH4 (amino) protons. Our results point to future directions in preparing more potent derivatives of Co(NH3)6(3+) for RNA binding or the induction of A-DNA.  相似文献   

11.
The synthesis and characterization of the diiron(II) complex [Fe(2)(μ-OTf)(2)-(PIC(2)DET)(2)](BARF)(2) (2), where PIC(2)DET is a 2,3-diethynyltriptycene-linked dipicolinic methyl ester ligand, are described. The dication in 2, contains, [Fe(2)(μ-OTf)(2)(PIC(2)DET)(2)](2+) two symmetry-equivalent iron atoms with octahedral coordination geometries. Each metal ion has a N(2)O(4) atom donor set that includes four atoms from two picolinic ester N,O chelate rings, as well as two oxygen atoms from the bridging trifluoromethanesulfonate groups. The Fe(2)(μ-OTf)(2) core of 2 is stabilized by two PIC(2)DET ligands that bind the two metal ions in a head-to-head fashion, leading to an Fe···Fe distance of 5.173(1)?. Molar conductivity data for 2 are consistent with Fe(2)(μ-OTf)(2)(PIC(2)DET)(2)](2+) retaining its identity in acetone solutions, where it behaves as a 2:1 electrolyte. (1)H NMR spectroscopic, solution (d(6)-acetone) and solid-state magnetic susceptibility data all indicate that the iron atoms of 2 are high-spin (S = 2). A fit of the magnetic data (2 - 300K) to a spin-only isotropic exchange Hamiltonian H = -2JS(1)·S(2) are consistent with weak antiferromagnetic coupling between the two iron atoms with J ~ -0.99(2) cm(-1) and g = 2.10(1).  相似文献   

12.
L J Ming  L Que  A Kriauciunas  C A Frolik  V J Chen 《Biochemistry》1991,30(50):11653-11659
The active site structure of isopenicillin N synthase (IPNS) has been previously studied by the use of M?ssbauer, EPR, electronic absorption, and NMR spectroscopies [Chen, V.J., Frolik, C.A., Orville, A.M., Harpel, M.R., Lipscomb, J.D., Surerus, K.K., & Münck, E. (1989) J. Biol. Chem. 264, 21677-21681; Ming, L.-J., Que, L., Jr., Kriauciunas, A., Frolik, C.A., & Chen, V.J. (1990) Inorg. Chem. 26, 1111-1112]. These studies have revealed three coordinated His residues along with three sites for substrate [delta-(L-alpha-aminoadipoyl)-L-cysteinyl-D-valine, ACV], NO, and water binding on the active Fe(II) of IPNS. We report here NMR studies of Fe(II)IPNS and its Co(II)-substituted derivative [Co(II)IPNS]. By the use of NOE techniques on the Co(II)IPNS-ACV complex, we have recognized a -CH2-CH less than spin system at 14.6, 24.3, and 38.6 ppm that is assigned to the alpha and beta protons of a coordinated Asp residue. Corresponding solvent nonexchangeable features are found near 40 ppm in Fe(II)IPNS and the Fe(II)IPNS-ACV complex, but the peaks are too broad for NOE effects to be observed. The binding of NO to the Fe(II) center results in a significant change in the configuration of the metal site: (a) The C beta H2 resonances due to the coordinated Asp residue disappear. The loss of the signal may indicate a change of the carboxylate configuration from syn-like to anti-like or, less likely, its displacement by NO.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Molecular modeling and energy minimisation calculations have been used to investigate the interaction of chromium(III) complexes in different ligand environments with various sequences of B-DNA. The complexes are [Cr(salen)(H(2)O)(2)](+); salen denotes 1, 2 bis-salicylideneaminoethane, [Cr(salprn)(H(2)O)(2)](+); salprn denotes 1, 3 bis- salicylideneaminopropane, [Cr(phen)(3)](3+); phen denotes 1, 10 phenanthroline and [Cr(en)(3)](3+); en denotes ethylenediamine. All the chromium(III) complexes are interacted with the minor groove and major groove of d(AT)(12), d(CGCGAATTCGCG)(2) and d(GC)(12) sequences of DNA. The binding energy and hydrogen bond parameters of DNA-Cr complex adduct in both the groove have been determined using molecular mechanics approach. The binding energy and formation of hydrogen bonds between chromium(III) complex and DNA has shown that all complexes of chromium(III) prefer minor groove interaction as the favourable binding mode.  相似文献   

14.
The reaction of N-acetylmethionine (N-AcMet) with the complex [Pt(Et(2)en)(D(2)O)(2)](2+) (Et(2)en=N,N-diethylethylenediamine) was studied by NMR spectroscopy and molecular mechanics calculations. Complexes containing two methionine residues coordinated to the platinum atom were calculated to be relatively high in energy unless the bulk of the methionine residues was directed away from the diethyl group of the Et(2)en ligand. In contrast, sulfur-oxygen chelates were found to be relatively free of steric clashes. Experimentally, two sets of NMR resonances were observed when [Pt(Et(2)en)(D(2)O)(2)](2+) was reacted with N-AcMet; variable temperature experiments indicated intermediate chemical exchange between the two sets of resonances. NMR studies indicated that the resonances corresponded to [Pt(Et(2)en)(N-AcMet-S,O)](+) complexes with the sulfur atom trans to the diethyl group of the Et(2)en ligand. No product with the sulfur atom cis to the diethyl group was observed experimentally even though molecular mechanics calculations suggested that such forms have few steric clashes. The NMR results suggested that the chemical exchange was a result of sulfur chirality inversion. In early stages of the reaction, a [Pt(Et(2)en)(N-AcMet-S)(D(2)O)](+) complex was observed, indicating that coordination of the oxygen to form the chelate is relatively slow.  相似文献   

15.
Snoussi K  Leroy JL 《Biochemistry》2002,41(41):12467-12474
We have investigated by NMR the effects of NH(4)(+) on the chemical shifts, on the structure, and on the imino proton exchange kinetics of two duplexes containing an A-tract, [d(CGCGAATTCGCG)](2) and [d(GCA(4)T(4)GC)](2), and of a B-DNA duplex,[d(CGCGATCGCG)](2). Upon NH(4)(+) addition to [d(CGCGAATTCGCG)](2), the adenosine H2 protons, the thymidine imino protons, and the guanosine imino proton of the adjacent G.C pair show unambiguous chemical shifts. Similar shifts are observed in the A-tract of [d(GCA(4)T(4)GC)](2) and for the A5(H2) proton of the B DNA duplex [d(CGCGATCGCG)](2). The localization of the shifted protons suggests an effect related to NH(4)(+) binding in the minor groove. The cross-peak intensities of the NOESY spectra collected at low and high NH(4)(+) concentrations are comparable, and the COSY spectra do not show any change of the sugar pucker. This indicates a modest effect of ammonium binding on the duplex structures. Nevertheless, the imino proton exchange catalysis by ammonia provides evidence for a substantial effect of NH(4)(+) binding on the A.T base-pair kinetics in the A-tracts. Proton exchange experiments performed at high and low NH(4)(+) concentrations show the occurrence of two native conformations in proportions depending on the NH(4)(+) concentration. The base-pair lifetimes and the open-state lifetimes of each conformation are distinct. Exchange from each conformation proceeds via a single open state. But if, and only if, the NH(4)(+) concentration is kept larger than 1 M, the A.T imino proton exchange times of A-tract sequences exhibit a linear dependence versus the inverse of the NH(3) proton acceptor concentration. This had been interpreted as an indication for two distinct base-pair opening modes (W?rml?nder, S., Sen, A., and Leijon, M. (2000) Biochemistry 39, 607-615).  相似文献   

16.
NMR spectroscopy has been used to observe the effects of the amine ligand on the rate of reaction of platinum diamine and triamine complexes with DNA and protein residues. Whereas [Pt(dien)Cl]Cl and [Pt(dien)(D(2)O)](2+) have been known to react faster with thioether residues such as N-AcMet than with 5'-GMP, we found that [Pt(Me(4)en)(D(2)O)(2)](2+) appeared to react faster with 5'-GMP. To quantitatively assess the factors influencing the rates of reaction, rate constants at pH 4 were determined for the reactions of [Pt(en)(D(2)O)(2)](2+) [en = ethylenediamine] and [Pt(Me(4)en)(D(2)O)(2)](2+) with N-AcMet, N-AcHis, 5'-GMP, and Guo (guanosine). In each case the less bulky complex ([Pt(en)(D(2)O)(2)](2+)) reacts more quickly than does the bulkier [Pt(Me(4)en)(D(2)O)(2)](2+), as expected. Both complexes reacted faster with 5'-GMP; however, analysis of the rate constants suggests that the [Pt(en)(D(2)O)(2)](2+) complex favors reaction with 5'-GMP due to hydrogen bonding with the 5'-phosphate, whereas [Pt(Me(4)en)(D(2)O)(2)](2+) disfavors reaction with N-AcMet due to steric clashes. Bulk had relatively little effect on the rate constant with N-AcHis, suggesting that peptides or proteins that coordinate via His residues would not have their reactivity affected by bulky diamine ligands.  相似文献   

17.
Despite their structural similarity, [Pt(dien)(1-MeC-N3)](2+) (1), [Pd(dien)(1-MeC-N3)](2+) (2), and [Pt(NH(3))(3)(1-MeC-N3)](2+) (3) (with dien=diethylenetriamine and 1-MeC=neutral 1-methylcytosine) behave in part markedly different at strongly alkaline pH (12-13) and at room temperature. While 1 and 2, yet not 3 show linkage isomerization from N3 to N4, deamination of the cytosine nucleobase to 1-methyluracilate occurs with 1 and 3, yet not with 2. Pathways leading to N3,N4-diplatinated 1-MeC(-) complexes (1-MeC(-)=1-methylcytosine, deprotonated at exocyclic amino group N4) have been studied at high pH by starting from 1 and 3, respectively, and adding (dien)Pt(II). It appears that initial migration of the metal entity from N3 to N4, followed by binding of the second metal to the available N3 site, is favored over sequential coordination to N3 and then N4. X-ray crystal data of 1-3 density functional theory (DFT) calculations, and NMR ((1)H, (195)Pt) data are presented.  相似文献   

18.
Desulfovibrio gigas desulforedoxin (Dx) consists of two identical peptides, each containing one [Fe-4S] center per monomer. Variants with different iron and zinc metal compositions arise when desulforedoxin is produced recombinantly from Escherichia coli. The three forms of the protein, the two homodimers [Fe(III)/Fe(III)]Dx and [Zn(II)/Zn(II)]Dx, and the heterodimer [Fe(III)/Zn(II)]Dx, can be separated by ion exchange chromatography on the basis of their charge differences. Once separated, the desulforedoxins containing iron can be reduced with added dithionite. For NMR studies, different protein samples were prepared labeled with (15)N or (15)N + (13)C. Spectral assignments were determined for [Fe(II)/Fe(II)]Dx and [Fe(II)/Zn(II)]Dx from 3D (15)N TOCSY-HSQC and NOESY-HSQC data, and compared with those reported previously for [Zn(II)/Zn(II)]Dx. Assignments for the (13)C(alpha) shifts were obtained from an HNCA experiment. Comparison of (1)H-(15)N HSQC spectra of [Zn(II)/Zn(II)]Dx, [Fe(II)/Fe(II)]Dx and [Fe(II)/Zn(II)]Dx revealed that the pseudocontact shifts in [Fe(II)/Zn(II)]Dx can be decomposed into inter- and intramonomer components, which, when summed, accurately predict the observed pseudocontact shifts observed for [Fe(II)/Fe(II)]Dx. The degree of linearity observed in the pseudocontact shifts for residues >/=8.5 A from the metal center indicates that the replacement of Fe(II) by Zn(II) produces little or no change in the structure of Dx. The results suggest a general strategy for the analysis of NMR spectra of homo-oligomeric proteins in which a paramagnetic center introduced into a single subunit is used to break the magnetic symmetry and make it possible to obtain distance constraints (both pseudocontact and NOE) between subunits.  相似文献   

19.
Berberine, an isoquinoline plant alkaloid, belongs to the structural class of protoberberines. Recently, the ability of these compounds to act as Topoisomerase I or II poisons, was related to the antitumor activity. The binding of protoberberins to DNA has been studied and the partial intercalation into the double helix has been considered responsible for their activity. We have studied the interaction of berberine with the double helix oligonucleotides d(AAGAATTCTT)(2), d(GCGATCGC)(2), d(CGTATACG)(2), d(CGTACG)(2), 5'-d(ACCTTTTTGATGT)-3'/5(ACATCAAAAAGGT)-3' and with the single strand 5'-d(ACATCAAAAAGGT)-3', by 1H, 31P NMR and UV spectroscopy. Phosphorus resonance experiments were performed to detect small conformational changes of the phosphoribose backbone, in the case that an intercalation process occurs. Our data reveal that berberine does not intercalate into the duplexes studied, and binds preferentially to AT rich sequences. The structure of the complex with d(AAGAATTCTT)(2) was determined by using proton 2D NOESY spectra, which allowed to obtain several NOE contacts between the drug and the nucleotide. Structural models were built up by Molecular Mechanics (MM) and Molecular Dynamics (MD) calculations, by using the inter-proton distances derived from the NOE values. Berberine results to be located in the minor groove, lying with the convex side on the helix groove and presenting the positively charged nitrogen atom close to the negative ionic surface of the oligomer. The large 1H chemical shifts variation, observed for the drug when it is added to the above duplexes, as well as to the single strand oligomer, was interpreted with non-specific ionic interactions. The binding constants were measured by UV and NMR spectroscopy. They are strongly affected by the ionic strength and by the self-association process, which commonly occurs with this type of drugs. A dimerisation constant was measured and the value was included in the calculations of the binding constants. The results obtained show that the non-specific ionic interactions represent the major contribution to the values of the binding constants. These parameters, as well as the protons chemical shift variation of the ligand, are thus not diagnostic for the identification of a drug/DNA complex.  相似文献   

20.
Mono- and di-phosphine diiron azadithiolate complexes [{(mu-SCH(2))(2)N(4-NO(2)C(6)H(4))}Fe(2)(CO)(5)(PMe(3))] (2), [{(mu-SCH(2))(2)N(4-NO(2)C(6)H(4))}{Fe(CO)(2)L}(2)] (3, L=PMe(3); 4, PMe(2)Ph) and the mu-hydride diiron complex [3(FeHFe)](+)[PF(6)](-) were prepared as biomimetic models of the active site of Fe-only hydrogenases. The complexes 2-4 and [3(FeHFe)](+)[PF(6)](-) were characterized by IR, (31)P, (1)H and (13)C NMR spectra and their molecular structures were determined by single crystal X-ray analyses. The PMe(3) ligand in complex 2 lies on the basal position. The PMe(3)-disubstituted complex 3 exists as two configuration isomers, transoid basal/basal and apical/basal, in the crystalline state, while two PMe(2)Ph ligands of 4 are in an apical/basal orientation. The variable temperature (31)P NMR spectra of 2 and 3 were made to have an insight into the existence of the possible conformation isomers of 2 and 3 in solution. The [3(FeHFe)](+) cation possesses the sole transoid ba/ba geometry as other reported mu-hydride diiron analogues. The electrocatalytic property of {(mu-SCH(2))(2)NC(6)H(5)}[Fe(CO)(2)PMe(3)](2) (5) was studied for proton reduction in the presence of HOAc.  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