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.
The synthesis of two new synthetic analogues of lecithin, two of phosphatidyl ethanolamine ("cephalin"), and one new phosphatidic acid analogue is described. They comprise one of each of the following types: the "isosteric" diether lecithin and cephalin analogues ROCH(2)CH(OR)- CH(2)CH(2)P(O) (O(-))OCH(2)CH(2)N(+)R'(3) (R = C(18)H(37); R' = H or CH(3)); and the "hydrocarbon" analogues of phosphatidic acid, lecithin, and cephalin, C(17)H(35)CH(2)CH(C(18)H(37))CH(2)P(O)(R) = (R'); [R = R' = OH; R = O(-), R' = OCH(2)CH(2)N(+)(CH(3))(3); and R = O(-), R' = OCH(2)CH(2)N(+)H(3)]. Infrared spectra and other properties of these compounds are described.  相似文献   

2.
A bioorganometallic approach to malaria therapy led to the discovery of ferroquine (FQ, SSR97193). To assess the importance of the electronic properties of the ferrocenyl group, cyclopentadienyltricarbonylrhenium analogues related to FQ, were synthesized. The reaction of [N-(7-chloro-4-quinolinyl)-1,2-ethanodiamine] with the cyrhetrenylaldehyde complexes (η(5)-C(5)H(4)CHO)Re(CO)(3) and [η(5)-1,2-C(5)H(3)(CH(2)OH)(CHO)]Re(CO)(3) produces the corresponding imine derivatives [η(5)-1,2-C(5)H(3)(R)(CHN-CH(2)CH(2)NH-QN)]Re(CO)(3) R=H 3a; R=CH(2)OH 3b; QN=N-(7-Cl-4-quinolinyl). Reduction of 3a and 3b with sodium borohydride in methanol yields quantitatively the amine complexes [η(5)-1,2-C(5)H(3)(R)(CH(2)-NH-CH(2)CH(2)NH-QN)]Re(CO)(3) R=H 4a; R=CH(2)OH 4b. To establish the role of the cyrethrenyl moiety in the antimalarial activity of this series, purely organic parent compounds were also synthesized and tested. Evaluation of antimalarial activity measured in vitro against the CQ-resistant strains (W2) and the CQ-susceptible strain (3D7) of Plasmodium falciparum indicates that these cyrhetrene conjugates are less active compared to their ferrocene and organic analogues. These data suggest an original mode-of-action of FQ and ferrocenyl analogues in relationship with the redox pharmacophore.  相似文献   

3.
In the formation reaction of Ni(2+) with the chiral racemic ligand, (R)(R)bdtp(-)/(S)(S)bdtp(-), bdtp(-) = [SSPOCH)CH(3))CH(CH(3))O](-), cyclo- O,O'-[1,2-dimethylethylene] dithiophosphato ion, the meso-complex Ni[(R)(R)(lambda)bdtp][(S)(S)(delta)-bdtp] is stereoselectively produced. The meso-complex was compared with the enantiopure crystals of (+)(589)Ni[(R)(R)(lambda)bdtp](2) or (-)(589)Ni[(S)(S)(delta)bdtp](2), as well as racemic crystals, rac-(+/-)Ni[bdtp](2), which were prepared from the solution containing the two enantiomers in a 1:1 ratio. Dissociation constants in solutions indicate different stability of the meso and enantiopure complexes depending on the solvent, whereas a more efficient crystal packing, weak H-bonding, and nonbonding interactions contribute to stabilization of the meso-species over the racemic one. Molecular structures show that the outer five-membered ligand ring adopts the half-chair conformation C(2) with either the lambda or the delta chirality and the methyl groups are in equatorial (e) positions. Enantiopure ligands of (+)(589)Ni[(R)(R)(lambda)bdtp](2) and (-)(589)Ni[(S)(S)(delta)bdtp](2) induce chirality into the symmetric SSNiSS chromophore with slightly helical distortion. Thus, their CD spectra exhibit weak negative or positive Cotton effects at 662 nm. CD spectra in L(+)- and D(-)diethyltartrate of the meso-complex and racemic crystal, rac-(+/-)Ni[bdtp](2), exhibit different weak Cotton effects of opposite sign. Complexes dissociate in methanol; rac-(+/-)Ni[bdtp](2) in methanol undergoes a crystallization-induced second-order asymmetric transformation which finally yields crystals of the meso-Ni[(R)(R)(lambda)bdtp][(S)(S)(delta)bdtp] complex.  相似文献   

4.
The reactions of RO(2)* radicals with Fe(H(2)O)(6)(2+) were studied, R[double bond]H; CH(3); CH(2)COOH; CH(2)CN; CH(2)C(CH(3))(2)OH; CH(2)OH; CHCl(2)/CCl(3). All these processes involve the following reactions: Fe(H(2)O)(6)(2+)+RO(2)*<==>(H(2)O)(5)Fe(III)[bond]OOR(2+) K(1) approximately 250 M(-1); (H(2)O)(5)Fe(III)[bond]OOR(2+)+H(3)O(+)/H(2)O-->Fe(H(2)O)(6)(3+)+ROOH+H(2)O/OH(-); (H(2)O)(5)Fe(III)[bond]OOR(2+)+2Fe(H(2)O)(6)(2+)-->3Fe(H(2)O)(6)(3+)+ROH; 2 RO(2)*-->Products; RO(2)*+(H(2)O)(5)Fe(III)[bond]OOR(2+)-->Fe(H(2)O)(6)(2+)+products. The values of k(1) and k(3) [reaction is clearly not an elementary reaction] approach the ligand exchange rate of Fe(H(2)O)(6)(2+), i.e. these reactions follow an inner sphere mechanism and the rate determining step is the ligand exchange step. The rate of reaction is several orders of magnitude faster than that of the Fenton reaction. Surprisingly enough the K(1) values are nearly independent of the redox potential of the radical and are considerably higher than calculated from the relevant redox potentials. These results indicate that the ROO(-) ligands considerably stabilise the Fe(III) complex, this stabilisation is smaller for radicals with electron withdrawing groups which raise the redox potential of the radical but decrease the basicity of the ROO(-) ligands, two effects which seem to nearly cancel each other. Finally, the results clearly indicate that reaction (5) is relatively fast and affects the nature of the final products. The contribution of these reactions to oxidation processes involving 'Fenton-like' processes is discussed.  相似文献   

5.
Deferriferrioxamine B (H3DFB) is a linear trihydroxamic acid siderophore with molecular formula NH2(CH2)5[N(OH)C(O)(CH2)2C(O)NH(CH2)5]2N(OH)C(O)CH3 that forms a kinetically and thermodynamically stable complex with iron(III), ferrioxamine B. Under the conditions of our study (pH = 4.30, 25 degrees C), ferrioxamine B, Fe(HDFB)+, is hexacoordinated and the terminal amine group is protonated. Addition of simple hydroxamic acids, R1C(O)N(OH)R2 (R1 = CH3, R2 = H; R1 = C6H5, R2 = H; R1 = R2 = CH3), to an aqueous solution of ferrioxamine B at pH = 4.30, 25.0 degrees C, I = 2.0, results in the formation of ternary complexes Fe(H2DFB)A+ and Fe(H3DFB)A2+, and tris complexes FeA3, where A- represents the bidendate hydroxamate anion R1C(O)N(O)R2-. The addition of a molar excess of ethylenediaminetetraacetic acid (EDTA) to an aqueous solution of ferrioxamine B at pH 4.30 results in a slow exchange of iron(III) to eventually completely form Fe(EDTA)- and H4DFB+. The addition of a hydroxamic acid, HA, catalyzes the rate of this iron exchange reaction: (formula; see text) A four parallel path mechanism is proposed for reaction (1) in which catalysis occurs via transient formation of the ternary and tris complexes Fe(H2DFB) A+, Fe(H3DFB)A2+, and FeA3. Rate and equilibrium constants for the various reaction paths to products were obtained and the influence of hydroxamic acid structure on catalytic efficiency is discussed. The importance of a low energy pathway for iron dissociation from a siderophore complex in influencing microbial iron bio-availability is discussed. The system represented by reaction (1) is proposed as a possible model for in vivo catalyzed release of iron from its siderophore complex at the cell wall or interior, where EDTA represents the intracellular storage depot or membrane-bound carrier and HA represents a low molecular weight hydroxamate-based metabolite capable of catalyzing interligand iron exchange.  相似文献   

6.
Six new methylenephosphonate analogues of P1P4-bis-(5',5'-adenosyl) tetraphosphate, Ap4A, having P2-P3 carbon bridges CF2, CCl2 and CH2CH2 or P1-P2 and P3-P4 carbon bridges CF2, CCl2 and CH2CH2 in the tetraphosphate chain, were examined as substrates or inhibitors for two specific Ap4A-degrading enzymes: (asymmetrical) Ap4A hydrolase (EC 3.6.1.17) from yellow-lupin seeds and (symmetrical) Ap4A hydrolase (EC 3.6.1.41) from Escherichia coli. All analogues in which the central oxygen atom was replaced by a stable carbon bridge were hydrolysed by the asymmetrical hydrolase (CF2 greater than CCl2 greater than O greater than CHBr greater than CH2 greater than CH2CH2). As expected, these analogues were not hydrolysed by the symmetrical hydrolase, which was also unable to act on analogues having P1-P2 and P3-P4 carbon bridges.  相似文献   

7.
2-Carboxylbenzaldehyde thiosemicarbazone (HL), and its three lanthanide (III) complexes, LnL(3) x 4H(2)O [Ln(III)=La, Sm, Eu], have been synthesized in water. The complexes were characterized by elemental analyses, molar conductivity and IR spectra. The crystal structure of [Sm(2)L(6)(CH(3)OH)(4)] x 7.5CH(3)OH x 0.5H(2)O obtained from methanol solution was determined by X-ray diffraction analysis, crystallized in the triclinic system, space group P-1, Z=1, a=12.217 (2)A, b=14.706 (2)A, c=15.035 (2)A, alpha=111.84(1) degrees , beta=103.47(1) degrees , gamma=104.24(1) degrees , R(1)=0.0290. It has symmetrical (mu-OCO)(2), (mu-O)(2) and disamarium(III) units. The coordination geometry of each Sm(III) ion is a distorted tetradecahedron with nine oxygen atoms. In addition, the DNA-binding properties of the ligand and its complexes have been investigated by absorption, fluorescence, and viscosity measurements. The experimental results indicate that the ligand and the Sm-complex can bind to DNA, but the other two complexes cannot; the binding affinity of the Sm-complex is higher than that of the ligand and the intrinsic binding constant K(b) of the complex is 3.22 x 10(5)M(-1).  相似文献   

8.
The reactions of phosphine Ph(2)P(CH(2))(2)SO(3)Na with Cp(2)M'Cl(2) (M'=Ti, Zr) in aqueous solution give the metallophosphines, Cp(2)Ti(OSO(2)(CH(2))(2)PPh(2))(2) (Cp=cyclopentadienyl) and CpZr(OH)(OSO(2)(CH(2))(2)PPh(2))(2). These react with CODM"Cl(2) (M"=Pd, Pt) (COD=1,5-cyclooctadiene) in dichloromethane to give heterobimetallic complexes Cp(2)Ti(OSO(2)(CH(2))(2)PPh(2))(2)M"Cl(2) and CpZr(OH)(OSO(2)(CH(2))(2) PPh(2))(2)M"Cl(2) respectively. The compounds are characterised by infrared and NMR spectroscopies and elemental analysis. Electrospray mass spectra of the complexes are reported and compared to those of Cp(2)M'Cl(2) in water and dimethylsulfoxide (DMSO). For zirconocene dichloride and its product heterobimetallic complexes, the addition of ethylenediamine tetraacetic acid disodium salt (Na(2)H(2)EDTA) was found to be an effective ionisation enhancement agent for the electrospray mass spectral studies. Cytotoxicity studies for the previously reported Cl(2)Pt(PPh(2)(CH(2))(2)SO(3)H)(2).3.5H(2)O (Wedgwood et al., Inorg. Chim. Acta 290 (1999) 189), and the compounds Cp(2)Ti(OSO(2)(CH(2))(2) PPh(2))(2).1.5H(2)O and Cp(2)Ti(OSO(2)(CH(2))(2)PPh(2))(2)PtCl(2).4H(2)O reported here, have been evaluated by colony formation assay against cisplatin-sensitive and -resistant cell lines L929 and L929/R to highlight potential chemotherapeutic activity. The compound Cl(2)Pt(PPh(2)(CH(2))(2)SO(3)H)(2).3.5H(2)O overcomes cisplatin resistance.  相似文献   

9.
Trans-[RuCl(NO)(dppe)2]2+ species were prepared. The complexes have been characterized by microanalysis, IR and 31P[1H] NMR spectroscopy and cyclic voltammetry. The trans-[RuCl(NO)(dppe)2](ClO4)2 complex shows a reversible one-electron-reduction process at E(1/2) = 0.200 V and another one-electron-reduction irreversible process at -0.620 V, both centered at the NO+ group. The dissociation of the NO group from the trans-[RuCl(NO)(dppe)2]2+ after two one-electron reductions results in the formation of the trans- and cis-[RuCl2(dppe)2] isomers. The product of an electrolyzed solution of the same complex at -0.300 V shows an EPR signal consistent with the presence of the [RuCl(NO(0))(dppe)2]+ complex. Crystal data for trans-[RuCl(NO)(dppe)2]2+*[RuCl4(NO)(H2O)]*1/2[RuCl6]4-*2[H2O] (I) and trans-[RuCl(NO)(dppe)(2)]2+*2[RuCl4(NO)(CH3O)]-*3[CH3OH] (II) are as follow: (I) Space group P-1, a=10.4040(3) A, b=12.3470(4) A, c=23.5620(8) A, alpha=95.885(2) degrees, beta=99.608(2) degrees, gamma=104.378(2) degrees, R=0.0521; (II) space group P-1, a=10.9769(2) A, b=13.2753(3) A, c=24.0287(4) A, alpha=99.743(1) degrees, beta=95.847(1) degrees, gamma=97.549(1) degrees; R=0.0496. The fac-[RuCl3(NO)(dppe)] (III) complex has been also prepared; its crystal data are: space group P2(1)/n (No. 14), a=11.841(2) A, b=13.775(2) A, c=16.295(4) A, beta=92.81(2) degrees; R1=0.0395.  相似文献   

10.
Arsenic exposure is associated with hypertension, diabetes, and cancer. Some mammals methylate arsenic. Saccharomyces cerevisiae hexose permeases catalyze As(OH)(3) uptake. Here, we report that mammalian glucose transporter GLUT1 catalyzes As(OH)(3) and CH(3)As(OH)(2) uptake in yeast or in Xenopus laevis oocytes. Expression of GLUT1 in a yeast lacking other glucose transporters allows for growth on glucose. Yeast expressing yeast HXT1 or rat GLUT1 transport As(OH)(3) and CH(3)As(OH)(2). The K(m) of GLUT1 is to 1.2mM for CH(3)As(OH)(2), compared to a K(m) of 3mM for glucose. Inhibition between glucose and CH(3)As(OH)(2) is noncompetitive, suggesting differences between the translocation pathways of hexoses and arsenicals. Both human and rat GLUT1 catalyze uptake of both As(OH)(3) and CH(3)As(OH)(2) in oocytes. Thus GLUT1 may be a major pathway uptake of both inorganic and methylated arsenicals in erythrocytes or the epithelial cells of the blood-brain barrier, contributing to arsenic-related cardiovascular problems and neurotoxicity.  相似文献   

11.
The reaction of [VO(CH3COO)2(phen)] (phen = 1,10-phenanthroline) with the sulfhydryl-containing pseudopeptides (scp), N-(2-mercaptopropionyl)glycine (H3mpg), N-(2-mercaptopropionyl)cysteine (H4m2pc), N-(3-mercaptopropionyl)cysteine (H4m3pc) and the dipeptides glycylglycine (H2glygly) and glycyl-L-alanine (H2glyala), in the presence of triethylamine, results in the formation of the compounds Et3NH[VO(mpg)(phen)] (1), (Et3NH)2[VO(m2pc)] (4), [(Et3NH)2[VO(m3pc) (5), [VO(glygly)(phen)] x 2CH3OH (2 x 2CH3OH) and [VO(glyala)(phen)] x CH3OH (3 x CH3OH). Evidence for the molecular connectivity in 2 x CH3OH was established by X-ray crystallography, showing the vanadium(IV) atom ligated to a tridentate glygly2- ligand at the N(amine), N(peptide) and O(carboxylato) atoms. Combination of the correlation plot of the EPR parameters gz versus Az, together with the additivity relationship supported the prediction of the equatorial donor atom sets of the V(IV)O2+ center at various pH values for the V(IV)O2+-glutathione system considered in this study. Model NMR studies (interaction of vanadium(V) with the scp H3mpg) showed that there is a possibility of vanadium(V) ligation to glutathione.  相似文献   

12.
T4 RNA ligase catalyzes the synthesis of ATP beta,gamma-bisphosphonate analogues, using the following substrates with the relative velocity rates indicated between brackets: methylenebisphosphonate (pCH(2)p) (100), clodronate (pCCl(2)p) (52), and etidronate (pC(OH)(CH(3))p) (4). The presence of pyrophosphatase about doubled the rate of these syntheses. Pamidronate (pC(OH)(CH(2)-CH(2)-NH(2))p), and alendronate (pC(OH)(CH(2)-CH(2)-CH(2)-NH(2))p) were not substrates of the reaction. Clodronate displaced the AMP moiety of the complex E-AMP in a concentration dependent manner. The K(m) values and the rate of synthesis (k(cat)) determined for the bisphosphonates as substrates of the reaction were, respectively: methylenebisphosphonate, 0.26+/-0.05 mM (0.28+/-0.05 s(-1)); clodronate, 0.54+/-0.14 mM (0.29+/-0.05 s(-1)); and etidronate, 4.3+/-0.5 mM (0.028+/-0.013 s(-1)). In the presence of GTP, and ATP or AppCCl(2)p the relative rate of synthesis of adenosine 5',5'-P(1),P(4)-tetraphosphoguanosine (Ap(4)G) was around 100% and 33%, respectively; the methylenebisphosphonate derivative of ATP (AppCH(2)p) was a very poor substrate for the synthesis of Ap(4)G. To our knowledge this report describes, for the first time, the synthesis of ATP beta,gamma-bisphosphonate analogues by an enzyme different to the classically considered aminoacyl-tRNA synthetases.  相似文献   

13.
Phosphotriester condensation (RO)(R'O)PO-2 (PDE) + R"OH (RO)(R'O)(R"O)PO (PTE) in the presence of arylsulfonyl chloride (ArSO2Cl) as well as arylsulfonyl azolides proceeds in two steps as revealed by 31P NMR spectroscopy. Pyrophosphotetraester (PPTE) accumulates in over 80% yield in the first step and converts to PTE in the second one. Nucleophilic catalysts of pyridine type (Nu1) are necessary in the first step. The second step is catalyzed by Nu1 as well as by catalysts of the tetrazole type (Nu2H). Base catalysis operates in the latter case. With Nu1 catalysts (pyridine, 4-N,N-dimethylaminopyridine, N-methylimidazole) the general scheme may be presented as follows: ArSO2Cl + Nu1 in equilibrium ArSO2Nu+1 + Cl-; ArSO2Nu+1 + PDE----(RO)(R'O)P(O)OSO2Ar (I); I + Nu+1----(RO)(R'O)P(O)Nu+1 (II); II + PDE in equilibrium [(RO)(R'O)PO]20; II + R"OH----(RO)(R'O)(R"O)PO. Catalysts of Nu2H type don't accelerate PPTE formation. In the second step they participate most probably in the process PPTE + Nu2H in equilibrium (RO)(R'O)P(O)Nu2 (III) + PDE; III + R"OH----(RO)(R'O)(R"O)PO + H+. The latter step is subjected to strong base catalysis.  相似文献   

14.
Previous resistive load detection (RLD) studies have ignored the nose, the usual route of breathing. Weber's law predicts the delta R50 (the added load detectable on 50% of presentations) to be a fixed percent of the background resistance (R0) and thus the delta R50/R0 ratio (the Weber fraction) is constant. We have noted the nose to be sensitive to added load, we wondered if the nose might play a role in RLD. To determine whether this was true and to characterize the effects of changes in R0 in the range of normal nasal resistance (RN), we determined R0 and delta R50 using standard techniques under the following conditions: nose vs. decongested nose, nose vs. nose with added external R0 (3.0 and 8.0 cmH2O X l-1 X s), nose vs. anesthetized nose, nose vs. mouth, and mouth vs. mouth with added load (3 cmH2O X l-1 X s). We found that decongestant decreased RN [4.3 +/- 0.6 (SE) to 3.1 +/- 0.5 cmH2O X l-1 X s, P less than 0.05] and delta R50 (1.7 +/- 0.5 to 1.1 +/- 0.3 cmH2O X l-1 X s, P less than 0.05). When an external load of 3 cmH2O X l-1 X s was added to the nose, delta R50 did not change significantly (1.4 +/- 0.2 to 1.1 +/- 0.2 cmH2O X l-1 X s), but the Weber fraction decreased (0.28 +/- 0.05 to 0.15 +/- 0.03, P less than 0.02).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
We prepared platinum(IV) complexes containing dipeptide and diimine or diamine, the [PtCl(dipeptide-N,N,O)(diimine or diamine)]Cl complex, where -N,N,O means dipeptide coordinated as a tridentate chelate, dipeptide=glycylglycine (NH(2)CH(2)CON(-)CH(2)COO(-), digly, where two protons of dipeptide are detached when the dipeptide coordinates to metal ion as a tridentate chelate), glycyl-L-alanine (NH(2)CH(2)CON(-)CHCH(3)COO(-), gly-L-ala), L-alanylglycine (NH(2)CH CH(3)CON(-)CH(2)COO(-), L-alagly), or L-alanyl-L-alanine (NH(2)CHCH(3)CON(-)CHCH(3)COO(-), dil-ala), and diimine or diamine=bipyridine (bpy), ethylenediamine (en), N-methylethylenediamine (N-Me-en), or N,N'-dimethylethylenediamine (N,N'-diMe-en). In the complexes containing gly-L-ala or dil-ala, two separate peaks of the (195)Pt NMR spectra of the [PtCl(dipeptide-N,N,O)(diimine or diamine)]Cl complexes appeared in, but in the complexes containing digly or L-alagly, one peak which contained two overlapped signals appeared. One of the two complexes containing gly-L-ala and bpy, [PtCl(gly-L-ala-N,N,O)(bpy)]NO(3), crystallized and was analyzed. This complex has the monoclinic space group P2(1)2(1)2(1) with unit cell dimensions of a=9.7906(3)A, b=11.1847(2)A, c=16.6796(2)A, Z=4. The crystal data revealed that this [PtCl(gly-L-ala-N,N,O)(bpy)]NO(3) complex has the near- (Cl, CH(3)) configuration of two possible isomers. Based on elemental analysis, the other complex must have the near- (Cl, CH(3))-[PtCl(gly-L-ala-N,N,O)(bpy)]NO(3) configuration. The (195)Pt NMR chemical shifts of the near- (Cl, CH(3))-[PtCl(gly-L-ala-N,N,O)(bpy)]NO(3) complex and the far- (Cl, CH(3))-[PtCl(gly-L-ala-N,N,O)(bpy)]NO(3) complex are 0 ppm and -19 ppm, respectively (0 ppm for the Na(2)[PtCl(6)] signal). The additive property of the (195)Pt NMR chemical shift is discussed. The (195)Pt NMR chemical shifts of [PtCl(dipeptide-N,N,O)(bpy)]Cl appeared at a higher field when the H attached to the dipeptide carbon atom was replaced with a methyl group. On the other hand, the (195)Pt NMR chemicals shifts of [PtCl(dipeptide-N,N,O)(diamine)] appeared at a lower field when the H attached to the diamine nitrogen atom was replaced with a methyl group, in the order of [PtCl(digly-N,N,O)(en)]Cl, [PtCl(digly-N,N,O)(N-Me-en)]Cl, and [PtCl(digly-N,N,O)(N,N'-diMe-en)]Cl.  相似文献   

16.
The acidity constants of twofold protonated, antivirally active, acyclic nucleoside phosphonates (ANPs), H(2) (PE)(±) , where PE(2-) =9-[2-(phosphonomethoxy)ethyl]adenine (PMEA(2-) ), 2-amino-9-[2-(phosphonomethoxy)ethyl]purine (PME2AP(2-) ), 2,6-diamino-9-[2-(phosphonomethoxy)ethyl]purine (PMEDAP(2-) ), or 2-amino-6-(dimethylamino)-9-[2-(phosphonomethoxy)ethyl]purine (PME(2A6DMAP)(2-) ), as well as the stability constants of the corresponding ternary Cu(Arm)(H;PE)(+) and Cu(Arm)(PE) complexes, where Arm=2,2'-bipyridine (bpy) or 1,10-phenanthroline (phen), are compared. The constants for the systems containing PE(2-) =PMEDAP(2-) and PME(2A6DMAP)(2-) have been determined now by potentiometric pH titrations in aqueous solution at I=0.1M (NaNO(3) ) and 25°; the corresponding results for the other ANPs were taken from our earlier work. The basicity of the terminal phosphonate group is very similar for all the ANP(2-) species, whereas the addition of a second amino substituent at the pyrimidine ring of the purine moiety significantly increases the basicity of the N(1) site. Detailed stability-constant comparisons reveal that, in the monoprotonated ternary Cu(Arm)(H;PE)(+) complexes, the proton is at the phosphonate group, that the ether O-atom of the ?CH(2) ?O?CH(2) ?P(O)$\rm{{_{2}^{-}}}$(OH) residue participates, next to the P(O)$\rm{{_{2}^{-}}}$(OH) group, to some extent in Cu(Arm)(2+) coordination, and that π?π stacking between the aromatic rings of Cu(Arm)(2+) and the purine moiety is rather important, especially for the H?PMEDAP(-) and H?PME(2A6DMAP)(-) ligands. There are indications that ternary Cu(Arm)(2+) -bridged stacks as well as unbridged (binary) stacks are formed. The ternary Cu(Arm)(PE) complexes are considerably more stable than the corresponding Cu(Arm)(R?PO(3) ) species, where R?PO$\rm{{_{3}^{2-}}}$ represents a phosph(on)ate ligand with a group R that is unable to participate in any kind of intramolecular interaction within the complexes. The observed stability enhancements are mainly attributed to intramolecular-stack formation in the Cu(Arm)(PE) complexes and also, to a smaller extent, to the formation of five-membered chelates involving the ether O-atom present in the ?CH(2) ?O?CH(2) ?PO$\rm{{_{3}^{2-}}}$ residue of the PE(2-) species. The quantitative analysis of the intramolecular equilibria involving three structurally different Cu(Arm)(PE) isomers shows that, e.g., ca. 1.5% of the Cu(phen)(PMEDAP) system exist with Cu(phen)(2+) solely coordinated to the phosphonate group, 4.5% as a five-membered chelate involving the ether O-atom of the ?CH(2) ?O?CH(2) ?PO$\rm{{_{3}^{2-}}}$ residue, and 94% with an intramolecular π?π stack between the purine moiety of PMEDAP(2-) and the aromatic rings of phen. Comparison of the various formation degrees of the species formed reveals that, in the Cu(phen)(PE) complexes, intramolecular-stack formation is more pronounced than in the Cu(bpy)(PE) species. Within a given Cu(Arm)(2+) series the stacking intensity increases in the order PME2AP(2-) 相似文献   

17.
The structure of a crystal complex of recombinant human immunodeficiency virus type 1 (HIV-1) protease with a peptide-mimetic inhibitor containing a dihydroxyethylene isostere insert replacing the scissile bond has been determined. The inhibitor is Noa-His-Hch psi [CH(OH)CH(OH)]Vam-Ile-Amp (U-75875), and its Ki for inhibition of the HIV-1 protease is < 1.0 nM (Noa = 1-naphthoxyacetyl, Hch = a hydroxy-modified form of cyclohexylalanine, Vam = a hydroxy-modified form of valine, Amp = 2-pyridylmethylamine). The structure of the complex has been refined to a crystallographic R factor of 0.169 at 2.0 A resolution by using restrained least-squares procedures. Root mean square deviations from ideality are 0.02 A and 2.4 degrees, for bond lengths and angles, respectively. The bound inhibitor diastereomer has the R configurations at both of the hydroxyl chiral carbon atoms. One of the diol hydroxyl groups is positioned such that it forms hydrogen bonds with both the active site aspartates, whereas the other interacts with only one of them. Comparison of this X-ray structure with a model-built structure of the inhibitor, published earlier, reveals similar positioning of the backbone atoms and of the side-chain atoms in the P2-P2' region, where the interaction with the protein is strongest. However, the X-ray structure and the model differ considerably in the location of the P3 and P3' end groups, and also in the positioning of the second of the two central hydroxyl groups. Reconstruction of the central portion of the model revealed the source of the hydroxyl discrepancy, which, when corrected, provided a P1-P1' geometry very close to that seen in the X-ray structure.  相似文献   

18.
Four imidazolate-bridged binuclear copper(II)-copper(II) and copper(II)-zinc(II) complexes viz., [(Bipy)(2)Cu-Im-Cu(Bipy)(2)](ClO(4))(3).CH(3)OH, [(Phen)(2)Cu-Im-Cu(Phen)(2)](BF(4))(3).2CH(3)OH, [(Bipy)(2)Cu-Im-Zn(Bipy)(2)](BF(4))(3), and [(Phen)(2)Cu-Im-Zn(Phen)(2)](BF(4))(3), (Bipy=2,2'-Bipyridyl, Phen=1-10-Phenanthroline and Im=imidazolate ion) were synthesized as a possible models for superoxide dismutase (SOD). Complex [(Bipy)(2)Cu-Im-Cu(Bipy)(2)](ClO(4))(3).CH(3)OH has been structurally characterized. This complex crystallizes in the triclinic space group P1, with the unit parameters a=8.88(5) A, b=13.79(17) A, c=20.18(18) A, alpha=76.424(8)(o), beta=85.888(6)(o), gamma=82.213(7). The metal-nitrogen bond length from 1.972-2.273 A and the distance Cu-Cu is 5.92 A. The five-coordinate geometry about the copper(II) ion is square pyramidal. Magnetic moment and electron paramagnetic resonance (e.p.r.) spectral measurements of the homobinuclear complexes have shown an antiferromagnetic exchange interaction. From the e.p.r. and UV-Vis spectral measurement studies, these complexes have been found to be stable (pH 8.5-10.5 for 1, 10.5 for 2,3 and 8.5 for 4). These complexes catalyse the dismutation of superoxide radical (O(2)(-)) at biological pH. All the observations indicate that these complexes act as good possible models for superoxide dismutase.  相似文献   

19.
Kumar NN  Swamy KC 《Chirality》2008,20(6):781-789
Diastereoselective synthesis and characterization of chiral unsymmetrical tris-spirocyclic cyclotriphosphazenes based on chiral 1,1'-bi-2-naphthol (BINOL) are reported. Specifically, the chiral compounds (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)Cl(2) [(-)-4] and (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](OCH(2)CH(2)NMe)(2) [(-)-5] are prepared by starting with the chiral mono-spiro compound (-)N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)]Cl(4) [(-)-3]. Synthesis of four other chiral spirocyclics, N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](OCH(2)CH(2) NMe)(O-2,2'C(6)H(4)-C(6)H(4)O)[(-)-6 and (+)-6], N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](NMe(2))(4) [(-)-7], N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)(NMeCH(2)CH(2)OH)(2) [(-)-8 and (+)-8], and N(3)P(3)[1,1'-O(2)(C(10)H(6))(2)](O-2,2'C(6)H(4)-C(6)H(4)O)[NHCH(2)CH(2)CH(2)Si(OEt)(3)](2) (9) is also reported herein. Compounds 4-6 are obtained in the solid state diastereoselectively and their X-ray structures have been determined and discussed. The diastereoselectivity is also shown by structural characterization of two distinct isomers in the case of 6 [(-)-6 and (+)-6, respectively] by starting with precursor of 3 having (R) or (S)-BINOL residue. The (1)H NMR spectra of 7 and 8 exhibit doublets with virtual coupling for the methyl protons, consistent with the chiral nature of the binaphthoxy residue. The potential of 9, which hydrolyzes readily in CDCl(3) solution, as a useful precursor for chiral polymer applications is highlighted.  相似文献   

20.
Two new complexes of Ru(III) with purine base derivatives, [mer-RuCl(3)(acv)(DMSO-S)(C(2)H(5)OH)].C(2)H(5)OH (1) (acv=acyclovir, DMSO=dimethyl sulfoxide) and [trans-RuCl(4)(guaH)(DMSO-S)].2H(2)O (2) (guaH=protonated molecule of guanine), were prepared from the same Ru(III) precursor, [trans-RuCl(4)(DMSO-S)(2)](-), by substitution of one DMSO-S. Coordination of acv induced also replacement of one chloride by an ethanol molecule. This reactivity difference was explained by striking contrasts in the hydrogen bonding schemes of the two complexes, evidenced in their X-ray crystal structures. In 1 the guanine derivative acyclovir is coordinated to ruthenium through the N(7) atom, while in 2 the protonated guanine molecule is bound through the N(9) atom. Both complexes were also characterized by various physico-chemical methods in the solid state and in the solution. In vitro, the biological activity of 2 and of the previously described complexes [mer-RuCl(3)(acv)(DMSO-S)(CH(3)OH)].0.5CH(3)OH (3) and [mer-RuCl(3)(acv)(DMSO-S)(H(2)O)].H(2)O (4) on tumour cells appear to be very similar to that of NAMI-A (NAMI-A=[ImH][trans-RuCl(4)(DMSO-S)Im]). All compounds are only weakly active on tumour cell proliferation but show an interesting proadhesive effect that suggest possible activity on tumour malignancy.  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