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 657 毫秒
1.
We used density functional theory to investigate the capacity for carbon monoxide (CO) release of five newly synthesized manganese‐containing CO‐releasing molecules (CO‐RMs), namely CORM‐368 ( 1 ), CORM‐401 ( 2 ), CORM‐371 ( 3 ), CORM‐409 ( 4 ), and CORM‐313 ( 5 ). The results correctly discriminated good CO releasers ( 1 and 2 ) from a compound unable to release CO ( 5 ). The predicted Mn? CO bond dissociation energies were well correlated (R2≈0.9) with myoglobin (Mb) assay experiments, which quantified the formation of MbCO, and thus the amount of CO released by the CO‐RMs. The nature of the Mn? CO bond was characterized by natural bond orbital (NBO) analysis. This allowed us to identify the key donor–acceptor interactions in the CO‐RMs, and to evaluate the Mn? CO bond stabilization energies. According to the NBO calculations, the charge transfer is the major source of Mn? CO bond stabilization for this series. On the basis of the nature of the experimental buffers, we then analyzed the nucleophilic attack of putative ligands (L′=HPO42?, H2PO4?, H2O, and Cl?) at the metal vacant site through the ligand‐exchange reaction energies. The analysis revealed that different L′‐exchange reactions were spontaneous in all the CO‐RMs. Finally, the calculated second dissociation energies could explain the stoichiometry obtained with the Mb assay experiments.  相似文献   

2.
Oxidative addition of diphenyl disulfide to the coordinatively unsaturated [Mn(CO)5]? led to the formation of low-spin, six-coordinate cis-[Mn(CO)4(SPh)2]?. The complex cis-[PPN][Mn(CO)4(SPh)2] crystallized in monoclinic space group P21/c with a = 9.965(2) Å, b = 24.604(5) Å, c = 19.291(4) Å, β = 100.05(2)°, V = 4657(2)Å3, and Z = 4; final R = 0.036 and Rw = 0.039. Thermal transformation of cis-[Mn(CO)4(SPh)2]? to [(CO)3Mn(μ-SPh)3Mn(CO)3]? was completed overnight in THF at room temperature. Additionally, reaction of [Mn(CO)5]? and PhSH in 1:2 mole ratio also led to cis-[PPN](Mn(CO)4(SPh)2]. Presumably, oxidative addition of PhSH to [Mn(CO)4]? was followed by a Lewis acid-base reaction to form cis-[Mn(CO)4(SPh)2]? with evolution of H2.  相似文献   

3.
Anionic iron(0) tetracarbonyl with terminal phenyltellurolate ligand PhTe?, [PhTeFe(CO)4]?, has been synthesized and characterized. The title compound was obtained by addition of (PhTe)2 to [PPN][HFe(CO)4] THF solution dropwise. [PPN][PhTeFe(CO)4] crystallizes in the monoclinic space group C c, with a = 16.119(4) Å, b = 13.141(3) Å, c = 19.880(8) Å, β = 93.04(3)°, V = 4205(2) Å3, and Z = 4. The [PhTeFe(CO)4]? anion is a trigonal-bipyramidal complex in which the phenyltellurolate ligand occupies an axial position with Fe-Te bond length 2.630(5) Å and the Fe-Te-C(Ph) angle is 103.4(5)°. The neutral iron(0)-telluroether compound, (PhTeMe)Fe(CO)4, was prepared by alkylation of the [PhTeFe(CO)4]?. Protonation of [PhTeFe(CO)4]?and reaction of H2Fe(CO)4 and PhTe)2 ultimately lead to formation of the known dimer Fe2(μ-TePh)2(CO)6 and H2.  相似文献   

4.
The arene complexes, (η6-C6H6)Cr(CO)2(CX) (X = S, Se), react with excess CO gas under pressure in tetrahydrofuran at about 60° C to produce the Cr(CO)5(CX) complexes in high yield. The IR and NMR (13C and 17O) spectra of these complexes are in complete accord with the expected C4v molecular symmetry. Like the analogous W(CO)5(CS) complex, both compounds react with cyclohexylamine to give Cr(CO)5(CNC6H11). However, while W(CO)5(CS) undergoes stereospecific CO substitution with halide ions (Y? to form trans-[W(CO)4(CS)Y]?, the two chromium chalcocarbonyl complexes apparently undergo both CO and CX substitution to afford mixtures of [Cr(CO)5Y]? and trans-[Cr(CO)4(CX)Y]?.  相似文献   

5.
The addition of CCl4 to hex-1-ene and to the methyl ester ofN-(trans-cinnamoyl)-l-proline (2) catalyzed by M3(CO)12 or by the M3(CO)12+DMF system (M=Fe, Ru, Os) was studied. The use of ruthenium and osmium dodecacarbonyls in combination with DMF increases the yields of adducts CCl3CH2CHClC4H9 (4) and PhCHClCH(CCl3)C(O)R′ (3) over those obtained in reactions catalyzed by the same carbonyls without DMF. In addition to adduct3, salts [M(CO3)Cl3][Me2NH2]+ were isolated from the products of the reaction between CCl4 and1 in the presence of M3(CO)12+DMF (M=Ru, Os). These salts do not catalyze this reaction and apparently result from chain termination. Experimental results in favor of a coordination mechanism of the addition of CCl4 to olefins in the presence of Ru3(CO)12 and Os3(CO)12 were obtained. Translated fromIzvestiya Akademii Nauk Seriya Khimicheskaya, No. 6, pp. 1174–1179, June, 1997.  相似文献   

6.
The thermal gas‐phase reactions of rhenium carbonyl complexes [Re(CO)x ]+ (x =0–3) with methane have been explored by using FT‐ICR mass spectrometry complemented by high‐level quantum chemical calculation. While it had been concluded in previous studies that addition of closed‐shell ligands in general decreases the reactivity of metal ions, the current work provides an exception: the previously demonstrated inertness of atomic Re+ towards methane is completely changed upon ligation with CO. Both [Re(CO)]+ and [Re(CO)2]+ bring about efficient dehydrogenation of the hydrocarbon under ambient conditions. However, addition of a third ligand to form [Re(CO)3]+ completely quenches the reactivity.  相似文献   

7.
Regioselectivity of the addition of the highly functionalized zinc-copper reagents to (η3-allyl)Fe(CO)4 cationic salts was studied. For 1,1-disubstituted allyl cation 1, the zinc-copper reagents added predominantly at the unsubstituted terminus. For 1,1,2-trisubstituted allyl cation 2, reactive zinc-copper reagents attacked mainly at the unsubstituted terminus while less reactive zinc-copper reagents added to a coordinated CO ligand. For 1,1,3-trisubstituted allyl cation 3, the addition occurred at both the less substituted allyl terminus and a coordinated CO ligand.  相似文献   

8.
The reaction of trans-IrCl(CO)L2 with pz?1 gives trans-Ir(pz-N)(CO)L2, where pzH is 3,5-dimethyl-, 3,5-dimethyl-4-nitro- or 3,5-bis(trifluoromethyl)-pyrazole, and L = PPh3. The nitrogen atom not involved in coordination can be protonated with HBF4 to give the corresponding [Ir(CO)L2(pzH-N]+ cation. The iridium(I) pyrazolates undergo oxidative addition, yielding Ir(H)2(pz-N)(CO)L2 species, while gaseous HCl cleaves the IrN bond, affording IrH(Cl)2(CO)L2. The iridium(I) derivatives can be obtained in several solid-state forms, each characterized by a slightly different CO stretching frequency. The presence of a monodentate pyrazolato ligand in trans-Ir(3,5-(CF3)2pz-N)(CO)L2, in the form with ν(CO) at 1975 cm?1, is supported also by an X-ray crystal structure determination. The compound crystallizes in the monoclinic system, space group P21/n, with cell dimensions a = 21.106(6), b = 19.700(5), c = 9.437(2) Å, and β = 94.34(2)° and Z = 4.  相似文献   

9.
We demonstrate that RuII(CO)2–protein complexes, formed by the reaction of the hydrolytic decomposition products of [fac‐RuCl(κ2‐H2NCH2CO2)(CO)3] (CORM‐3) with histidine residues exposed on the surface of proteins, spontaneously release CO in aqueous solution, cells, and mice. CO release was detected by mass spectrometry (MS) and confocal microscopy using a CO‐responsive turn‐on fluorescent probe. These findings support our hypothesis that plasma proteins act as CO carriers after in vivo administration of CORM‐3. CO released from a synthetic bovine serum albumin (BSA)–RuII(CO)2 complex leads to downregulation of the cytokines interleukin (IL)‐6, IL‐10, and tumor necrosis factor (TNF)‐α in cancer cells. Finally, administration of BSA–RuII(CO)2 in mice bearing a colon carcinoma tumor results in enhanced CO accumulation at the tumor. Our data suggest the use of RuII(CO)2–protein complexes as viable alternatives for the safe and spatially controlled delivery of therapeutic CO in vivo.  相似文献   

10.
We demonstrate that RuII(CO)2–protein complexes, formed by the reaction of the hydrolytic decomposition products of [fac‐RuCl(κ2‐H2NCH2CO2)(CO)3] (CORM‐3) with histidine residues exposed on the surface of proteins, spontaneously release CO in aqueous solution, cells, and mice. CO release was detected by mass spectrometry (MS) and confocal microscopy using a CO‐responsive turn‐on fluorescent probe. These findings support our hypothesis that plasma proteins act as CO carriers after in vivo administration of CORM‐3. CO released from a synthetic bovine serum albumin (BSA)–RuII(CO)2 complex leads to downregulation of the cytokines interleukin (IL)‐6, IL‐10, and tumor necrosis factor (TNF)‐α in cancer cells. Finally, administration of BSA–RuII(CO)2 in mice bearing a colon carcinoma tumor results in enhanced CO accumulation at the tumor. Our data suggest the use of RuII(CO)2–protein complexes as viable alternatives for the safe and spatially controlled delivery of therapeutic CO in vivo.  相似文献   

11.
Conclusions The reaction of [CpW(CO)3]2 with PhCH2Cl and the decomposition of CpW(CO)3Me were studied by ESR spectroscopy using 2,4,6-tri-tert-butyinitrobenzene (BNB) and -phenyl-N-tert-butylnitrone (PNB) as spin traps. The rate constants for the addition of CpW(CO)3 radicals to the spin traps at 20°C were determined by identification of the intermediate radicals in these reactions: k ad BNB =6.7·105-7.0·105 and k ad PNB =5.8·104-6.1·104 liters/mole · sec.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 11, pp. 2631–2633, November, 1984.  相似文献   

12.
Three half‐sandwich iron(II) complexes, [Fe(η5‐Cp)(cis‐CO)2X] (X?=Cl?, Br?, I?), were synthesized and characterized. The kinetics of the CO‐releasing behaviour of these complexes upon illumination by visible irradiation in various media was investigated. Our results indicated that the CO release was significantly affected by the auxiliary ligands. Of the three light sources used (blue, green, and red), blue light exhibited the highest efficiency. In the photoinduced CO release, the solvents and exogenous nucleophiles in the media were involved, which allowed their CO‐releasing reaction to comply with pseudo first‐order model rather than the characteristic zero‐order model for a photochemical reaction. In aqueous media (D2O), an intermediate bearing the core of {FeII(cis‐CO)2} involving cleavage of cyclopentadiene was detected. Despite the non‐absorption of the red light, its illumination combined with nucleophilic substitution did cause considerable CO release. Assessment of the cytotoxicity of the three complexes indicated that they showed good biocompatibility.  相似文献   

13.
The possibility of formation of arylphosphinidene complexes with tungsten pentacarbonyl ArP=W(CO)5 upon the nucleophilic substitution of the chlorine atoms in aryldichlorophosphines ArPCl2 by the electrochemically generated [W(CO)5]2− anion is demonstrated. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1356–1359, June, 2005.  相似文献   

14.
Nitrosylation reactions are rare in the context of low valent Re(I)- and Tc(I)-tricarbonyl complexes so far. We herein describe a method for the conversion of a “M(CO)3-moiety” (M = Re, Tc) into a dicarbonyl-nitrosyl moiety “M(CO)2NO”, the synthesis of important precursor complexes and intermediates and possible applications for this new kind of Re- and Tc-chemistry.The behavior of the complex [ReCl3(CO)2(NO)] in water was studied in detail and compared to that of [ReCl3(CO)3]2−. Contrary to the conversion of [ReCl3(CO)3]2− to the mixed aquo-carbonyl complex [Re(OH2)3(CO)3]+ in water, one chloride remains initially bound to the metal center in the dicarbonyl-nitrosyl complex, making [ReCl(OH2)2(CO)2(NO)]+ the main species for further reactions. In this context, we isolated and characterized the complex [Re(μ3-O)(CO)2(NO)]4. Examples of complexes with different bi- and tridentate ligands based on ReCl3(CO)2(NO)] are discussed.For the development of potential new radiopharmaceuticals we also adapted the nitrosylation technique to the n.c.a. level with 99mTc. [99mTc(OH2)3(CO)3]+ served as starting material to form a 99mTc(CO)2(NO)-core. Labelling reactions with ligands such as iminodiacetic acid (IDA), nitrilotriacetic acid (NTA) and diethylenetriamine pentaacetic acid (DTPA) were performed, resulting in the complexes [99mTc(IDA)(CO)2(NO)], [99mTc(NTA)(CO)2(NO)] and [99mTc(DTPA)(CO)2(NO)]. In this way, the “nitrosyl-approach” adds a new and challenging synthetic tool to the already established organometallic chemistry of Re- and Tc-tricarbonyl complexes.  相似文献   

15.
The heat of reaction for addition of iodine to the planar complexes RhCl(CO)dppe, Rh(dppen)2+, Rh(dppe)2+ and to the dimer Rh2Cl2(CO)2(dppm)2 has been obtained by a calorimetric method. Iodine forms stronger Rh---I bonds with RhCl(CO)dppe than with the bichelate complexes. The presence of metal-metal interaction in the iodine addition compound of Rh2Cl2(CO)2(dppm)2 makes a significant contribution to the enthalpy change for the oxidative addition. The stereochemistry of the complexes are discussed on the basis of IR and 31P NMR spectra.  相似文献   

16.
The thermal dissociation of gaseous Mo(CO)6 and W(CO)6 in an argon carrier gas, Mo(CO)6 → Mo(CO)5 + CO (1) and W(CO)6 → W(CO)5 + CO (2), is studied over temperature ranges of ∼585–685 K for (1) and ∼690−810 K for (2) at a total gas concentrations of 4 × 10−6 and 4 × 10−5 mol/cm3 by using the shock tube technique in conjunction with absorption spectrophotometry. The measured rate constants are extrapolated to the high-pressure limit by means of a newly developed procedure, with the resultant expressions for the indicated temperature ranges reading as kd1,∞(T),[s−1] = 1016.12 ± 0.68exp[(−148.8 ± 8.1 kJ/mol)/RT] and kd2,∞(T),[s−1] = 1015.93 ± 0.63exp[(−171.7 ± 8.9 kJ/mol)/RT]. Comparison of the high-pressure dissociation rate constants with the published data revealed a considerable discrepancy, a tentative explanation of which is given. Based on the obtained high-pressure dissociation rate constants and the available data on the high-pressure room-temperature rate constants for the reverse reaction of recombination, the first bond dissociation energies for these molecules are evaluated and compared with previous determinations, both theoretical and experimental. The enthalpies of formation of Mo(CO)5 and W(CO)5 are determined: ΔfH°(Mo(CO)5, g, 298.15 K) = −644.1 ± 5.6 kJ/mol and ΔfH°(W(CO)5, g, 298.15 K) = −581.9 ± 6.6 kJ/mol. Based on the enthalpies of formation of Mo(CO)5, W(CO)5, Mo(CO)6, and W(CO)6, and the published molecular parameters of these four species, their thermochemical functions are calculated and presented in the form of NASA seven-term polynomials.  相似文献   

17.
Imines, Im, such as MeN=C(Ph)H (5), 2-methyl 4,5-dihydrothiazole (8a), 2-methyl 4,5-dihydrooxazole (8b) and MeN=C(OMe)Me (13) add to the α-carbon atom of the vinylidene ligand in [(CO)5Cr=C=CMe2] (4) to give isolable zwitterionic adducts, [(CO)5Cr–C(=CMe2)(Im+)]. The reaction of [(CO)5W=C=CPh2] (12) with 13 also yields an adduct, [(CO)5W–C(=CPh2){NMe=C(OMe)Me}+] (15), whereas from the corresponding reaction of 4 with xanthylideneimine, H–N=C(C6H4)2O (16), a carbene complex, [(CO)5Cr=C(i-Pr)–N=C(C6H4)2O] (17), is obtained. Complex 17 presumably is formed by initial addition of 16 to 4 and subsequently rapid rearrangement. In solution, the adduct [(CO)5Cr–C(=CMe2)(NMe=C(Ph)H)+] (6) slowly cyclizes to form the 2-azetidin-1-ylidene complex [(CO)5Cr= Me2] (7). In contrast, when solution of those zwitterions are heated that are formed by addition of 4,5-dihydrothiazole or 4,5-dihydrooxazole to 4, no cyclization is observed but rather the formation of 4,5-dihydrothiazole and 4,5-dihydrooxazole complexes, respectively. The structures of two adducts, [(CO)5Cr–C(=CMe2)(Im+)] (Im=MeN=C(Ph)H, 2-methyl 4,5-dihydrothiazole) and of the substitution product [(CO)5W(2-methyl 4,5-dihydrothiazole)] have been established by X-ray structural analyses.  相似文献   

18.
Linear trimetallic MPPtIIL2M complexes (M = Cr(CO)3(η-C5H5), Mo(CO)3- (η-C5H5), W(CO)3(η-C5-H5), Mn(CO)5, Fe(CO)3NO, Co(CO)4; L = t-BuNC, cyclo- C0H11 NC) are reduced on platinum and gold electrodes in non-aqueous medium. All these complexes undergo irreversible one electron reductions, which result in the rupture of one Ptmetal bond and the liberation of one M? ion per mole reduced. Coupled ESR spectroscopy and coulometry show that a radical is generated during the reduction of the trimetallic complexes. The several ESR signals obtained for these paramagnetic Pt1 species exhibit no hyperfine structure.The electrochemical behaviour of MPtL2M complexes is compared with that of the following linear trimetallic complexes: MHgM and (MAuM)?.  相似文献   

19.
The reaction of H2Os5(CO)15 with nucleophiles Y leads either to deprotonation (Y = OH? or Me?) or addition (Y = I?, P(OMe)3, or CO). The geometrical consequences of these reactions are discussed.  相似文献   

20.
Hydrogenation of Aromatic Nitriles on the Fe3(CO)9 Cluster The μ3-nitrile bridged clusters Fe3(CO)932-N≡CR) ( 3 , R = phenyl, p-tolyl, p-anisyl) consume hydrogen upon heating in solution with formation of the acimidoyl- and the alkylideneimido-bridged clusters HFe3(CO)932-HN=CR) ( 1 ) and HFe3(CO)932-N=CHR) ( 2 ). These can be obtained in a better way by successive H+ and H addition with NaBH4 and H3PO4. HFe3(CO)932-N=CHR) ( 2 ) adds P(OMe)3 with concomitant hydrogen migration to form Fe3(CO)9P(OMe)331-N–CH2R) ( 6 ). The phosphite-substituted cluster Fe3(CO)8P(OMe)332-N≡CPh) ( 5 a ) on the other hand is converted by the H+/H addition to the products HFe3(CO)8P(OMe)332-HN=CPh) ( 7 a ) and HFe3(CO)8P(OMe)332-N=CHPh) ( 8 a ).  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