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.
Oehlsen ME  Qu Y  Farrell N 《Inorganic chemistry》2003,42(18):5498-5506
A possible explanation for the low bioavailability of platinum antitumor compounds is their high reactivity with the sulfur-containing tripeptide glutathione (GSH; deprotonated GSH = SG). GSH is located in the intracellular matrix of the cell with a normal concentration of 5-10 mM. In vivo, only a small fraction of the administered drug will migrate into the cell, resulting in relatively high concentrations of GSH compared to that of the drug. The products of the reactions of [[trans-PtCl(NH(3))(2)](2)-mu-[trans-Pt(NH(3))(2)(NH(2)(CH(2))(6)NH(2))(2)]](NO(3))(4) (BBR3464; 1,0,1/t,t,t, n = 6), [[trans-PtCl(NH(3))(2)](2)-mu-(H(2)N(CH(2))(6)NH(2))](NO(3))(2) (BBR3005; 1,1/t,t, n = 6), [[trans-PtCl(NH(3))(2)](2)-mu-(H(2)N(CH(2))(3)NH(2)(CH(2))(4)NH(2))]Cl(3) (BBR3571; 1,1/t,t-spermidine, n = 3, 4), and trans-[PtCl(2)(NH(3))(2)] (t-DDP) with reduced GSH in phosphate-buffered saline (pH 7.35) have been characterized by (1)H, (195)Pt, and (1)H(-)(15)N gradient heteronuclear single-quantum coherence NMR spectroscopy and high-performance liquid chromatography (HPLC) coupled with electrospray ionization time-of-flight mass spectrometry to determine likely metabolites of the complexes with GSH. Chemical shifts (NMR) and retention times (HPLC) established via analysis of the t-DDP profile served as a fingerprint to compare results obtained for the products afforded by the degradation of the polynuclear compounds by GSH. Identical kinetic profiles and chemical shifts between the metabolites and the t-DDP/GSH products allowed identification of the final product for the 1:2 Pt:GSH reaction as a dinuclear species [[trans-Pt(SG)(NH(3))(2)](2)-mu-SG], in which glutathione bridges the two platinum centers via only the sulfur atom.  相似文献   

2.
A series of mixed ammine/amine platinum(IV) complexes with lipophilic ligands in their axial positions were designed, synthesized, and spectrally characterized. In vitro cytotoxicity evaluation of these complexes and their lead compounds have been carried out against A549, SMMC-7721, MCF-7, and SW480 human cancer cell lines. The introduction of carboxylate ions as leaving group can improve the aqueous solubility and stability of the platinum(II) complexes. The carboxylato ligands and chloride ligands in the axial position markedly increased the lipophilicity and cytotoxicity of compounds C4 and C5. Particularly, compound C5 showed two to eight times higher cytotoxicity than cisplatin and satraplatin against selected cell lines. For its oral activity and no cross-resistance potentiality, C5 is expected to be an antitumor platinum drug candidate. This novel class of platinum compounds represents a valuable lead in the development of new-generation agents capable of demonstrating cytotoxicity superior to that of the clinically established cisplatin.  相似文献   

3.
A series of mononuclear cis-diamineplatinum(II) pyrophosphato complexes containing ammine (am), trans-1,2-cyclohexanediamine (dach), and 1,2-ethanediamine (en) as the amine ligands were synthesized and characterized by (31)P and (195)Pt NMR spectroscopy. Chemical shifts of (31)P NMR resonances of these completely deprotonated complexes appear at 2.12, 1.78, and 1.93 ppm, indicating a coordination chemical shift of at least 8 ppm. The (195)Pt NMR chemical shifts for the am and dach complexes were observed at -1503 and -1729 ppm. The complexes are highly stable at neutral pH; no aquation due to the release of either phosphate or amine ligands was observed within 48 h. Furthermore, no partial deligation of the pyrophosphate ligand was detected within several days at neutral pH. At lower pH, however, release of a pyrophosphate ion was observed with concomitant formation of a bridged pyrophosphatoplatinum(II) dinuclear complex. The extended crystal structure containing the dach ligand revealed a zigzag chain stacked in a head-to-tail fashion. Moreover, two zigzag chains are juxtaposed in a parallel fashion and supported by additional hydrogen bonds reminiscent of DNA structures where two strands of DNA bases are held by hydrogen bonds. Theoretical calculations support the notion that the two dinuclear units are held together primarily by hydrogen bonds between the amine and phosphate moieties. Platinum(II) pyrophosphato complexes were readily oxidized by hydrogen peroxide to yield cis-diamine-trans-dihydroxopyrophosphatoplatinum(IV) complexes. Two of these complexes, containing am and en, were characterized by X-ray crystallography. Notable structural features include Pt-O (phosphate) bond distances of 2.021-2.086 A and departures from 180 degrees in trans-HO-Pt-OH bond angles, >90 degrees in O-Pt-O, and >90 degrees in cis-N-Pt-N bond angles. The departure in the trans-HO-Pt-OH angle is more pronounced in the 1,2-ethanediamine complex compared to the dach analogue because of the existence of two molecules possessing enantiomeric conformations within the asymmetric unit. (31)P NMR spectra exhibited well-resolved (195)Pt satellites with coupling constants of 15.4 Hz for the ammine and 25.9 Hz for both the 1,2-ethanediamine and trans-1,2-cyclohexanediamine complexes. The (195)Pt NMR spectrum of the ammine complex clearly showed coupling with two equivalent N atoms.  相似文献   

4.
5.
The chemical shifts and coupling constants of [1,2-15N2]pyrazole, 2-(1-[1,2- 15N2]pyrazolyl)-2-[l,3-2H6]propanol, 1-nitro[1,215N2] and 3-nitro[1,2-15N2]pyrazole are reported.  相似文献   

6.
By the use of [1H,15N] heteronuclear single quantum coherence (HSQC) 2D NMR spectroscopy and electrochemical methods we have determined the hydrolysis profile of the bifunctional dinuclear platinum complex [[trans-PtCl(15NH3)2]2(mu-15NH2(CH2)(6)15NH2)]2+ (1,1/t,t (n = 6), 15N-1), the prototype of a novel class of potential antitumor complexes. Reported are estimates for the rate and equilibrium constants for the first and second aquation steps, together with the acid dissociation constant (pKa1 approximately pKa2 approximately pKa3). The equilibrium constants determined by NMR at 25 and 37 degrees C (I = 0.1 M) were similar, pK1 approximately pK2 = 3.9 +/- 0.2, and from a chloride release experiment at 37 degrees C the values were found to be pK1 = 4.11 +/- 0.05 and pK2 = 4.2 +/- 0.5. The forward and reverse rate constants for aquation determined from this chloride release experiment were k1 = (8.5 +/- 0.3) x 10(-5) s-1 and k-1 = 0.91 +/- 0.06 M-1 s-1, where the model assumed that all the liberated chloride came from 1. When the second aquation step was also taken into account, the rate constants were k1 = (7.9 +/- 0.2) x 10(-5) s-1, k-1 = 1.18 +/- 0.06 M-1 s-1, k2 = (10.6 +/- 3.0) x 10(-4) s-1, k-2 = 1.5 +/- 0.6 M-1 s-1. The rate constants compare favorably with other complexes with the [PtCl(am(m)ine)3]+ moiety and indicate that the equilibrium of all these species favors the chloro form. A pKa value of 5.62 was determined for the diaquated species [[trans-Pt(15NH3)2(H2O)]2(mu-15NH2(CH2)(6)15NH2)]4+ (3) using [1H,15N] HSQC NMR spectroscopy. The speciation profile of 1 and its hydrolysis products under physiological conditions is explored.  相似文献   

7.
The Schiff base ligands I–V, made by condensing either 2-acetylpyridine (I), 8-quinolinecarboxaldehyde (II and III), or o-methylthiobenzaldehyde (IV and V) with either N,N′-dimethyl-1,3-diaminopropane (I, II, and IV), 2-aminomethylpyridine (III), or 2-(2-aminoethyl)-pyridine (V), give ionic PtIVMe3 complexes containing tridentate NNN- or SNN-bonded ligands. With PtMe3Br ligand V gives a neutral complex XI in which it is coordinated only via the two N atoms. A monomeric PtIVMe3 salicyladiminate complex results on treating the dimeric trimethylplatinum(IV) salicylaldehyde complex with the bidentate amine H2N (CH2)3NMe2. The complexes have been fully characterised by 1H NMR spectroscopy.  相似文献   

8.
9.
Guanosine derivatives with a nucleophilic group at the 5' position (G-5') are oxidized by the Pt (IV) complex Pt( d, l)(1,2-(NH 2) 2C 6H 10)Cl 4 ([Pt (IV)(dach)Cl 4]). The overall redox reaction is autocatalytic, consisting of the Pt (II)-catalyzed Pt (IV) substitution and two-electron transfer between Pt (IV) and the bound G-5'. In this paper, we extend the study to improve understanding of the redox reaction, particularly the substitution step. The [Pt (II)(NH 3) 2(CBDCA-O,O')] (CBDCA = cyclobutane-1,1-dicarboxylate) complex effectively accelerates the reactions of [Pt (IV)(dach)Cl 4] with 5'-dGMP and with cGMP, indicating that the Pt (II) complex does not need to be a Pt (IV) analogue to accelerate the substitution. Liquid chromatography/mass spectroscopy (LC/MS) analysis showed that the [Pt (IV)(dach)Cl 4]/[Pt (II)(NH 3) 2(CBDCA-O,O')]/cGMP reaction mixture contained two Pt (IV)cGMP adducts, [Pt (IV)(NH 3) 2(cGMP)(Cl)(CBDCA-O,O')] and [Pt (IV)(dach)(cGMP)Cl 3]. The LC/MS studies also indicated that the trans, cis-[Pt (IV)(dach)( (37)Cl) 2( (35)Cl) 2]/[Pt (II)(en)( (35)Cl) 2]/9-EtG mixture contained two Pt (IV)-9-EtG adducts, [Pt (IV)(en)(9-EtG)( (37)Cl)( (35)Cl) 2] and [Pt (IV)(dach)(9-EtG)( (37)Cl)( (35)Cl) 2]. These Pt (IV)G products are predicted by the Basolo-Pearson (BP) Pt (II)-catalyzed Pt (IV)-substitution scheme. The substitution can be envisioned as an oxidative addition reaction of the planar Pt (II) complex where the entering ligand G and the chloro ligand from the axial position of the Pt (IV) complex are added to Pt (II) in the axial positions. From the point of view of reactant Pt (IV), an axial chloro ligand is thought to be substituted by the entering ligand G. The Pt (IV) complexes without halo axial ligands such as trans, cis-[Pt(en)(OH) 2Cl 2], trans, cis-[Pt(en)(OCOCF 3) 2Cl 2], and cis, trans, cis-[Pt(NH 3)(C 6H 11NH 2)(OCOCH 3) 2Cl 2] ([Pt (IV)(a,cha)(OCOCH 3) 2Cl 2], satraplatin) did not react with 5'-dGMP. The bromo complex, [Pt (IV)(en)Br 4], showed a significantly faster substitution rate than the chloro complexes, [Pt (IV)(en)Cl 4] and [Pt (IV)(dach)Cl 4]. The results indicate that the axial halo ligands are essential for substitution and the Pt (IV) complexes with larger axial halo ligands have faster rates. When the Pt (IV) complexes with different carrier ligands were compared, the substitution rates increased in the order [Pt (IV)(dach)Cl 4] < [Pt (IV)(en)Cl 4] < [Pt (IV)(NH 3) 2Cl 4], which is in reverse order to the carrier ligand size. These axial and carrier ligand effects on the substitution rates are consistent with the BP mechanism. Larger axial halo ligands can form a better bridging ligand, which facilitates the electron-transfer process from the Pt (II) to Pt (IV) center. Smaller carrier ligands exert less steric hindrance for the bridge formation.  相似文献   

10.
The complexes [MCl(2)(kappa2-N approximately N')] (N approximately N' = 2-C(5)H(4)N-CH2-NHAr; Ar = 4-MeC(6)H(4), a; 2,6-Me(2)C(6)H(3), b; 4-MeOC(6)H(4), c; 4-CF(3)C(6)H(4), d; M = Pd, 1a-d; Pt, 2a-d) have been prepared and fully stereochemically characterized both in the solid state and in solution. Their behavior in DMSO-d6 solution is dependent on the substituents of the aryl group and on the metal. Complexes of palladium with substituents at the para position (1a, 1c, 1d) display a dynamic 1H NMR pattern when the solutions are heated. An enantiomeric exchange Slambda/Rdelta is suggested to explain such behavior. On the basis of the calculated negative DeltaS values, an associative mechanism involving the solvent is proposed. Under the same conditions, analogous complexes of platinum (2a, 2c, 2d) proved to be unstable, and release of the N approximately N' ligand was observed. Complexes 1b and 2b show temperature-variable 1H NMR spectra without any evidence accounting for enantiomeric exchange or decoordination. DFT calculations on models of 1a and 1b show that diastereomeric exchange Sdelta/Slambda is a process where the complex with the higher sterical hindrance, 1b, has a lower energy barrier.  相似文献   

11.
In this study the components of the nitrogen chemical shift (CS) tensor are examined for a series of para substituted N,N-dimethylaniline derivatives. This is done through measurement of the 15N NMR spectra of powder samples and through quantum chemical calculations on the isolated molecules. Experiments and calculations show that the isotropic CS, delta(iso), decreases with increasing electron donating ability of the para substituent, in agreement with previous solution studies. More importantly, this study shows that this decrease in the isotropic (solution) CS is due to decreasing values of the CS tensor component delta(11) and component delta(33). The component delta(22) is essentially invariant to the electron donating/withdrawing ability of the para substituent. Through Ramsey's theory of nuclear magnetic shielding, it can be seen that the variation in delta(11) and delta(33), and hence delta(iso), is due to changes in the n-pi* and the sigma-pi* energy gaps in N,N-dimethylaniline. This, in turn, is a result of the change in the energy of the pi* molecular orbital with change in the pi-electron donating ability of the para substituent. The effects of nitrogen inversion on the components of the nitrogen CS tensor components are also discussed. This study also shows the feasibility of performing 15N cross-polarization experiments on nonspinning powder samples at natural isotopic abundance.  相似文献   

12.
1H NMR assignment, including the values of delta(H) and J(H,H) for the cyclopropane moiety, and 13C NMR and 15N NMR spectral data for ciprofloxacin are presented.  相似文献   

13.
According to the density functional theory calculations, the X···H···N (X?N, O) intramolecular bifurcated (three‐centered) hydrogen bond with one hydrogen donor and two hydrogen acceptors causes a significant decrease of the 1hJ(N,H) and 2hJ(N,N) coupling constants across the N? H···N hydrogen bond and an increase of the 1J(N,H) coupling constant across the N? H covalent bond in the 2,5‐disubsituted pyrroles. This occurs due to a weakening of the N? H···N hydrogen bridge resulting in a lengthening of the N···H distance and a decrease of the hydrogen bond angle at the bifurcated hydrogen bond formation. The gauge‐independent atomic orbital calculations of the shielding constants suggest that a weakening of the N? H···N hydrogen bridge in case of the three‐centered hydrogen bond yields a shielding of the bridge proton and deshielding of the acceptor nitrogen atom. The atoms‐in‐molecules analysis shows that an attenuation of the 1hJ(N,H) and 2hJ(N,N) couplings in the compounds with bifurcated hydrogen bond is connected with a decrease of the electron density ρH···N at the hydrogen bond critical point and Laplacian of this electron density ?2ρH···N. The natural bond orbital analysis suggests that the additional N? H···X interaction partly inhibits the charge transfer from the nitrogen lone pair to the σ*N? H antibonding orbital across hydrogen bond weakening of the 1hJ(N,H) and 2hJ(N,N) trans‐hydrogen bond couplings through Fermi‐contact mechanism. An increase of the nitrogen s‐character percentage of the N? H bond in consequence of the bifurcated hydrogen bonding leads to an increase of the 1J(N,H) coupling constant across the N? H covalent bond and deshielding of the hydrogen donor nitrogen atom. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

14.
The 15N chemical shifts and 1H? 15N and 13C? 15N coupling constants of nine monolabelled indazoles were measured and assigned. The experimental values are discussed in terms of the indazolic and iso-indazolic structures, and compared with literature data for other related heterocycles. All the results are consistent with an N-1(H) tautomeric structure for indazole in DMSO-d6.  相似文献   

15.
Four new oxovanadium(IV) compounds were prepared by template reaction of salicyl-, 5-bromosalicyl-and 3-methoxysalicyl-aldehyde S-methylthiosemicarbazones with 2-hydroxy-, 5-bromo-2-hydroxy-and 3-methoxy-2-hydroxy-benzaldehyde in various combinations. The compounds were isolated as stable solid compounds with general formula [VO(L)] and characterized by elemental analysis, conductivity and magnetic measurements, electronic, IR and EPR spectroscopy. The X-band EPR signals recorded from powder forms of all samples have a single asymmetric line shape and theoretical fit studies proved the presence of axial symmetry around the paramagnetic vanadium ions. The anisotropic Lande splitting factors take values of g < g < ge = 2.0023. Orbital energy levels for magnetic electrons were determined from theoretically well fitted Spin Hamiltonian parameters. The EPR spectra recorded from solution forms almost have isotropic character.  相似文献   

16.
Representative polyethyleneimine-based polymers have been studied by 13C- and 15N-nuclear magnetic resonance spectroscopy and an extensive set of chemical shift assignments has been proposed for the complex spectra obtained. The samples in question represent a complex mixture of both protonated and nonprotonated primary, secondary, and tertiary as well as quaternary monomeric units.  相似文献   

17.
We have studied gramicidin A, an environmentally sensitive polymorphic pentadecapeptide, fully 15N-labelled and dispersed in a highly deuterated phospholipid bilayer system. By submitting the sample to fast magic angle spinning, we were able to reduce the polypeptide amide hydrogen linewidths to 160 Hz, and hence to partially resolve them. By correlating these resonances with the 40 Hz wide dipolar coupled 15N in a 2D-CROPSY (cross-polarization spectroscopy) experiment, it was possible to observe the 20 partially overlapping 1H-15N signal pairs from the polypeptide backbone and sidechains. Both chemical shift distributions closely match those of the same peptide in SDS micelles, but only poorly match those of conformationally different gramicidin A in trifluoroethanol, dimethylsulfoxide, or methanol/chloroform mixture. Our results are indicative of the N-to-N right-handed beta6.3-helix conformation of gramicidin A and offer sufficient resolution to encourage development of experiments to measure orientational or distance restraints using through-space dipolar couplings.  相似文献   

18.
A new series of oxovanadium(IV)-lanthanide(III) heteronuclear complexes [Yb(H2O)8]2[(VO)2(TTHA)](3)21 H2O (1), {[Ho(H2O)7(VO)2(TTHA)][(VO)2(TTHA)](0.5)} 8.5 H2O (2), {[Gd(H2O)7(VO)2(TTHA)][(VO)2(TTHA)](0.5)}8.5 H2O (3), {[Eu(H2O)7][(VO)2(TTHA)](1.5)} 10.5 H2O (4), and [Pr2(H2O)6(SO4)2][(VO)2(TTHA)] (5) (H6TTHA=triethylenetetraaminehexaacetic acid) were prepared by using the bulky flexible organic acid H(6)TTHA as structure-directing agent. X-ray crystallographic studies reveal that they contain the same [(VO)2(TTHA)]2- unit as building block, but the Ln3+ ion lies in different coordination environments. Although the lanthanide ions always exhibit similar chemical behavior, the structures of the complexes are not homologous. Compound 1 is composed of a [Yb(H2O)8]3+ ion and a [(VO)2(TTHA)]2- ion. Compounds 2 and 3 are isomorphous; both contain a trinuclear [Ln(H2O)7(VO)2(TTHA)]+ (Ln=Ho for 2 and Gd for 3) ion and a [(VO)2(TTHA)]2- ion. Compound 4 is an extended one-dimensional chain, in which each Eu3+ ion links two [(VO)2(TTHA)]2- ions. For 5, the structure is further assembled into a three-dimensional network with an interesting framework topology comprising V2Pr2 and V4Pr2 heterometallic lattices. Moreover, 4 and 5 are the first oxovanadium(IV)-lanthanide(III) coordination polymers and thus enlarge the realm of 3d-4f complexes. The IR, UV/Vis, and EPR spectra and the magnetic properties of the heterometallic complexes were studied. Notably, 2 shows unusual ferromagnetic interactions between the VO2+ and Ho3+ ions.  相似文献   

19.
Platinum(IV) halides formed complexes of the type PtL2X4 [L=1-vinyl imidazole (1,-VIm), 1-methylimidazole (1-MeIm), 1,2-dimethylimidazole (1,2-Me2Im), 1-vinyl-2-methylimidazole (1-V-2-MeIm), 2-methylimidazole (2-MeIm), 2-ethylimidazole (2-EtIm), 2-isopropylimidazole (2-i-PrIm), and 4-methylimidazole (4-MeIm); X=Cl, Br] in neutral aqueous solution. The 1-n-butylimidazole (1-n-BuIm) ligand yielded only (LH)2PtX6 compound in the same medium. The compounds were characterised by elemental analyses, IR, UV-VIS and 1HNMR spectra.  相似文献   

20.
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