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.
Platinum anticancer drug DNA intrastrand cross-link models, LPt(d(G*pG*)) (G* = N7-platinated G residue, L = R(4)dt = bis-3,3'-(5,6-dialkyl)-1,2,4-triazine), and R = Me or Et), undergo slow Pt-N7 bond rotation. NMR evidence indicated four conformers (HH1, HH2, ΔHT1, and ΛHT2); these have different combinations of guanine base orientation (head-to-head, HH, or head-to-tail, HT) and sugar-phosphodiester backbone propagation relative to the 5'-G* (the same, 1, or opposite, 2, to the direction in B DNA). In previous work on LPt(d(G*pG*)) adducts, Pt-N7 rotation was too rapid to resolve conformers (small L with bulk similar to that in active drugs) or L was too bulky, allowing formation of only two or three conformers; ΛHT2 was not observed under normal conditions. The (R(4)dt)Pt(d(G*pG*)) results support our initial hypothesis that R(4)dt ligands have Goldilocks bulk, sufficient to slow G* rotation but insufficient to prevent formation of the ΛHT2 conformer. Unlike the (R(4)dt)Pt(5'-GMP)(2) adducts, ROESY spectra of (R(4)dt)Pt(d(G*pG*)) adducts showed no EXSY peaks, a result providing clear evidence that the sugar-phosphodiester backbone slows conformer interchange. Indeed, the ΛHT2 conformer formed and converted to other conformers slowly. Bulkier L (Et(4)dt versus Me(4)dt) decreased the abundance of the ΛHT2 conformer, supporting our initial hypothesis that steric crowding disfavors this conformer. The (R(4)dt)Pt(d(G*pG*)) adducts have a low abundance of the ΔHT1 conformer, consistent with the proposal that the ΔHT1 conformer has an energetically unfavorable phosphodiester backbone conformation; its high abundance when L is bulky is attributed to a small d(G*pG*) spatial footprint for the ΔHT1 conformer. Despite the Goldilocks size of the R(4)dt ligands, the bases in the (R(4)dt)Pt(d(G*pG*)) adducts have a low degree of canting, suggesting that the ligand NH groups characteristic of active drugs may facilitate canting, an important aspect of DNA distortions induced by active drugs.  相似文献   

2.
Complexes of the types LPtCl2 and [L2Pt]X2 [L = substituted 3-(pyridin-2'-yl)-1,2,4-triazine] were synthesized and characterized by NMR spectroscopy and, for the first time, by X-ray crystallography in an effort to determine the coordination properties of this novel class of inorganic medicinal agents possessing HIV-1 virucidal activity. The agents containing either one or two sp2 N-donor bidentate ligands are referred to as ptt (platinum triazine) agents. The X-ray structures of three LPtCl2 compounds revealed the expected pseudo-square-planar geometry. The X-ray structure of [(pyPh2t)2Pt](BF4)2 [pyPh2t = 3-(pyridin-2'-yl)-5,6-diphenyl-1,2,4-triazine] has the expected trans relationship of the unsymmetrical L and is essentially planar, an unusual property for a Pt(II) complex with two bidentate sp2 N donors. HIV-1 is an RNA virus; the guanosine ribonucleoside (Guo) binds (MepyMe2t)PtCl2 at both (inequivalent) available coordination sites to form [(MepyMe2t)Pt(Guo)2]2+ [MepyMe2t = 3-(4'-methylpyridin-2'-yl)-5,6-dimethyl-1,2,4-triazine]. This adduct has four nearly equally intense Guo H8 signals attributed to two pairs of head-to-tail (HT) and head-to-head (HH) conformers, which interchange rapidly within each pair. However, equilibration between pairs requires rotation of the Guo cis to the MepyMe2t pyridyl ring, and the H6' proton on this ring projects toward the Guo and hinders Guo rotation about the Pt-N7 bond. Thus, the HT/HH pairs do not interchange; such behavior is rare. Guo did not add to [(MepyMe2t)2Pt]2+, a result suggesting the possibility that the virucidal activity of LPtCl2 and [L2Pt]2+ ptt agents arises respectively from covalent and noncovalent (possibly intercalative interactions favored by [L2Pt]2+ planarity) binding to biomolecular targets.  相似文献   

3.
Early studies on cis-PtA2(d(G*pG*)) (A2 = diamine or two amines, G = N7-platinated G) and cis-Pt(NH3)2(d(G*pG*)) models for the key cisplatin-DNA cross-link suggested that they exist exclusively or mainly as the HH1 conformer (HH1 = head-to-head G bases, with 1 denoting the normal direction of backbone propagation). These dynamic models are difficult to characterize. Employing carrier A2 ligands designed to slow dynamic interchange of conformers, we found two new conformers, DeltaHT (head-to-tail G* bases with a Delta chirality) and HH2 (with 2 denoting the backbone propagation direction opposite to normal). However, establishing that the non-HH1 conformations exist as an intrinsic feature of the 17-membered Pt(d(G*pG*)) ring requires exploring a range of different carrier ligands. Here we employ the planar aromatic sp(2) N-donor 5,5'-Me(2)bipy (5,5'-dimethyl-2,2'-bipyridine) ligand, having a shape very different from those of previously used nonplanar sp(3) N-donor bidentate carrier ligands, which often bear NH groups. The 5,5'-Me(2)bipy H6 and H6' protons project toward the d(G*pG*) moiety and hinder the dynamic motion of 5,5'-Me(2)bipyPt(d(G*pG*)). We again found HH1, HH2, and DeltaHT conformers with typical properties, supporting the conclusions that the new DeltaHT and HH2 conformers exist universally in dynamic cis-PtA2(d(G*pG*)) adducts, including cis-Pt(NH3)2(d(G*pG*)), and that the carrier ligand typically has little influence on the overall structure of the Pt(d(G*pG*)) macrocyclic ring of a given conformer. The sizes of the G H8 to H6/H6' NOE cross-peaks indicate little base canting in all 5,5'-Me(2)bipyPt(d(G*pG*)) conformers, suggesting that carrier-ligand NH groups favor the canting of one G base in the HH1 and HH2 conformers of typical cis-PtA2(d(G*pG*)) adducts.  相似文献   

4.
We employ retro models, cis-PtA2G2 (A2 = a diamine, G = guanine derivative), to assess the cross-linked head-to-head (HH) form of the cisplatin-DNA d(GpG) adduct widely postulated to be responsible for the anticancer activity. Retro models are designed to have minimal dynamic motion to overcome problems recognized in models derived from cisplatin [A2 = (NH3)2]; the latter models are difficult to understand due to rapid rotation of G bases about the Pt-N7 bond in solution and the dominance of the head-to-tail (HT) form in the solid. Observation of an HH form is unusual for cis-PtA2G2 models. Recently, we found the first HH forms for a cis-PtA2G2 model with A2 lacking NH groups in a study of new Me2ppzPtG2 models. (Me2ppz, N,N'-dimethylpiperazine, has inplane bulk which reduces dynamic motion by clashing with the G O6 as the base rotates into the coordination plane from the ground state position approximately perpendicular to this plane G = 5'-GMP and 3'-GMP.) The finding of an HH form (albeit in a mixture with HT forms) with both G H8 signals unusually downfield encouraged us to study additional Me2ppzPtG2 analogues in order to explain the unusual spectral features and to identify factors that influence the relative stability of HT and HH forms. Molecular modeling techniques suggest HH structures with the H8's close to the deshielding region of the z axis of the magnetically anisotropic Pt atom, explaining the atypical shift pattern. When G = 1-Me-5'-GMP, we obtained NMR evidence that the HH rotamer has a high abundance (34%) and that the three rotamers have nearly equal abundance. These findings and the observation that the relative HT distributions varied little or not at all as a function of pH when G = Guo, 1-MeGuo, or 1-Me-5'-GMP are consistent with two of our earlier proposals concerning phosphate groups in HT forms of cis-PtA2(GMP)2 complexes. We proposed that a G phosphate group can form hydrogen bonds with the cis G N1H ("second-sphere" communication) and (for 5'-phosphate) A2 NH groups. The new results with 1-Me-5'-GMP led us to propose a new role for a 5'-phosphate group; it can also favor the HH form by counteracting the natural preference for the G bases to adopt an HT orientation. Finally, the HH form was also sufficiently abundant to allow observation of a distinct 195Pt NMR signal (downfield of the resonance observed for the HT forms) for several complexes. This is the first report of an HH 195Pt NMR signal for cis-PtA2G2 complexes.  相似文献   

5.
Most simple cis-PtA2G2 complexes that model the G-G cross-link DNA lesions caused by the clinically used anticancer drug cis-PtCl2(NH3)2 undergo large fluxional motions at a rapid rate (A2 = two amines or a diamine; G = guanine derivative). The carrier amine ligands in active compounds have NH groups, but the fundamental role of the NH groups has been obscured by the dynamic motion. To assess carrier ligand effects, we examine retro models, cis-PtA2G2 complexes, in which dynamic motion has been reduced by the incorporation of steric bulk into the carrier ligands. In this study we introduce a new approach employing the chirality-neutral chelate (CNC) ligand, Me2ppz (N,N'-dimethylpiperazine). Because they lie in the Pt coordination plane, the methyl groups of Me2ppz do not clash with the 06 of the base of G ligands in the ground state, but such clashes sterically hinder dynamic motion. NMR spectroscopy provided conclusive evidence that Me2ppzPt(GMP)2 complexes (GMP = 5'- and 3'-GMP) exist as a slowly interconverting mixture of two dominant head-to-tail (HT) conformers and a head-to-head (HH) conformer. Since the absence of carrier ligand chirality precluded using NMR methods to determine the absolute conformation of the two HT conformers, we used our recently developed CD pH jump method to establish chirality. The most abundant HT Me2ppzPt(5'-GMP)2 form had A chirality. Previously this chirality was shown to be favored by phosphate-cis G NIH hydrogen-bonding interligand interactions; such interactions also favor the HT conformers over the HH conformer. For typical carrier ligands, G O6 and phosphate interactions with the carrier ligand NH groups also favor the HT forms. These latter interactions are absent in Me2ppzPt(GMP)2 complexes, but the HT forms are still dominant. Nevertheless, we do find the first evidence for an HH form of a simple cis-PtA2G2 model with A2 lacking any NH groups. In previous studies, the absence of the HH conformer in cis-PtA2G2 complexes lacking carrier NH groups may be due to the presence of out-of-plane carrier ligand bulk. Such bulk forces both G O6-G O6 and G O6-carrier ligand clashes, thereby disfavoring the HH form. The major DNA cross-link adduct has the HH conformation. Thus, for anticancer activity, the small bulk of the NH group may be more important than the H-bonding interaction.  相似文献   

6.
Rapid atropisomerization of cisplatin-DNA cross-link models, cis-PtA(2)G(2) (A(2) = two amines or a diamine, G = guanine derivative, bold font indicating a guanine not linked to another guanine), makes their NMR spectra uninformative. The conformers [two head-to-tail (DeltaHT and LambdaHT) conformers, one head-to-head (HH) conformer] are detected in (CCC)PtG(2) retro models (CCC = chirality-controlling chelates designed to reduce rotation around the G N7 to Pt bond by destabilizing the transition state). Clear trends are found with four CCC ligands, 2,2'-bipiperidine (Bip) and N,N'-dimethyl-2,3-diaminobutane (each with S,R,R,S and R,S,S,R configurations at the chelate ring N, C, C, and N atoms, respectively). S,R,R,S ligands favor left-handed G base canting and the LambdaHT form; R,S,S,R ligands favor right-handed canting and the DeltaHT form. The HH conformer is normally negligible unless G = 5'-GMP. However, understanding this 5'-phosphate effect is complicated by possible interligand interactions of the 5'-phosphate with the N1H of the cis-5'-GMP and a CCC NH; these interactions are referred to as second-sphere (SSC) and first-to-second-sphere (FSC) communication, respectively. We now investigate the four (CCC)PtG(2) models with 1-Me-5'-GMP, a G lacking the N1H needed for SSC. The phosphate location makes FSC possible in the major but not the minor HT form. The major form should increase from pH 3 to pH 7 because the phosphate is deprotonated at pH 7. However, the major DeltaHT form for the R,S,S,R pair did not change in abundance, and the major LambdaHT form for the S,R,R,S pair actually decreased. Thus, FSC is weak. At pH approximately 7 the HH conformer of the S,R,R,S pair had an abundance (40-44%) higher than that in any reported cis-PtA(2)G(2) adduct. FSC involving one 1-Me-5'-GMP could play a role. The high HH abundance and use of a pH jump experiment with (S,R,R,S)-BipPt(1-Me-5'-GMP)(2) allowed us to obtain the first deconvoluted CD spectrum for a cis-PtA(2)G(2) HH conformer. The CD features for the HH conformer are much weaker than for the HT conformers. Our findings are best interpreted to indicate that FSC is not important in aqueous solution, especially for the HT form. Weak FSC is consistent with recent models of the cross-link in duplexes. In contrast, crystals of fluxional models often reveal FSC, but not the more important SSC. SSC was unrecognized until our retro model studies, and the new results reinforce the value of studying retro models for identifying interactions in solution.  相似文献   

7.
Treatment of two precursors, fac-[Re(CO)(3)(L)(CH(3)CN)]BF(4) [L = 5,5'-dimethyl-2,2'-bipyridine (5,5'-Me(2)bipy) (1) and 6,6'-dimethyl-2,2'-bipyridine (6,6'-Me(2)bipy) (2)], with five C(2)-symmetrical saturated heterocyclic amines yielded 10 new amidine complexes, fac-[Re(CO)(3)(L)(HNC(CH(3))N(CH(2)CH(2))(2)Y)]BF(4) [Y = CH(2), (CH(2))(2), (CH(2))(3), NH, or O]. All 10 complexes possess the novel feature of having only one isomer (amidine E configuration), as established by crystallographic and (1)H NMR spectroscopic methods. We are confident that NMR signals of the other possible isomer (amidine Z configuration) would have been detected, if it were present. Isomers are readily detected in closely related amidine complexes because the double-bond character of the amidine C-N3 bond (N3 is bound to Re) leads to slow E to Z isomer interchange. The new fac-[Re(CO)(3)(L)(HNC(CH(3))N(CH(2)CH(2))(2)Y)]BF(4) complexes have C-N3 bonds with essentially identical double-bond character. However, the reason that the Z isomer is so unstable as to be undetectable in the new complexes is undoubtedly because of unfavorable clashes between the equatorial ligands and the bulky N(CH(2)CH(2))(2)Y ring moiety of the axial amidine ligand. The amidine formation reactions in acetonitrile (25 °C) proceeded more easily with 2 than with 1, indicating that the distortion in 6,6'-Me(2)bipy resulting from the proximity of the methyl substituents to the inner coordination sphere enhanced the reactivity of the coordinated CH(3)CN. Reaction times for 1 and 2 exhibited a similar dependence on the basicity and ring size of the heterocyclic amine reactants. Moreover, when the product of the reaction of 1 with piperidine, fac-[Re(CO)(3)(5,5'-Me(2)bipy)(HNC(CH(3))N(CH(2)CH(2))(2)CH(2))]BF(4), was challenged in acetonitrile-d(3) or CDCl(3) with a 5-fold excess of the strong 4-dimethylaminopyridine ligand, there was no evidence for replacement of the amidine ligand after two months, thus establishing that the piperidinylamidine ligand is a robust ligand. This chemistry offers promise as a suitable means for preparing isomerically pure conjugated fac-[(99m)Tc(CO)(3)L](n±) imaging agents, including conjugates with known bioactive heterocyclic amines.  相似文献   

8.
In [PtCl2(cis-1,4-DACH)] (DACH = diaminocyclohexane), the N-Pt-N bite angle (> or =97 degrees , as determined by X-ray diffraction analysis) is much larger than those found in other Pt complexes with bidentate diamines or in cisplatin (approximately 91 degrees ). Hence, the possibility exists that in (cis-1,4-DACH)PtG 2 adducts, rotation of the G's around the Pt-N7 bonds is slowed enough to allow observation of different conformers. In accord with this prevision, decreasing the temperature to 238 K enabled us to observe different conformers of (cis-1,4-DACH)Pt(5'-GMP) 2 (GMP = guanosine monophosphate). This observation is the first case in which such conformers for a platinum derivative with primary diamines and untethered guanines have been resolved and represents the closest model to clinically effective cisplatin obtained to date. We also found that the presence of the 1,4-DACH ligand increased the intensity of the circular dichroism signal stemming from the dominance of an HT conformer (DeltaHT in the adduct with 3'-GMPs and LambdaHT in the adduct with 5'-GMPs).  相似文献   

9.
A series of pyridine-substituted derivatives of octaacetatotetraplatinum(II), [Pt4(CH3COO)8-n(L)2n]n+ (L= 4-dimethylaminopyridine (dmap), pyridine (py), 4-cyanopyridine (cpy); n = 1-4) were prepared, and the tetra- and octasubstituted forms (n = 2 and 4) were isolated. 1HNMR spectra showed that this type of cluster undergoes a comproportionation reaction. Reactions between clusters in which n = 0 and 2, n = 0 and 4, and n = 2 and 4 afforded Pt4 clusters with n = 1, 2, and 3, respectively, as a main product in acetonitrile. The dmap-substituted clusters, trans-[Pt4(CH3COO)6(dmap)4](ClO4)2 x 3CH3NO2 (3a(ClO4)2 x 3CH3NO2) and [Pt4(CH3COO)4(dmap)8](ClO4)4 x 4 H2O (5a(ClO4)4-4H2O), have been structurally characterized. Both 3a and 5a have a square-planar cluster core comprised of four PtII ions, and all eight out-of-plane coordination sites are occupied by acetate ligands in a bridging mode. In 5a, all of the in-plane sites are occupied by dmap ligands. In 3a, four dmap ligands occupy the coordination sites at the two mutually opposite edges of the square planar cluster skeleton, giving a trans tetrasubstituted form of [Pt4(CH3COO)8-] (1). In octasubstituted 5a, adjacent dmap ligands are so closely arranged that the Pt-N distances (2.20(3), 2.30(3) A) are longer than those in tetrasubstituted 3a (2.13(1), 2.15(1) A) and related Pt4 clusters. Furthermore, rotation of the dmap ligand about the Pt-N bond in 5a was restricted, and the rate constant of the rotation was 4.5s(-1) at 20 degrees C from dynamic NMR study. Cluster [Pt4(CH3COO)5(dmap)6]3+ (4a) also exhibited similar hindered rotation with the rate constants of 2.0s(-1), 12s(-1) and approximately 10(4)s(-1) at 20 degrees C depending on the coordination sites of the dmap ligands in 4a.  相似文献   

10.
11.
Substitution reactions of the complexes [Pt(terpy)(H(2)O)](2+), [Pt(terpy)(cyst-S)](2+) and [Pt(terpy)(guo-N(7))](2+), where terpy = 2,2':6',2"-terpyridine, cyst = L-cysteine and guo = guanosine, with some biologically relevant ligands such as inosine (INO), inosine-5'-monophosphate (5'-IMP), guanosine-5'-monophosphate (5'-GMP), l-cysteine, glutathione, thiourea, thiosulfate and diethyldithiocarbamate (DEDTC), were studied in aqueous 0.10 M NaClO(4) at pH 2.5 and 6.0 using variable-temperature and -pressure stopped-flow spectrophotometry. The reactions of [Pt(terpy)(H(2)O)](2+) with INO, 5'-IMP and 5'-GMP showed that these ligands are very good nucleophiles. The second order rate constants varied between 4 x 10(2) and 6 x 10(2) M(-1) s(-1) at 25 degree C. The [Pt(terpy)(cyst-S)](2+) complex is unreactive towards nitrogen donor nucleophiles, and cysteine cannot be replaced by N(7) from INO, 5'-IMP and 5'-GMP. However, sulfur donor nucleophiles such as thiourea, thiosulfate and diethyldithiocarbamate could displace the Pt-cysteine bond. Diethyldithiocarbamate is the best nucleophile and the order of reactivity is: thiourea < thiosulfate < DEDTC with rate constants of 0.936 +/- 0.002, 5.99 +/- 0.02 and 8.88 +/- 0.07 M(-1) s(-1) at 25 degree C, respectively. The reactions of [Pt(terpy)(guo-N(7))](2+) with sulfur donor ligands showed that these nucleophiles could substitute guanosine from the Pt(ii) complex, of which diethyldithiocarbamate and thiosulfate are the strongest nucleophiles. The tripeptide glutathione is also a very efficient nucleophile. Activation parameters (Delta H(++), Delta S(++) and Delta V(++)) were determined for all reactions. The crystal structures of [Pt(terpy)(cyst-S)](ClO(4))(2).0.5H(2)O and [Pt(terpy)(guo-N(7))](ClO(4))(2).0.5guo.1.5H(2)O were determined by X-ray diffraction. Crystals of [Pt(terpy)(cyst-S)](ClO(4))(2).0.5H(2)O are orthorhombic with the space group P2(1)2(1)2(1), whereas [Pt(terpy)(guo-N(7))](ClO(4))(2).0.5guo.1.5H(2)O crystallizes in the orthorhombic space group P2(1)2(1)2. A typical feature of terpyridine complexes can be found in both molecular structures: the Pt-N (central) bond distance, 1.982(7) and 1.92(2) A, respectively, is shorter than the other two Pt-N distances, being 2.043(7) and 2.034(7) A in [Pt(terpy)(cyst-S)](ClO(4))(2).0.5H(2)O and 2.03(2) and 2.04(2) A in [Pt(terpy)(guo-N(7))](ClO(4))(2).0.5guo.1.5H(2)O, respectively. In both crystal structures two symmetrically independent cations representing different conformers are present in the asymmetric unit. The results are analysed in reference to the antitumour activity of Pt(II) complexes, and the importance of the rescue agents are discussed.  相似文献   

12.
Compounds based on the Mn-tda unit (tda=S(CH(2)COO)(2)(-2) ) and N co-ligands have been analyzed in terms of structural, spectroscopic, magnetic properties and DFT calculations. The precursors [Mn(tda)(H(2)O)](n) (1) and [Mn(tda)(H(2)O)(3)]·H(2)O (2) have been characterized by powder and X-ray diffraction, respectively. Their derivatives with bipyridyl-type ligands have formulas [Mn(tda)(bipy)](n) (3), [{Mn(N-N)}(2)(μ-H(2)O)(μ-tda)(2)](n) (N-N=4,4'-Me(2)bipy (4), 5,5'-Me(2)bipy, (5)) and [Mn(tda){(MeO)(2)bipy}·2H(2)O](n) (6). Depending on the presence/position of substituents at bipy, the supramolecular arrangement can affect the metal coordination type. While all the complexes consist of 1D coordination polymers, only 3 has a copper-acetate core with local trigonal prismatic metal coordination. The presence of substituents in 4-6, together with water co-ligands, reduces the supramolecular interactions and typical octahedral Mn(II) ions are observed. The unicity of 3 is also supported by magnetic studies and by DFT calculations, which confirm that the unusual Mn coordination is a consequence of extended noncovalent interactions (π-π stacking) between bipy ligands. Moreover, 3 is an example of broken paradigm for supramolecular chemistry. In fact, the desired stereochemical properties are achieved by using rigid metal building blocks, whereas in 3 the accumulation of weak noncovalent interactions controls the metal geometry. Other N co-ligands have also been reacted with 1 to give the compounds [Mn(tda)(phen)](2)·6H(2)O (7) (phen=1,10-phenanthroline), [Mn(tda)(terpy)](n) (8) (terpy=2,2':6,2'-terpyridine), [Mn(tda)(pyterpy)](n) (9) (pyterpy=4'-(4-pyridyl)-2,2':6,2'-terpyridine), [Mn(tda)(tpt)(H(2)O)]·2H(2)O (10) and [Mn(tda)(tpt)(H(2)O)](2)·2H(2)O (11) (tpt=2,4,6-tris(2-pyridyl)-1,3,5-triazine). Their identified mono-, bi- or polynuclear structures clearly indicate that hydrogen bonding is variously competitive with π-π stacking.  相似文献   

13.
The N7-Pt-N7 adjacent G,G intrastrand DNA cross-link responsible for cisplatin anticancer activity is dynamic, promotes local "melting" in long DNA, and converts many oligomer duplexes to single strands. For 5'-d(A1T2G3G4G5T6A7C8C9C10A11T12)-3' (G3), treatment of the (G3)2 duplex with five pairs of [LPt(H2O)2]2+ enantiomers (L = an asymmetric diamine) formed mixtures of LPt-G3 products (1 Pt per strand) cross-linked at G3,G4 or at G4,G5 in all cases. L chirality exerted little influence. For primary diamines L with bulk on chelate ring carbons (e.g., 1,2-diaminocyclohexane), the duplex was converted completely into single strands (G3,G4 coils and G4,G5 hairpins), exactly mirroring results for cisplatin, which lacks bulk. In sharp contrast, for secondary diamines L with bulk on chelate ring nitrogens (e.g., 2,2'-bipiperidine, Bip), unexpectedly stable duplexes having two platinated strands (even a unique G3,G4/G4,G5 heteroduplex) were formed. After enzymatic digestion of BipPt-G3 duplexes, the conformation of the relatively nondynamic G,G units was shown to be head-to-head (HH) by HPLC/mass spectrometric characterization. Because the HH conformation dominates at the G,G lesion in duplex DNA and in the BipPt-G3 duplexes, the stabilization of the duplex form only when the L nitrogen adducts possess bulk suggests that H-bonding interactions of the Pt-NH groups with the flanking DNA lead to local melting and to destabilization of oligomer duplexes. The marked dependence of adduct properties on L bulk and the minimal dependence on L chirality underscore the need for future exploration of the roles of the L periphery in affecting anticancer activity.  相似文献   

14.
Treatment of cis-Pt(Me2SO)2Cl2 with DNSH-tren afforded [Pt(DNSH-tren)Cl]Cl and with DNSH-dienH, under increasingly more basic conditions, led to Pt(DNSH-dienH)Cl(2), Pt(DNSH-dien)Cl, and Pt(DNS-dien). (DNSH = 5-(dimethylamino)naphthalene-1-sulfonyl, linked via a sulfonamide group to tris(2-aminoethyl)amine (DNSH-tren) and diethylenetriamine (DNSH-dienH); the H's in DNSH-dienH designate protons sometimes lost upon Pt binding, i.e., sulfonamide NH for the dienH moiety and H8 for the DNSH moiety). Respectively, the three neutral DNSH-dienH-derived complexes are difunctional, monofunctional, and nonfunctional and exhibit decreasing fluorescence in this order as the dansyl group distance to Pt decreases. 2D NMR data establish that Pt(DNS-dien) has a Pt-C8 bond and a Pt-N(sulfonamido) bond. Pt(DNSH-dien)Cl and [Pt(DNSH-tren)Cl]Cl bind to N7 of 6-oxopurines (e.g., 5'-GMP, 3'-IMP, and 9-ethylguanine) and sulfur of methionine (met). Competition and challenge reactions for Pt(II) with met and 5'-GMP typically reveal that met binding is favored kinetically but that 5'-GMP binding is favored thermodynamically. This common type of behavior was found for [Pt(DNSH-tren)Cl]Cl. In contrast, Pt(DNSH-dien)Cl had reduced kinetic selectivity for met. This unusual behavior undoubtedly arises as a consequence of the bound Pt-N(sulfonamido) group, which donates strongly to Pt (as indicated by relatively upfield dien NH signals) and which places the bulky DNSH moiety close to the monofunctional reaction site. The decrease in the relatively upfield shifts of the DNSH group signals indicates that this group stacks with the purine. This stacking could explain the unprecedented, relatively low reactivity of a Pt complex bearing a dien-type ligand toward met vs 5'-GMP.  相似文献   

15.
Typical cis-PtA(2)G(2) models of key DNA lesions formed by cis-type Pt anticancer drugs are very dynamic and difficult to characterize (A(2) = diamine or two amines; G = guanine derivative). Retro models have A(2) carrier ligands designed to decrease dynamic motion without eliminating any of three possible conformers with bases oriented head-to-tail (two: DeltaHT and LambdaHT) or head-to-head (one: HH). All three were found in NMR studies of eight Me(2)DABPtG(2) retro models (Me(2)DAB = N,N'-dimethyl-2,3-diaminobutane with S,R,R,S and R,S,S,R configurations at the chelate ring N, C, C, and N atoms, respectively; G = 5'-GMP, 3'-GMP, 5'-IMP, and 3'-IMP). The bases cant to the left (L) in (S,R,R,S)-Me(2)DABPtG(2) adducts and to the right (R) in (R,S,S,R)-Me(2)DABPtG(2) adducts. Relative to the case in which the bases are both not canted, canting will move the six-membered rings closer in to each other ("6-in" form) or farther out from each other ("6-out" form). Interligand interactions between ligand components near to Pt (first-first sphere communication = FFC) or far from Pt (second-sphere communication = SSC) influence stability. In typical cases at pH < 8, the "6-in" form is favored, although the larger six-membered rings of the bases are close. In minor "6-out" HT forms, the proximity of the smaller five-membered rings could be sterically favorable. Also, G O6 is closer to the sterically less demanding NH part of the Me(2)DAB ligand, possibly allowing G O6-NH hydrogen bonding. These favorable FFC effects do not fully compensate for possibly stronger FFC dipole effects in the "6-in" form. SSC, phosphate-N1H cis G interactions favor LambdaHT forms in 5'-GMP and 5'-IMP complexes and DeltaHT forms in 3'-GMP and 3'-IMP complexes. When SSC and FFC favor the same HT conformer, it is present at >90% abundance. In six adducts [four (S,R,R,S)-Me(2)DABPtG(2) and (R,S,S,R)-Me(2)DABPtG(2) (G = 3'-GMP and 3'-IMP)], the minor "6-out" HT form at pH approximately 7 becomes the major form at pH approximately 10, where G N1H is deprotonated, because the large distance between the negatively charged N1 atoms minimizes electrostatic repulsion and probably because the G O6-(NH)Me(2)DAB H-bond (FFC) is strengthened by N1H deprotonation. At pH approximately 10, phosphate-negative N1 repulsion is an unfavorable SSC term. This factor disfavors the LambdaHT R form of two (R,S,S,R)-Me(2)DABPtG(2) (G = 5'-GMP and 5'-IMP) adducts to such an extent that the "6-in" DeltaHT R form remains the dominant form even at pH approximately 10.  相似文献   

16.
The fac-[Re(CO)3(H2O)3]+ cation, the putative DNA-binding species accounting for the biological activity of related Re(I) complexes, binds reversibly to N7 of 6-oxopurine nucleotide monophosphates (NMPs), in contrast to Pt(II) anticancer drugs. A relatively high amount of NMP is needed to convert all of the fac-[Re(CO)3(H2O)3]+ to adducts. The Re/nucleotide 1:1 adduct forms more rapidly and builds up to a higher concentration for guanosine 5'-monophosphate (5'-GMP) and inosine 5'-monophosphate (5'-IMP) than for the respective 3'-monophosphates (3'-GMP and 3'-IMP). These results are attributable to the 5'-positioning of the 5'-NMP phosphate group that allows it to approach the metal inner sphere for more favorable cation electrostatic and aqua ligand H-bonding interactions, both in the initial productive ion pair encounter complexes and in the N7-bound 1:1 adducts. A higher reactivity of 5'-GMP over 3'-GMP is known for cisplatin. In contrast, more Re/nucleotide 1:2 adduct was formed by 3'-GMP (and 3'-IMP) than by 5'-GMP (and 5'-IMP). Because the 3'-phosphate group cannot closely approach the metal inner coordination sphere, the greater stability for the 3'-GMP 1:2 adduct reflects the more favorable G N1H-phosphate interligand GMP-GMP interactions for 3'-GMP vs 5'-GMP (G=guanine base derivative). This type of interaction is known for platinum adducts. In 1:2 adducts the bound nucleotides are inequivalent, prompting us to perform mixed 5'-GMP/3'-GMP experiments, leading to the observation of major (M) and minor (m) mixed Re/5'-GMP/3'-GMP 1:1:1 adducts. The order of abundance at equilibrium in a typical experiment was M>bis 3'-GMP>m>or=bis 5'-GMP. This stability order was rationalized by invoking the phosphate interactions described above. When methionine and 5'-GMP were allowed to compete for fac-[Re(CO)3(H2O)3]+, the Re/5'-GMP 1:1 adduct was the kinetic product and the S-bound Re/methionine adduct was the thermodynamic product, a result opposite to that typically found for cisplatin.  相似文献   

17.
The synthesis and the NMR and X-ray structural characterization of a cisplatin analogue designed to reduce the Pt-N7 rotation of a coordinated guanine base by a billion times are reported. The [Pt(dmm){(+/-)-bip}] (dmm=dimethylmalonato; bip=2,2'-bipiperidine) complex crystallizes in the C2/m space group, which contemplates a mirror plane bisecting the bip and dmm ligands. Because the bip moiety (R, R or S, S configuration at the 2,2'-carbon atoms) does not have planes of symmetry, the requirements of the crystal symmetry are satisfied by a statistical disorder made of bip molecules of R, R or S, S configurations alternating at the same crystallographic site. Such an unexpected arrangement has been permitted by a "quasi planarity" of the bip ligand [maximum deviation from the mean plane through the C and N atoms of 0.2927(9) A], which allows bip molecules of different chiralities to fit in the same space. The bip array of heavy atoms is overlaid, from both sides, by a layer of "quasi axial" (C)H and (N)H atoms (six per side). Those on one side are hydrogen-bonded to the dmm oxygen atoms of another complex molecule joined in a pair. The distance between the average platinum coordination planes is as short as 3.498(1) A, comparable to those found in crystals of the [PtCl 2(bipy)] complex (bipy=2,2'-bipyridine) and of graphite, in which, however, all atoms of each unit are rigorously coplanar and there are no out-of-plane hydrogen atoms. The NMR data show a net chemical shift separation between geminal methylene protons, with the "quasi axial" protons being always at higher field with respect to the "quasi equatorial" ones. This is in accordance with a rigid bip ligand frame and the inability of the bip methylene protons adjacent to the coordinated nitrogen to rotate away from a cis-G base (G=guanine) during G rotation around the Pt-N7 bond.  相似文献   

18.
姜琴  施鹏飞  朱建辉 《无机化学学报》2009,25(10):1764-1770
本文合成了三种铂(Ⅱ)配合物PtLCl(HL1=2-(3′,5′-二甲基-吡唑-1′-基)-N-(2″-吡啶甲基)乙基胺;HL2=2-(3′,5′-二甲基-吡唑-1′-基)-N-(2″-吡啶乙基)乙基胺;HL3=2-(3′,5′-二甲基-吡唑-1′-基)-N-(喹啉-8″-基)乙基胺,通过元素分析和质谱进行结构表征。利用荧光和圆二色光谱研究了3种配合物与小牛胸腺DNA的相互作用,结果发现配体结构对配合物与DNA的作用方式及强度产生极大影响。PtL3Cl具有较大的共轭平面而易以嵌入模式与DNA结合,而PtL1Cl和PtL2Cl的空间位阻较小,易与DNA以共价模式结合。通过质谱跟踪发现,配合物PtL1Cl和PtL2Cl均能与5′-鸟苷酸(5′-GMP)发生共价结合,但是没有发现PtL3Cl与5′-GMP的加合物。3种配合物对人宫颈癌细胞的体外毒活性数据表明:PtL3Cl的细胞毒活性最强。  相似文献   

19.
Capillary electrophoresis-electrospray ionization-mass spectrometry (CE-ESI-MS) has been used for investigating the influence of the sulfur containing amino acid L-methionine (L-Met) on the binding behavior of oxaliplatin (trans-R,R-diaminocyclohexane-(oxalato)platinum(II)) to 5'-GMP. L-Methionine competes with 5'-GMP for the platinum binding site and forms as well as 5'-GMP adducts with oxaliplatin. The formation of the prognosed complexes [Pt(DACH)(L-Met-S,N)]+ and [Pt(DACH)(5'-GMP)2]2- (DACH = 1,2-diaminocyclohexane) could be proved directly by using CE-ESI-MS. Furthermore, we could now bring forward proofs, that the coordination of 5'-GMP with oxaliplatin is inhibited by L-methionine and could show, that the 5'-GMP ligands of the [Pt(DACH) (5'-GMP)2]2- complex can be replaced slowly by L-methionine whereas methionine can not be replaced by GMP.  相似文献   

20.
A series of platinum(II) complexes of the type [Pt(NN)(pyB)2](NO3)2 (NN = bipy, phen; pyB = 3- or 4-pyridineboronic acid) were successfully prepared and fully characterised by 1D- and 2D-multinuclear NMR spectroscopy and ESI-MS. Using VT 1H NMR spectroscopy, rotational isomers for [Pt(NN)(3-pyB)2](NO3)2 were identified and the free energies of activation for rotation of 3-pyB about the Pt-N bond were determined to be DeltaG++310) = 69.2 +/- 0.1 kJ mol(-1) and DeltaG++(305) = 66.0 +/- 0.1 kJ mol(-1) for [Pt(bipy)(3-pyB)2](NO3)2 and [Pt(phen)(3-pyB)2](NO3)2, respectively. The 3- and 4-pyB ligands readily deboronate in boiling H2O to afford [Pt(NN)(py)2](NO3)2; the structure of [Pt(phen)(py)2](2+) (as its PF6- salt) was confirmed by X-ray crystallography. Preliminary thermal denaturation studies revealed only minimal interactions between [Pt(NN)(pyB)2](NO3)2 and calf-thymus DNA and is attributed to hydroxylation of the boronic acid groups at pH 7.4 to afford the corresponding zwitterionic boronate species. This was confirmed by means of variable pH 1H and 11B{1H} NMR spectroscopy.  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