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 parallel lability trend for the anation reactions of RRh(DH)2H2O (RCH3, CH3CH2, CF3CH2, ClCH2) and RCo(DH)2H2O complexes suggests a dissociative activation process for the reactions of the organorhodoximes. The high lability of these complexes, arising from the stabilization of the transition state, is not entirely due to the trans-effect of the R group, but, at least partially, to the labilizing effect of the equatorial macrocycle.  相似文献   

2.
Second-order rate constants (k2) are reported for the reduction of 9-R-10-methylacridinium cations (5:R = H, CH3, CH3CH2, C6H5CH2, (CH3)2CH, C6H5, 4-(CH3)2NC6H4) by 1-benzyl-1,4-dihydronicotinamide (2:R = C6H5CH2) in 20% CH3CN-80% H2O at 25°C. All 5:R ≠ H are reduced in the range 20- to 140-fold more slowly than 5:R = H. However, there is no simple relationship between k2 and the nature of R, nor between k2 and the second-order rate constant for hydroxide ion attack at C-9 of these cations in pseudobase formation. Rates of reduction of 5 by 1-benzyl-4,4-dideuterio-1,4-dihydronicotinamide allow the calculation of the following kinetic isotope effects in this solvent medium: 5:R, kHkD:H, 1.56; C6H5CH2, 2.7; C6H5, 5.4. Substituent effects upon k2 were evaluated for the reduction of 5 by 1-(X-benzyl)-1,4-dihydronicotinamides, and lead to the following Hammett ? parameters: 5:R, ?: H, ?0.68; C6H5CH2, ?0.92; C6H5, ?0.96. The latter two values require essentially complete unit positive charge generation on the nicotinamide moiety in the rate-determining transition state. It is shown that these Hammett ? values and the above isotope effects can only be rationalized by a two-step e? + H? mechanism for hydride transfer from 2 to 5 in this solvent system. This result contrasts with our earlier conclusion of direct, one-step hydride transfer in the reduction of isoquinolinium cations by 2, but is consistent with our observation that acridinium cations are reduced 37500-fold faster by 2 than predicted on the basis of the relative rates of nucleophilic attack (hydroxide ion) on acridinium and isoquinolinium cations. It is suggested that the availability of both Hammett ? values and primary kinetic isotope effects will generally allow the establishment of the mechanism of hydride transfer in these systems. Application of these ideas to literature data suggests that 5:R = H is reduced by direct hydride transfer in acetonitrile solution, in contrast to the above result in predominantly aqueous solution. The ready formation of acridanyl radicals by electron transfer to acridinium cations is demonstrated by the formation of Wurster's Blue radical cation upon mixing solutions of acridinium cations with N,N,N′,N′-tetramethyl-p-phenylenediamine.  相似文献   

3.
《Inorganica chimica acta》1987,134(1):139-142
A new series of heterocyclic telluronium salts (C4H8OTeRX: R=CH3CH2, CH2CHCH2, CH3, C4H9, X=I; R=CH2CHCH2, X=Br; R=CH3, X=ClO4; R=CH3, CH3CH2, C6H5, X=BPh4) have been prepared. Conductivity measurements in dimethylsulphoxide (DMSO) and N,N-dimethytl- formamide (DMF) have shown that there is an ion pair interaction between the anion and the tellurium cation.1H NMR studies showed that there is no reaction between the solute and the solvent. All compounds are stable in solution in DMSO. Infra-red spectra are reported and discussed.  相似文献   

4.
Gallium(III) tris-dialkyldithiophosphates, Ga[S2P(OR)2]3 (R = C2H5, n-C3H7, i-C3H7, n-C4H9 and i-C4H9) and gallium(III) tris-alkylenedithiophosphates, Ga(S2POGO)3 [G = -CH2C(C2H5)2CH2-, -C(CH3)2C(CH3)2 and -C(CH3)2CH2CH(CH3)] have been synthesized for the first time by the reactions of gallium(III) chloride with the alkali metal salt of the corresponding ligand in anhydrous benzene in 1:3 molar ratio respectively.These compounds are crystalline solids or viscous liquids and are soluble in common organic solvents, in which they show monomeric behaviour. Based on elemental analyses, molecular weight determinations, IR and NMR (1H and 31P) spectral data, chelate octahedral structures have been proposed for these derivatives.  相似文献   

5.
N-substituted ethylcarbamates form with thorium nitrate the complexes Th(NO3)4·3RHNC(O)OC2H5 (where R = CH3, C2H5, C6H5(CH3)CH) and with lanthanum nitrate the complexes La(NO3)3· 2RR′NC(O)OC2H5·3H2O (where R = CH3, C2H5, C6H5(CH3)CH; R′ = H and R = CH3, C6H5; R′ = C2H5 or R = R′ = CH3). In addition the anhydrous La(NO3)3·3(C2H5)2NC(O)OC2H5 has been isolated. From the IR spectra it is deduced that the carbamates coordinate the metal through the carbonyl oxygen atom and that the nitrato groups act as chelated ligands. 1H nmr spectral data of the complexes are reported and discussed.  相似文献   

6.
Resolution of rac‐3,3,3‐trifluorolactic acid by diastereomeric salt formation was reinvestigated. The use of (S)‐1‐phenylethylamine gives coprecipitation of two diastereomeric phases, 1 (S)‐[NH3CH(CH3)Ph](S)‐[CF3CH(OH)COO] and 2 (S)‐[NH3CH(CH3)Ph](R)‐[CF3CH(OH)COO]·H2O. Pure phase 1 may be obtained using molecular sieves as desiccants. Resolution by (S,S)‐2‐amino‐1‐phenylpropan‐1,3‐diol gives monoclinic (S,S)‐[NH3CH(CH2OH)CHOHPh] (R)‐[CF3CH(OH)‐COO] 3 with minor (S)‐3,3,3‐trifluorolactate contamination, which is precluded in the recrystallized orthorhombic form 4 . A new resolution using inexpensive phenylglycinol gives pure phase 5 (S)‐[NH3CH(CH2OH)Ph] (S)‐[CF3CH(OH)COO] in 76% yield, 94% ee in a single step, in preference to its (S)‐(R) diastereomer 6 . Overall efficient resolution for both enantiomers of the trifluorolactic acid (each ca. 70% yield, 99% ee) may be achieved by various two‐step “tandem” crystallizations, involving direct addition of either water or a second base to the filtrate from the initial reaction.  相似文献   

7.
The compounds W(CO)5P(C6H4-4-CH2CH2(CF2)7CF3)3 (1) and W(CO)5P(CH2CH2(CF2)5CF3)3 (2) were synthesized in order to probe the electronic and physical effects of ligation by perfluorocarbon substituted tertiary phosphine ligands in a W(CO)5L complex. The π-accepting ability of the fluorous phosphines was found to rank with non-fluorous comparators as P(CH2CH2(CF2)5CF3)3 > P(C6H4-4-CH2CH2(CF2)7CF3)3 > PPh3 > P(p-tolyl)3 > P(n-octyl)3. The X-ray crystal structure of W(CO)5P(C6H4-4-CH2CH2(CF2)7CF3)3 shows strong intermolecular association of fluorous components but confirms that the para fluorocarbon subtituents have an insignificant effect on the tungsten coordination environment. Partition coefficients (toluene/perfluoromethylcyclohexane) were measured for compounds 1 and 2.  相似文献   

8.
The 1,3-oxazine complexes cis- and trans-[PtCl2{ C(R)OCH2CH2C}H22] (cis: R=CH3 (1a), CH2CH3 (2a), (CH3)3C (3a), C6H5 (4a); trans:R =CH3 (1b), C6H5 (4b)) were obtained in 51-71% yield by reaction in THF at 0 °C of the corresponding nitrile complexes cis- and trans-[PtCl2(NCR)2] with 2 equiv. of OCH2CH2CH2Cl, generated by deprotonation of 3-chloro-1-propanol with n-BuLi. The cationic nitrile complexes trans-[Pt(CF3)(NCR)(PPh3)2]BF4 (R=CH3, C6H5) react with 1 equiv, of OCH2CH2CH2Cl to give a mixture of products, including the corresponding oxazine derivatives trans-[Pt(CF3){ CH2}(PPh3)2]BF4 (5 and 6), the chloro complex trans- [Pt(CF3)Cl(PPh3)2] and free oxazine H2. For short reaction times (c. 5–15 min) the oxazine complexes 5 and 6 could be isolated in modest yield (37–49%) from the reaction mixtures and they could be separated from the corresponding chloro complex (yield 40%) by taking advantage of the higher solubility of the latter derivative in benzene. For longer reaction times (> 2 h), trans-[Pt(CF3)Cl(PPh3)2] was the only isolated product. Complex 6 was crystallographically characterized and it was found to contain also crystals of trans- [PtCl{ H2}(PPh3)2]BF4, which prevented a more detailed analysis of the bond lengths and angles within the metal coordination sphere. The 1,3-oxazine ring, which shows an overall planar arrangement, is characterized by high thermal values of the carbon atoms of the methylene groups indicative of disordering in this part of the molecule in agreement with fast dynamic ring processes suggested on the basis of 1H NMR spectra. It crystallizes in the trigonal space group P , with a=22.590(4), b=15.970(3) Å, γ=120°, V=7058(1) Å3 and Z=6. The structure was refined to R=0.059 for 3903 unique observed (I3σ(I)) reflections. A mechanism is proposed for the conversion of nitrile ligands to oxazines in Pt(II) complexes.  相似文献   

9.
[1+1] macrocyclic and [1+2] macroacyclic compartmental ligands (H2L), containing one N2O2, N3O2, N2O3, N4O2 or O2N2O2 Schiff base site and one O2On (n=3, 4) crown-ether like site, have been prepared by self-condensation of the appropriate formyl- and amine precursors. The template procedure in the presence of sodium ion afforded Na2(L) or Na(HL) · nH2O. When reacted with the appropriate transition metal acetate hydrate, H2L form M(L) · nH2O, M(HL)(CH3COO) · nH2O, M(H2L)(X)2 · nH2O (M=Cu2+, Co2+, Ni2+; X=CH3COO, Cl) or Mn(L)(CH3COO) · nH2O according to the experimental conditions used. The same complexes have been prepared by condensation of the appropriate precursors in the presence of the desired metal ion. The Schiff bases H2L have been reduced by NaBH4 to the related polyamine derivatives H2R, which form, when reacted with the appropriate metal ions, M(H2R)(X)2 (M= Co2+, Ni2+; X=CH3COO, Cl), Cu(R) · nH2O and Mn(R)(CH3COO) · nH2O. The prepared ligands and related complexes have been characterized by IR, NMR and mass spectrometry. The [1+1] cyclic nature of the macrocyclic polyamine systems and the site occupancy of sodium ion have been ascertained, at least for the sodium (I) complex with the macrocyclic ligand containing one N3O2 Schiff base and one O2O3 crown-ether like coordination chamber, by an X-ray structural determination. In this complex the asymmetric unit consists of one cyclic molecule of the ligand coordinated to a sodium ion by the five oxygen atoms of the ligand. The coordination geometry of the sodium ion can be described as a pentagonal pyramid with the metal ion occupying the vertex. In the mononuclear complexes with H2L or H2R the transition metal ion invariantly occupies the Schiff base site; the sodium ion, on the contrary, prefers the crown-ether like site. Accordingly, the heterodinuclear complexes [MNa(L)(CH3COO)x] (M=Cu2+, Co2+, Ni2, x=1; M=Mn3+, x=2) have been synthesised by reacting the appropriate formyl and amine precursors in the presence of M(CH3COO)n · nH2O and NaOH in a 1:1:1:2 molar ratio. The reaction of the mononuclear transition metal complexes with Na(CH3COO) · nH2O gives rise to the same heterodinuclear complexes. Similarly [MNa(R)(CH3COO)x] have been prepared by reaction of the appropriate polyamine ligand H2R with the desired metal acetate hydrate and NaOH in 1:1:2 molar ratio.  相似文献   

10.
《Inorganica chimica acta》1986,115(2):147-151
In the presence of Fe3+, template condensation of the fluorinated keto-alcohol CH3C(O)CH2C- (CF3)2OH with the triamine CH3C(CH2NH2)3 leads to two products: a fully condensed, imino-alkoxy, iron(III) complex, Fe{CH3C[CH2NC(CH3)CH2C(CF3)2O]3}, and a partially condensed iron(III) complex, O{FeCH3C[CH2NC(CH3)CH2C(CF3)2O]2(CH2NH2)}2, in which two six-coordinate iron(III) centers are linked by an oxide ion. A complete crystal and molecular structure determination of the latter has been made.Crystals are monoclinic, space group C2/c, a= 13.886(4); b=23.206(5); c=15.241(4) Å; β= 106.55(2)°; V=4708 Å3; Z=4. Least-squares refinement on F of 322 variables using 2627 observations converged at a conventional agreement factor of 3.8%. The Fe to bridging oxide distance is 1.811(1) Å, the FeFe distance 3.468 Å, and the FeOFe angle 146.6(2)°. A comparison is made between this structure and those of natural hemerythrin systems.  相似文献   

11.
The green sulfur photosynthetic bacterium Chlorobaculum (Cba.) tepidum was grown in liquid cultures containing perfluoro-1-decanol, 1H,1H,2H,2H-heptadecafluoro-1-decanol [CF3(CF2)7(CH2)2OH] or 1H,1H-nonadecafluoro-1-decanol [CF3(CF2)8CH2OH], to introduce rigid and fluorophilic chains into the esterifying moiety of light-harvesting bacteriochlorophyll (BChl) c. Exogenous 1H,1H,2H,2H-heptadecafluoro-1-decanol was successfully attached to the 172-carboxy group of bacteriochlorophyllide (BChlide) c in vivo: the relative ratio of the unnatural BChl c esterified with this perfluoroalcohol over the total BChl c was 10.3%. Heat treatment of the liquid medium containing 1H,1H,2H,2H-heptadecafluoro-1-decanol with β-cyclodextrin before inoculation increased the relative ratio of the BChl c derivative esterified with this alcohol in the total BChl c in Cba. tepidum. In a while, 1H,1H-nonadecafluoro-1-decanol was not attached to BChlide c in Cba. tepidum, which was grown by its supplementation. These results suggest that the rigidity close to the hydroxy group of the esterifying alcohol is not suitable for the recognition by the BChl c synthase called BchK in Cba. tepidum. The unnatural BChl c esterified with 1H,1H,2H,2H-heptadecafluoro-1-decanol participated in BChl c self-aggregates in chlorosomes.  相似文献   

12.
The mass spectra six silver(I) carboxylates, AgO2CR, (R = Me, Et, Prn, Ph, CF3 and (CF2)2CF3) show these compounds to be dimeric in the vapour phase, the base peak being the ion Ag2(O2CR)+ in each case. Two fragmentation pathways are observed. The alkyl carboxylates initially lose mainly RCO2· from the radical ion Ag2O2CR)2+, whereas the benzoate and the perfluorocarboxylates also easily lose carbon dioxide from the radical ion leading to the formation of abundant organosilver ions. The low frequency (500−40 cm−1) infrared spectra of these silver(I) carboxylates are compared with the spectra of the copper(I) analogues and bands selected which may be assigned to predominantly skeletal modes.  相似文献   

13.
The sec, rac-CH3Co(H2O)L2+ (L=5,7,7,12,14,14-hexamethyl-1,4,8,11-tetraazacyclotetradeca-4,11-diene) was prepared successfully via meso-CH3Co(H2O)L2+ in aqueous solution. The isomerizations from meso-RCo(H2O)L2+ (R=CH3, C2H5 and C3H7) and sec, rac-CH3Co(H2O)L2+ to pri, rac-RCo(H2O)L2+ were both base catalyzed in aqueous solution. The kinetic results showed the reaction to be first order in both organocobalt complex and hydroxide ion with the reactivity order for the alkyl group being C3H7 ∼ C2H5 ? CH3. However, the conversion from the most steric hindered isomer form of sec, rac- was slow. The ratio of the isomerization rate constants between meso-CH3Co(H2O)L2+ and sec, rac-CH3Co(H2O)L2+ to pri, rac-CH3Co(H2O)L2+ is almost a factor of 100. The thermodynamic activation parameters for these isomerization reactions were investigated.  相似文献   

14.
《Inorganica chimica acta》1988,153(4):219-225
The preparations are reported of [Rh(RCO2)2L]2 [where R = CH3, C2H5, and CH3OCH2; L = 6-chloro-2-methoxy-9-[2(NR′2)ethyl]aminoacridine (R′ = H, CH3)]. X-ray structural studies have been carried out on two of the compounds [ R = C2H5, R′ = H, (1); R = CH3, R′ = CH3, (2)]. Compound 1 is monoclinic, space group C2/c, with a = 20.864(11), b = 15.736(4), c = 14.402(4) Å, β = 93.14(4)°, V = 4721 Å3, and Z = 4; 2 is monoclinic, space group P21/n, a = 8.861(2), b = 23.089(10), c = 12.014(2) Å, β = 105.84(2)°, V = 2365 Å3, and Z = 2. Both compounds comprise the standard dinuclear rhodium(II) carboxylate unit with the substituted acridine ligands coordinated to rhodium in the axial positions, via the NH2 group nitrogen in 1 and the N(CH3)2 nitrogen in 2.The dimethyl substitution on the tertiary amine group in 2, and an associated conformational change in the diamine chain, result in an increased separation of the acridine ligand from the metal centre. There is a pronounced acridine base stacking in 1 but not in 2.  相似文献   

15.
《Inorganica chimica acta》2006,359(8):2448-2454
The three novel gadolinium(III) containing compounds (NH3CH3)[Gd(CF3CF2COO)4(H2O)] (1), (NH3C2H5)[Gd(CF3CF2COO)4(H2O)] (2) and ((CH3)4N)[Gd(CF3CF2COO)3(H2O)2]CF3CF2COO (3) were synthesized and structurally characterized by X-ray crystallography. In the crystal structures of 1 and 2, the gadolinium ions are bridged by carboxylate groups to dimers with a Gd3+–Gd3+ distance of 451.6(2) (1) and 451.8(3) pm (2), respectively. In the crystal structure of 3 the Gd3+ ions are bridged by carboxylate groups to chains with almost the same Gd3+–Gd3+ distances (494.0(8) and 503.4(7) pm). The magnetic behaviour of 1 and 2 was investigated in the temperature range of 1.76–300 K. The magnetic data indicate weak antiferromagnetic interactions within the dimeric unit.  相似文献   

16.
Electrospray (ESI) mass spectra analysis of acetonitrile solutions of a series of neutral chloro dimers, pincer type, and monomeric palladacycles has enabled the detection of several of their derived ionic species. The monometallic cationic complexes Pd[κ1-C1-N1-S-C(CH3S-2-C6H4)C(Cl)CH2N(CH3)2]+ (1a) and [Pd[κ1-C1-N1-S-C(CH3S-2-C6H4)C(Cl)CH2N(CH3)2](CH3CN)]+ (1b) and the bimetallic cationic complex [κ1-C1-N1-S-C(CH3S-2-C6H4)C(Cl)CH2N(CH3)2]Pd-Cl-Pd[κ1-C1-N1-S-C(CH3S-2-C6H4)C(Cl)CH2N(CH3)2]+ (1c) were detected from an acetonitrile solution of the pincer palladacycles Pd[κ1-C1-N1-S-C(CH3S-2-C6H4)C(Cl)CH2N(CH3)2](Cl) 1. For the dimeric compounds {Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2](μ-Cl)}2 (2, Y=H and 3, CF3), highly electronically unsaturated palladacycles [Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2]+ (2d, 3d) and their mono and di-acetonitrile adducts, namely, [Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2](CH3CN)]+ (2e, 3e) and [Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2](CH3CN)2]+ (2f and 3f) were detected together with the bimetallic complex [Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2]-Cl-Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N](CH3)2]+ (2a, 3a) and its acetonitrile adducts [κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2](CH3CN)Pd-Cl-Pd[ κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2]+ (2b, 3b) and [κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2](CH3CN)Pd-Cl-Pd[κ1-C, κ1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2(CH3CN)]+ (2c, 3c). The dimeric palladacycle {Pd[κ1-C1-N-C(CH3O-2-C6H4)C(Cl)CH2N(CH3)2](μ-Cl)}2 (4) is unique as it behaves as a pincer type compound with the OCH3 substituent acting as an intramolecular coordinating group which prevents acetonitrile full coordination, thus forming the cationic complexes [(C6H4(o-CH3O)CC(Cl)CH2N(CH3)2OCN)Pd]+ (4b), [(C6H4(o-CH3O)CC(Cl)CH2N(CH3)2- κOCN)Pd(CH3CN)]+ (4c) and [(C6H4 (o-MeO)CC(Cl)CH2N(CH3)2O, κCN)Pd-Cl-Pd(C6H4(o-CH3O)CC(Cl)CH2N(CH3)2OCN)]+ (4a). ESI-MS spectra analysis of acetonitrile solutions of the monomeric palladacycles Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2](Cl)(Py) (5, Y=H and 6, Y=CF3) allows the detection of some of the same species observed in the spectra of the dimeric palladacycles, i.e., monometallic cationic 2d-3d, 2e-3e and {Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2](Py)}+ (5a, 6a) and {Pd[κ1-C1-N-C(Y-2-C6H4)C(Cl)CH2N(CH3)2](CH3CN)(Py)}+ (5b, 6b) and the bimetallic 2a, 3a, 2b, 3b, 2c and 3c. In all cationic complexes detected by ESI-MS, the cyclometallated moiety was intact indicating the high stability of the four or six electron anionic chelate ligands. The anionic (chloride) or neutral (pyridine) ligands are, however, easily replaced by the acetonitrile solvent.  相似文献   

17.
Two gadonilium DOTAM complexes [Gd(DOTAM)H2O](CF3SO3)3 · 3H2O (1) and [Gd(DOTAM)H2O](CF3SO3)3 · 0.5H2O · CH3CN (2) have the structure of the M isomer, with coordination geometry around the Gd ion capped square antiprismatic (SA). They differ for the Gd-Owater bond lengths of 2.396(6) and 2.474(7) Å in (2) where two independent molecules are present in the crystal cell. The factors influencing the Gd-Owater bond distance, important for the water exchange rate in MRI experiments, have been correlated to the different network of hydrogen bonds involving the Gd coordinated water molecule. The X-ray structure of the parent compound [Pr(DOTAM)H2O](CF3SO3)3 · H2O · CH3CN (3) reveals an unexpected twisted square antiprismatic (TSA) coordination geometry characteristic of the m isomer.  相似文献   

18.
Phosphorus-31 and cadmium-113 NMR spectroscopy has been used to study the interaction between tertiary phosphines (P(c-C6H11)3 and PBu3) and Cd(O3SCF3)2, Cd(ClO4)2, Cd(CF3COO)2 and Cd(SCN)2 salts in solution. the NMR data imply the formation in solution of novel 1:1 adducts CdX2(phos) (X = O3SCF3, ClO4, CF3COO, phos = P(c-C6H11)3, PBu3) in which there is substantial interaction between the anions and cadmium. Data are also presented for mixed phosphine complexes CdX2[P(c-C6H11)3][PBu3] (X = O3SCF3, ClO4, NO3, CF3COO, CH3COO, Cl, Br, I, SCN). The two bond coupling constants 2J(P′P) of these mixed phosphine complexes decrease as the coordination ability of the anion increases and cover the narrow range from 95 Hz to 66 Hz.  相似文献   

19.
Electron poor cationic complexes [(CF3PCP)Pt(L)]+ (where L = CO, NC5F5, or acetone) react with H2O in polar solvents via selective hydrolysis of a single P-CF3 substituent to afford the spectroscopically-characterized phosphinoyl-bridged complex {k3-P,C,P,k1-O-(CF3)2PCH2C6H3CH2P(CF3)O}2Pt2 (1) in good yield. X-ray diffraction confirms the presence of a six-member Pt-P-O-Pt-P-O ring in a chair conformation. The presumed intermediate aqua complex, (CF3PCP)Pt(H2O)+, is stable in dichloromethane, but when dissolved in more polar solvents readily converts to 1.  相似文献   

20.
Lithioamidines {R′N(Li)C(R)NR′, I; R = CH3, R′ = C6H5, p-CH3,C6H4} react with iron(III) chloride
in monoglyme to produce navy-blue, high spin Fe{R′NC(R)NR′}3 complexes which are extremely air and moisture sensitive. The corresponding reaction when R = R′ = C6H5 produces a soluble red complex and an air-stable green complex, whereas when R = H, R′ = C6H5 and R = R′ = C6H5 and the reaction is started at ca. ?20°, red and green complexes respectively are formed. Though all the complexes are formulated Fe{R′NC(R)NR′}3, their properties reflect association through bridging amidino-groups. Iron(II) chloride reacts with I(R = CH3, R′ = p-CH3C6H4) to form two complexes, one crimson and soluble in organic solvents, and one brown and insoluble, which are fomulated [Fe{R′NC(R)NR′}2]n. The iron(III) complexes failed to react with, or were decomposed by, a variety of reducing, electrophilic and nucleophilic reagents, though blue Fe{p-CH3C6H4NC(CH3)N-p-CH3C6H4}3 reacts readily with nitric oxide to form a purple addition complex from which the N-nitroso-compound p-CH3C6H4NC(CH3)N(NO)-p-CH3C6H4 was obtained in high yield. Treatment of the corresponding brown iron(II) complex with nitric oxide gave no reaction.  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