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 93 毫秒
1.
Detailed studies were carried out on the binding of the enantiomers of [PtCl2(mepyrr)] (mepyrr = N-methyl-2-aminomethylpyrrolidine) to dG, d(GpG) and a 52-mer oligonucleotide. The pyrrolidine ligand structure was found to be neither sufficiently rigid nor bulky to enforce a single chirality at the exocyclic amine site in this complex, resulting in the presence of diastereomers that complicated the binding studies. Reaction of the (GpG) dinucleotide with R- and S-[PtCl2(mepyrr)] resulted in formation of four [Pt{d(GpG)}(mepyrr)] isomers for each enantiomer as a consequence of the existence of two orientational isomers and two diastereomers. These isomers formed in different amounts most likely as a consequence of the unequal formation of the diastereomers together with stereoselectivity induced by interactions between the dinucleotide and the mepyrr ligand. The [PtCl2(mepyrr)] complexes displayed stereoselectivity and enantioselectivity in their reactions with a 52-mer duplex designed to allow formation of only GpG intrastrand adducts. All four bifunctional adducts formed for each enantiomer, providing further evidence of the lack of directing ability of the ligand in formation of the 1,2-intrastrand adduct. Significant amounts of monofunctional species remained in these assays suggesting that the introduction of the methyl substituent to the exocyclic amine inhibited ring-closure to the bifunctional adduct. This was not sufficient to achieve enantiospecificity, but in the case of the R-enantiomer, one of the bifunctional adducts formed in only small amounts.  相似文献   

2.
The rate and extent of binding of [PtCl2(hpip)] (hpip=homopiperazine-1,4-diazacycloheptane) and cis-[PtCl2(NH3)2] to calf thymus DNA was measured using atomic absorption spectroscopy and it was found that [PtCl2(hpip)] bound both more rapidly and to a greater extent than did cis-[PtCl2(NH3)2]. The binding of [PtCl2(hpip)] and [PtCl2(en)] (en=ethane-1,2-diamine) to salmon sperm DNA and to synthetic, self-complementary 10-base-pair and 52-base-pair oligonucleotides was studied using enzymatic digestion and HPLC analysis of the products. [PtCl2(hpip)] forms approximately two-fold fewer GpG and ApG intrastrand adducts and concomitantly more monofunctional adducts than does [PtCl2(en)]. In the case of [PtCl2(hpip)], two GpG adducts, corresponding to the different orientations of the hpip ligand with respect to the DNA, were observed in a 1:3.3 ratio. The minor product corresponds to the orientation in which the bulkier propylene chain of the hpip ligand is adjacent to, and makes close contacts with, the floor of the major groove. When the reaction was repeated with a synthetic oligonucleotide decamer duplex, the ratio of the two forms was approximately 1:1.9 and with the 52-mer duplex it was 1:2.4, revealing an apparent systematic dependence of stereoselectivity on nucleotide size. Computer modeling of the two adducts formed by [PtCl2(hpip)] and those formed by [PtCl2(en)] and cis-[PtCl2(NH3)2] revealed that non-bonded interactions between the hpip ligand and the DNA were probably responsible for both the decreased proportion of GpG adducts formed by [PtCl2(hpip)] and the stereoselectivity exhibited in the formation of these adducts. This is the first case in which the stereoselectivity can be ascribed to steric factors alone.  相似文献   

3.
The reaction of H2[PtCl6] · 6H2O and (H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O (18C6 = 18-crown-6) with 9-methylguanine (MeGua) proceeded with the protonation of MeGua forming 9-methylguaninium hexachloroplatinate(IV) dihydrate (MeGuaH)2[PtCl6] · 2H2O (1).The same compound was obtained from the reaction of Na2[PtCl6] with (MeGuaH)Cl.On the other hand, the reaction of guanosine (Guo) with (H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O in methanol at 60 °C proceeded with the cleavage of the glycosidic linkage and with ligand substitution to give a guaninium complex of platinum(IV), [PtCl5(GuaH)] · 1.5(18C6) · H2O (2).Within several weeks in aqueous solution a slow reduction took place yielding the analogous guaninium platinum(II) complex, [PtCl3(GuaH)] · (18C6) · 2Me2CO (3).H2[PtCl6] · 6H2O and guanosine was found to react in water, yielding (GuoH)2[PtCl6] (4) and in ethanol at 50 °C, yielding [PtCl5(GuoH)] · 3H2O (5).Dissolution of complexes 2 and 5 in DMSO resulted in the substitution of the guaninium and guanosinium ligands, respectively, by DMSO forming [PtCl5(DMSO)].Reactions of 1-methylcytosine (MeCyt) and cytidine (Cyd) with H2[PtCl6] · 6H2O and(H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O resulted in the formation of hexachloroplatinates with N3 protonated pyrimidine bases as cation (MeCytH)2[PtCl6] · 2H2O (6) and (CydH)2[PtCl6] (7), respectively. Identities of all complexes were confirmed by 1H, 13C and 195Pt NMR spectroscopic investigations, revealing coordination of GuoH+ in complex 5 through N7 whereas GuaH+ in complex 3 may be coordinated through N7 or through N9. Solid state structure of hexachloroplatinate 1 exhibited base pairing of the cations yielding (MeGuaH+)2, whereas in complex 6 non-base-paired MeCytH+ cations were found. In both complexes, a network of hydrogen bonds including the water molecules was found. X-ray diffraction analysis of complex 3 exhibited a guaninium ligand that is coordinated through N9 to platinum and protonated at N1, N3 and N7. In the crystal, these NH groups form hydrogen bonds N–HO to oxygen atoms of crown ether molecules.  相似文献   

4.
Cobalt(III) complexes with a thiolate or thioether ligand, t-[Co(mp)(tren)]+ (2), t-[Co(mtp)(tren)]2+ (1Me) and t-[Co(mta)(tren)]2+ (2Me), (mp = 3-mercaptopropionate, MA = 3-(methylthio)propionate and MTA = 2-(methylthio)acetate) have been prepared in aqueous solutions. The crystal structures of 1, 2, 1Me and 2Me were determined by X-ray diffraction methods. The crystal data are as follows, t-[Co(mp)(tren)]ClO4 (1CIO4): monoclinic, P21/n, A = 10.877(8), B = 11.570(4), c = 12.173(7) Å, β = 92.20(5)°, V = 1531(1) Å3, Z = 4 and R = 0.060; t-[Co(ma)(tren)]Cl·3H2O (2Cl·3H2O): monoclinic, P21/n, a = 7.7688(8), B = 27.128(2), C = 7.858(1) Å, β = 100.63(1)°, V = 1627.7(3) Å3, Z = 4 and R = 0.066; (+)465CD-t-[Co(mtp)(tren)](ClO4)2 ((+)465CD-1Me(ClO4)2): orthorhombic, P212121, A = 10.6610(7), B = 11.746(1), C = 15.555(1) Å, V = 1947.9(3) Å3, Z = 4 and R = 0.068; (+)465CD-t-[Co(mta)(tren)](ClO4)2 ((+)465CD-2Me(ClO4)2): orthorhombic, P212121, a = 10.564(1), B = 11.375(1), C = 15.434(2) Å, V = 1854.7(4) Å3, Z = 4 and R = 0.047. All central Co(III) atoms have approximately octahedral geometry, coordinated by four N, one O, and one S atoms. All of the complexes are only isomer, of which the sulfur atom in the didentate-O,S ligands are located at the trans position to the tertiary amine nitrogen atom of tren. 1 and 1Me contain six-membered chelate ring, and 2 and 2Me do five-membered chelate ring in the didentate ligand. The chirality of the asymmetric sulfur donor atom in (+)465CD-1Me is the S configuration and that in (+)465CD-2Me is the R one. The 1H NMR, 13C NMR and electronic absorption spectral behaviors and electrochemical properties of the present complexes are discussed in relation to their stereochemistries.  相似文献   

5.
The chloro complexes trans-[Pt(Me)(Cl)(PPh3)2], after treatment with AgBF4, react with 1-alkynes HC---C---R in the presence of NEt3 to afford the corresponding acetylide derivatives trans-[Pt(Me) (C---C---R) (PPh3)2] (R = p-tolyl (1), Ph (2), C(CH3)3 (3)). These complexes, with the exception of the t-butylacetylide complex, react with the chloroalcohols HO(CH2)nCl (n = 2, 3) in the presence of 1 equiv. of HBF4 to afford the alkyl(chloroalkoxy)carbene complexes trans-[Pt(Me) {C[O(CH2)nCl](CH2R) } (PPh3)2][BF4] (R = p-tolyl, N = 2 (4), N = 3 (5); R=Ph, N = 2 (6)). A similar reaction of the bis(acetylide) complex trans-[Pt(C---C---Ph)2(PMe2Ph)2] with 2 equiv. HBF4 and 3-chloro-1-propanol affords trans-[Pt(C---CPh) {C(OCH2CH2CH2Cl)(CH2Ph) } (PMe2Ph)2][BF4] (7). T alkyl(chloroalkoxy)-carbene complex trans-[Pt(Me) {C(OCH2CH2Cl)(CH2Ph) } (PPh3)2][BF4] (8) is formed by reaction of trans-[Pt(Me)(Cl)(PPh3)2], after treatment with AgBF4 in HOCH2CH2Cl, with phenylacetylene in the presence of 1 equiv. of n-BuLi. The reaction of the dimer [Pt(Cl)(μ-Cl)(PMe2Ph)]2 with p-tolylacetylene and 3-chloro-1-propanol yields cis-[PtCl2{C(OCH2CH2CH2Cl)(CH2C6H4-p-Me}(PMe2Ph)] (9). The X-ray molecular structure of (8) has been determined. It crystallizes in the orthorhombic system, space group Pna21, with a = 11.785(2), B = 29.418(4), C = 15.409(3) Å, V = 4889(1) Å3 and Z = 4. The carbene ligand is perpendicular to the Pt(II) coordination plane; the PtC(carbene) bond distance is 2.01(1) Å and the short C(carbene)-O bond distance of 1.30(1) Å suggests extensive electronic delocalization within the Pt---C(carbene)---O moietry.  相似文献   

6.
In the present work, we measured survival and the platinum on the genome after treatment of repair-proficient or repair-deficient Escherichia coli strains with trans-[PtCl2(E-iminoether)2] and compared these results with the effects of “classical” cisplatin. We found that toxicity of antitumor trans-[PtCl2(E-iminoether)2] in repair-deficient trains was much less than that of cisplatin. This markedly reduced toxicity was not a consequence of the reduced uptake or low levels of DNA binding in the bacteria cells but rather appeared to reflect DNA binding mode of this trans-platinum drug different from that of cisplatin.  相似文献   

7.
Reaction of 4,6-dimethylpyrimidine-2(1H)-thione (Me2pymSH) with mer-[ReOCl3(Me2S)(OPPh3)] synthon in 1:1 molar ratio in refluxing acetone, results in the replacement of the Me2S ligand to form the mer-[ReOCl3(Me2pymSH)(OPPh3)] species. X-ray diffraction shows that the structure of the title compound consists of monomeric units with a distorted octahedral coordination around the rhenium(V) centre which includes the axial ReO and Re---OPPh3 bonds, and in which three Cl ions and a S-monodentate neutral Me2pymSH ligand act as equatorial ligands. The compound was also characterised using electrochemical measurements and UV–Vis–NIR and IR spectroscopy.  相似文献   

8.
Cytotoxicity and mutagenicity of trans,trans,trans-[PtCl2(CH3COO)2(NH3)(1-adamantylamine)] [trans-adamplatin(IV)] and its reduced analog trans-[PtCl2(NH3)(1-adamantylamine)] [trans-adamplatin(II)] were examined. In addition, the several factors underlying biological effects of these trans-platinum compounds using various biochemical methods were investigated. A notable feature of the growth inhibition studies was the remarkable circumvention of both acquired and intrinsic cisplatin resistance by the two lipophilic trans-compounds. Interestingly, trans-adamplatin(IV) was considerably less mutagenic than cisplatin. Consistent with the lipophilic character of trans-adamplatin complexes, their total accumulation in A2780 cells was considerably greater than that of cisplatin. The results also demonstrate that trans-adamplatin(II) exhibits DNA binding mode markedly different from that of ineffective transplatin. In addition, the reduced deactivation of trans-adamplatin(II) by glutathione seems to be an important determinant of the cytotoxic effects of the complexes tested in the present work. The factors associated with cytotoxic and mutagenic effects of trans-adamplatin complexes in tumor cell lines examined in the present work are likely to play a significant role in the overall antitumor activity of these complexes.  相似文献   

9.
The syntheses of nitrosyl–dimethylsulfoxide–ruthenium(II) complexes with general formula mer-[RuCl3(L)(DMSO)(NO)] (L=DMSO or CD3CN) is reported. The mer-[RuCl3(DMSO)2(NO)] (1) complex was obtained from the reaction of [RuCl3(NO)] with the sulfoxide ligand in acetone. The mer-[RuCl3(CD3CN)(DMSO)(NO)] (2) compound was obtained from mer-[RuCl3(DMSO)2(NO)] maintained in deuterated acetonitrile. These data suggest a slow kinetic reaction due the low lability of the DMSO ligand coordinated to the {RuII–NO+} species. The crystal and molecular structures of (1) and (2) have been determined from X-ray studies. Crystal data: for (1), monoclinic, P21/c, a=8.8340(2) Å, b=12.0230(3) Å, c=13.7064(4) Å, β=114.546(2)°, Z=4, R1=0.0429; for (2), monoclinic, P21/n, a=10.0180(7) Å, b=9.5070(7) Å, c=13.3340(9) Å, β=102.264(4)°, Z=4, R1=0.0472. The spectroscopic characterization of (1), in solid state (infrared spectrum) and in solution (nuclear magnetic resonance and cyclic voltammetry) is also described.  相似文献   

10.
The cancer chemotherapeutic drug cis-diamminedichloroplatinum(II) (cis-DDP) is active as a result of its bifunctional reactions with DNA. Many other platinum complexes also have therapeutic activity. Of current interest are complexes containing 1,2-diaminocyclohexane (DACH). The DACH ligand exists in three isomeric forms with reported differences in therapeutic activity in the order R,R greater than S,S greater than R,S-DACH-Pt. The reaction of the sulphate form of each of these three isomers with DNA has been characterized as a possible explanation for the apparent differences in antitumor activity. These reactions have been characterized by platinating pure DNA followed by enzyme digestion, HPLC separation and analysis by atomic absorption and nuclear magnetic resonance. The spectrum of adducts produced was similar for each isomer and similar to that reported for cis-DDP with adduction at d(GpG), d(ApG) and (dG)2. The R,S-isomer additionally demonstrated isomeric adducts at d(GpG) and d(ApG). The kinetics of formation of the various adducts was the same for each isomer; total platination of DNA was complete in 15 min as were bifunctional adducts at d(GpG) and (dG)2. However, rearrangement to bifunctional adducts took several hours in the case of adducts at d(ApG) sequences. These results did not provide a reason for the different activities of the isomers. It is suggested that the interaction of these adducts with metabolic processes such as DNA repair might explain these differences.  相似文献   

11.
Duplex oligonucleotides containing a single intrastrand [Pt(NH3)2]2+ cross-link or monofunctional adduct and either 15 or 22 bp in length were synthesized and chemically characterized. The platinum-modified and unmodified control DNAs were polymerized in the presence of DNA ligase and the products studied on 8% native polyacrylamide gels. The extent of DNA bending caused by the various platinum-DNA adducts was revealed by their gel mobility shifts relative to unplatinated controls. The bifunctional adducts cis-[Pt(NH3)2[d(GpG)]]+, cis-[Pt(NH3)2[d(ApG)]]+, and cis-[Pt(NH3)2[d(G*pTpG*)]], where the asterisks denote the sites of platinum binding, all bend the double helix, whereas the adduct trans-[Pt(NH3)2[d(G*pTpG*)]] imparts a degree of flexibility to the duplex. When modified by the monofunctional adduct cis-[Pt(NH3)2(N3-cytosine)(dG)]Cl the helix remains rod-like. These results reveal important structural differences in DNAs modified by the antitumor drug cisplatin and its analogs that could be important in the biological processing of the various adducts in vivo.  相似文献   

12.
Unsymmetrical di(phosphine) ligands (dpp)2Rop (1a, b = bis(diphenylphosphino)-2-alkyl-3-oxapropane (alkyl = methyl and ethyl)) and (dpp)2oCy (1c = trans-2-diphenylphosphinocyclohexyl diphenylphosphinite) and their Pt(II) dichloride complexes, PtCl2((dpp)2mop) (2a), PtCl2((dpp)2eop) (2b) and PtCl2((dpp)2oCy) (2c), have been synthesized and characterized by NMR spectroscopy. The crystal structures of 2b and 2c show that the geometry about the platinum centers is square planar. In 2b, the metal and di(phosphine) ligand chelate ring are in a chair conformation, whereas in 2c, the chelate ring conformation is a skewed boat. Initial reaction of sodium borohydride with 2a, b, c yields the monohydride monochloride complexes PtHCl((dpp)2mop) (5a), PtHCl((dpp)2eop) (5b) and PtHCl((dpp)2oCy) (5c). At longer reaction times, fluxional dimeric species are obtained, [PtH((dpp)2mop)]2 (4a), [PtH((dpp)2eop)]2 (4b) and [PtH((dpp)2oCy)]2 (4c),and in the case of 4c two different isomers exist. The dihydride complexes PtH2((dpp)2mop) (3a), PtH2((dpp)2eop) (3b) and PtH2((dpp)2oCy) (3c), are prepared by further reaction of NaBH4 and 2. Hydrogen cycling is facile in the dihydride complexes 3a, b, c, and oxidative addition of H2 proceeds in a pairwise manner as determined by the observation of parahydrogen induced polarization (PHIP) in the 1H NMR spectra. The reductive elimination of H2 is also shown to be concerted by reaction of dihydride complexes with D2. Crystal data: 2b (C30H32Cl6OP2Pt), monoclinic, space group P21/c (No. 14), a = 13.7040(1), b = 11.3430(7), c = 21.3880(9) Å, β = 97.923(9)°, V = 3292.9(2) Å3 and Z = 4; 2c (C30H30Cl2OP2Pt), monoclinic, space group P21 (No. 4), a = 11.7360(2), b = 8.4311(2), c = 14.2789(2) Å, β = 101.290(1)°, V = 1385.52(4) Å3 and Z = 2.  相似文献   

13.
The reaction of cis-[PtCl2(PPh3)2] with trisubstituted thioureas [R1R2NC(=S)NHR3] in refluxing methanol with triethylamine base, followed by addition of NaBPh4 gives the salts [Pt{SC(=NR1R2)NR3}(PPh3)2]BPh4 in high yield; a range of thiourea substituents, including chiral, fluorescent and chromophoric groups can be incorporated. The azo dye-derived complex [Pt{SC(=N(CH2CH2)2O)NC6H4N=NC6H4NMe2}(PPh3)2]BPh4 has been characterised by a single-crystal X-ray diffraction study. The formation of a fluorescein-derivatised platinum–thiourea complex is also described. Reaction of cis-[PtCl2(PPh3)2] with PhNHC(S)NHPh or EtNHC(S)NHEt, triethylamine and NaBPh4 gives, respectively, [Pt{SC(=NHPh)NPh}(PPh3)2]+ and the known cation [Pt{SC(=NHEt)NEt}(PPh3)2]+, isolated as tetraphenylborate salts. Reaction of cis-[PtCl2(PPh3)2] with an excess of Na[MeNHC(S)NCN] in methanol gives the bis(thiourea monoanion) complex trans-[Pt{SC(=N---CN)NHMe}2(PPh3)2], characterised by NMR spectroscopy and an X-ray crystal structure determination. When cis-[PtCl2(PPh3)2] is reacted with 1 equiv. of Na[MeNHC(S)N---CN] in methanol, with added NaBPh4, a mixture of isomers of the [Pt{SC(=NHCN)NMe}(PPh3)2]+ cation is obtained.  相似文献   

14.
Condensation of Z-PPh2CH2C(But)=NNH2 with 4-nitroacetophenone gave the azine phosphine Z,E-PPh2CH2C(But)=N-N=CMe(C6H4NO2-4) (I). The corresponding phsophine oxide II was prepared by treatment of I with H2O2. The phosphine I with [Mo(CO)4(nbd)] (nbd=norbornadiene) gave [Mo(CO)4{PPh2CH2C(But)=N-N=CMe(C6H4NO2-4)}] (1a); the corresponding tungsten 1b and chromium 1c complexes were made similarly. The crystal structure of 1a was determined by X-ray diffraction and showed the presence of a six-membered chelate ring with the bulky 4-nitrophenyl group held close to the metal. Oxidation of 1a with bromine gave the seven-coordinate molybdenum (II) complex 2. Treatment of [PtMe2(cod)] (cod=cycloocta-1,5-diene) with I at 20°C gave the dimethyl-platinum (II) complex [PtMe2{PPh2CH2C(But)=N-N=CMe(C6H4NO2-4)}] (3a) which with MeI gave the iodotrimethylplatinum(IV) complex 4. Treatment of 3a with C≡O opened the chelate ring to give the dimethyl(carbonyl)platinum(II) complex 5 containing a monodentate phosphine ligand. When 3a was heated in toluene solution at 110°C it gave the cyclometallated methylplatinum(II) complex [PtMe{PPh2CH2C(But)=N-N=CMe(C6H3NO2-4)}] (6). Treatment of 6 with MeI gave the platinum(IV) complex 7. The dichloropalladium(II) complex [PdCl2{PPh2CH2C(But)=N-N=CMe(C6H4NO2-4)}] (3b) was prepared by treatment of [PdCl2(NCPh)2] with I in CH2Cl2. Treatment of [PtCl2(NCMe)2] with 2 equiv. of I gave the trans-bis(phosphine) complex 8. When 2 equiv. of I were treated with [PtCl2(cod)] followed by NH4PF6 this gave the salt 9a containing two six-membered chelate rings; the analogous palladium(II) 9b) salt was also prepared. Treatment of 2 equiv. of I with [PtCl2(cod)] followed by NH4PF6 gave the PF6 salt 10 containing a six-membered chelate ring and a monodentate ligand. When 10 was treated with AgNO3 followed by NH4PF6 this gave the bis-chelate complex 11 containing five- and six-membered chelate rings. Treatment of [IrCl(CO)2(p-toluidine)] with I gave the cyclometallated iridium(III) hydride complex [IrHClCO{PPh2CH2C(But)=N-N=CMe(C6H3NO2-4)}] (12). [RuCl2(PPh3)3] with the phosphine I resulted in the Ru(II) complex 13 in which the ortho hydrogens of the 4-nitrophenyl group are agostically interacting with ruthenium. Proton, Phosphorus-31, some carbon-13 NMR and IR data have been obtained. Crystals of 1a are orthorhombic, space group Pna21, with a = 1819.3(2), b = 1050.0(1), c = 1614.8(2) pm and Z = 4; final R = 0.0191 for 2616 observed reflections.  相似文献   

15.
The reaction of the monoalkyl complex trans-[Pt(DMSO)2Cl(CH3)] with a large variety of heterocyclic nitrogen bases L, in chloroform solution, leads to the formation of uncharged complexes of the type [Pt(DMSO)(L)Cl(CH3)], containing four different groups coordinated to the metal center. Only two out of the three different possible isomers were detected in solution. These two trans(C,N) and cis(C,N) species can be unambiguously identified through 1H NMR spectroscopy. For the trans(C,N) isomers, average values of 2JPtH=75±4 Hz and 3JPtH=36±4 Hz have been observed for the coordinated methyl and DMSO ligands, respectively. In the case of the cis(C,N) isomers, these values increase to 2JPtH=83±2 Hz, and decrease to 3JPtH=26±3 Hz due to the mutual exchange of ligands in trans position to CH3 and DMSO. In the case of bulky asymmetric ligands, such as quinoline, 2-quinolinecarboxaldehyde, 2-methylquinoline, 5-aminoquinoline, 2-phenylpyridine and 2-chloropyridine, slow rotation of the hindered group around the Pt---N bond makes the coordinated DMSO ligand prochiral. NMR experiments have shown that the first reaction product is the trans(C,N) isomer as a consequence of the very fast removal of one DMSO ligand by the nitrogen bases from the starting complex trans-[Pt(DMSO)2Cl(CH3)]. This trans kinetic product undergoes a geometrical conversion into the more stable cis(C,N) isomer through the intermediacy of fast exchanging aqua-species. The rate of isomerization and the relative stability of the two isomers depends essentially on the rate of aquation and on the steric congestion imposed by the new L ligand on the metal.  相似文献   

16.
Fulvenes (1a–e) derived from condensation of cyclopentadiene with acetone or a variety of aldehydes were treated with LiPAr2 (Ar = phenyl, p-tolyl) to yield the respective substituted (diarylphosphinomethyl)cyclopentadienides (2, 3). Subsequent reaction with ZrCl4(THF)2 gave the respective bis[(diarylphosphinomethy])cyclopentadienyl]zirconium dichlorides ( Ar = phenyl (4), p-tolyl (5)). The complex rac-[C5H4-CH(CH3)-PPh2]2ZrCl2 (rac-4b) was characterized by X-ray diffraction. The reaction of complexes 4a and 5a [(Cp-CMe2-PAr2)2ZrCl2] with PdCl2(NCPh)2 or PtCl2(NCPh)2 leads to the formation of the trans-(metallocene-chelate-phosphane)metal complexes 6–9 (e.g. trans-Cl2Pd(Ph2P-CMe2-Cp)2ZrCl2]. Chloride abstraction from the reaction product of [Cp-CH(CMe3)PPh2]2ZrCl2 with PdCl2(NCPh)2 eventually gave the cationic complex [meso,trans-(Cp-CH(CMe3)PPh2)2(Cl)Zr(μ-Cl)Pd(Cl)]+ (10) that was also characterized by X-ray diffraction. It features a dimetallabicyclic framework with two Cp-CHR-PPh2 ligands and a chloride bridging between the early and the late transition metal center.  相似文献   

17.
The platinum-DNA adduct profile formed by the R- and S-enantiomers of [PtCl2(ahaz)] (ahaz = 3(R)-aminohexahydroazepine or 3(S)-aminohexahydroazepine) on reaction with salmon sperm DNA were characterised using HPLC and GFAAS (graphite furnace atomic absorption spectrometry) analyses. At a platinum to nucleotide ratio (Rt) equalling 0.05, the R-enantiomer forms a substantially larger amount (approximately 60%) of monofunctional adducts than the S-enantiomer (less than 35%). Fewer intrastrand GpG adducts are formed by the R-enantiomer (approximately 21%) than the S-enantiomer (approximately 37%). For both enantiomers, two isomeric GpG adducts, corresponding to the different orientations of the primary amine of ahaz ligand with respect to the O6 atom of the 5′ guanine, were observed in the ratios of 1:1.3 and 1:4.3 for the R- and S-enantiomers, respectively. The reasons for this enantioselectivity and stereoselectivity are discussed.  相似文献   

18.
-erythro-5,6,7,8-Tetrahydrobiopterin (BH4), which is the cofactor of aromatic amino acid hydroxylases, plays an important role in the biosyntheses of monoamine neurotransmitters. BH4 exists as natural (6R)- and unnatural (6S)-isomers. In our previous reports, only (6R)-isomer significantly stimulated cofactor activity for tyrosine, tryptophan and phenylalanine hydroxylases (TH, TPH, PAH) in whole animals or in tissue slices. In this study we have compared the in situ cofactor activity on TH between natural (6R)- and unnatural (6S)-isomers in clonal cells. We have transfected human TH type 2 cDNA into the normal rat kidney (NRK) fibroblasts. These cells expressed TH protein, but had neither DOPA decarboxylase (DDC) nor BH4. Thus, TH activity was observed only in the presence of exogenous BH4. We compared the difference in in situ DOPA formation by TH activity in the presence of (6R)- or (6S)-BH4 in the human TH-transfected cells. The effect of exogenous BH4 was also compared between (6R)- and (6S)-isomers in rat pheochromocytoma PC12h cells, which contained approximately 100 μM endogenous (6R)-BH4. The rate of uptake of both BH4 isomers into these cells increased in proportion to the pterin cofactor concentrations in the incubation medium up to 400 μM but was nearly saturated at 1 mM BH4. TH-transfected NRK fibroblasts formed DOPA only in the presence of exogenously added (6R)- or (6S)-BH4 dose-dependently and released DOPA into the medium. At a saturating concentration of 1 mM, (6R)-BH4 was approximately three times as active as (6S)-BH4. In contrast, in PC12h cells which contained endogenous (6R)-BH4 (approximately 100 μM), exogenous (6R)-BH4 activated DOPA formation maximally at 500 μM about 10-fold, while (6S)-BH4 activated it only slightly, about 2.5-fold. These results suggest that (6S)-isomer has lower cofactor activity with TH in the cells than (6R)-isomer. This TH transfected fibroblasts should be useful to assess cofactor activities of tetrahydropteridines in the cell.  相似文献   

19.
The asymmetrical platinum complex [PtCl2(N,N-dmen)] (N,N-dmen = N,N-dimethylethylenediamine) reacts with the dinucleotide GpG to form two isomeric chelates of the formula [Pt(N,N-dmen)(GpG)]+ [9]. One of the isomers forms two stable rotamers separable by HPLC, whereas the other apparently prefers one single rotameric form. The favored conformations of these three forms were elucidated by means of molecular mechanics and dynamics techniques. In parallel, we have prepared the adduct, isolated the three rotamers, and recorded their solution circular dichroism (CD) spectra. For the first time we were thus able to correlate the CD features of individual rotamers of a cis-Pt(GpG) chelate with their structures. We show here that the two forms labeled in Inagaki's paper 1'e and 2e have the same right-handed helicoidal arrangement of the guanine bases but display different CD spectra in which the prominent bands have inverted signs. Thus, base-base interactions cannot be the principal cause of the CD of these compounds.  相似文献   

20.
The bulky, asymmetric analog of the antitumor drug cisplatin, [PtCl(2)(tmen)] (tmen = N,N,N'-trimethylethylenediamine), was used to produce crosslinks with the dinucleotide d(GpG), modeling the most frequent lesions that cisplatin and its analogs cause to DNA. The ligand tmen was chosen because it is expected to constrain the guanine cis to the NMe(2) group in the adduct [Pt(tmen){d(GpG)}](+) to an orientation perpendicular to the coordination plane and to stabilize the other guanine in an oblique orientation, thus maintaining a head-to-head geometry typical of cisplatin-d(GpG) crosslinks within single- and double-stranded DNA. Of the four possible combinations of tmen chirality (R or S symmetry of the coordinated NHMe group) and crosslink direction (5'-G bound cis to the secondary or the tertiary amino group of tmen), two isomers were preponderantly formed, [Pt(R-tmen){d(GpG)}](+) with 5'-G bound cis to NMe(2) and [Pt(S-tmen){d(GpG)}](+) with 5'-G bound cis to NHMe. The former was shown to have a right-handed R2 orientation of guanines similar to that found in duplex DNA, whereas the latter had a left-handed L1 orientation that modeled cisplatin-d(GpG) adducts within single-stranded DNA. The R2 rotamer was found to be in an equilibrium (as observed using EXSY spectroscopy) with a minor fraction (< or =4%) of a Delta-HT rotamer related to R2 by rotation of the 3'-G about the Pt-N7 bond. The major rotamers R2 and L1 were isolated using reverse-phase HPLC, and their NMR and CD signatures were compared to those of the corresponding rotamers of the less hindered adduct [Pt(dmen)(GpG)](+) (dmen = N,N-dimethylethylenediamine). From this and other comparisons with previously reported platinum dinucleotide complexes, and from molecular modeling, it could be concluded that both steric repulsion between guanine and substituents of the cis amino group and N-H...O6 hydrogen bonding are significant effects favoring the oblique orientation of one guanine base typical of the HH rotamers of [Pt(diamine){d(GpG)}](+) and [Pt(diamine)(GpG)](+) complexes.  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