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.
Syntheses and NMR Spectroscopic Ivestigations of Salts containing the Novel Anions [PtXn(CF3)6‐n]2— (n = 0 ‐ 5, X = F, OH, Cl, CN) and Crystal Structure of K2[(CF3)2F2Pt(μ‐OH)2PtF2(CF3)2]·2H2O The first syntheses of trifluoromethyl‐complexes of platinum through fluorination of cyanoplatinates are reported. The fluorination of tetracyanoplatinates(II), K2[Pt(CN)4], and hexacyanoplatinates(IV), K2[Pt(CN)6], with ClF in anhydrous HF leads after working up of the products to K2[(CF3)2F2Pt(μ‐OH)2PtF2(CF3)2]·2H2O. The structure of the salt is determined by a X‐ray structure analysis, P21/c (Nr. 14), a = 11.391(2), b = 11.565(2), c = 13.391(3)Å, β = 90.32(3)°, Z = 4, R1 = 0.0326 (I > 2σ(I)). The reaction of [Bu4N]2[Pt(CN)4] with ClF in CH2Cl2 generates mainly cis‐[Bu4N]2[PtCl2(CF3)4] and fac‐[Bu4N]2[PtCl3(CF3)3], but in contrast that of [Bu4N]2[Pt(CN)6] with ClF in CH2Cl2 results cis‐[Bu4N]2[PtX2(CF3)4], [Bu4N]2[PtX(CF3)5] (X = F, Cl) and [Bu4N]2[Pt(CF3)6]. In the products [Bu4N]2[PtXn(CF3)6‐n] (X = F, Cl, n = 0—3) it is possibel to exchange the fluoro‐ligands into chloro‐ and cyano‐ligands by treatment with (CH3)3SiCl und (CH3)3SiCN at 50 °C. With continuing warming the trifluoromethyl‐ligands are exchanged by chloro‐ and cyano‐ligands, while as intermediates CF2Cl and CF2CN ligands are formed. The identity of the new trifluoromethyl‐platinates is proved by 195Pt‐ and 19F‐NMR‐spectroscopy.  相似文献   

2.
Copolymerization of fluorine ring-substituted 2-phenyl-1,1-dicyanoethenes, RC6H3CH?C(CN)2 (R is 2,3-F,F, 2,4-F,F, 2,5-F,F, 2,6-F,F, and 4-CF3) with 4-fluorostyrene were prepared in the presence of a radical initiator (ABCN) at 70°C. The composition of the copolymers was calculated from nitrogen analysis, and the copolymers were characterized by IR, 1H and 13C-NMR, GPC, DSC, and TGA. The monomer reactivity ratios for 4-fluorostyrene (M1), r1 = 0.6 and 2-(2,4-difluorophenyl)-1,1-dicyanoethene (M2), r2 = 0 were determined from Fineman-Ross plot. The order of relative reactivity (1/r1) for difluoro-substituted monomers is 2,4-F,F (0.31) > 2,3-F,F (0.25) > 2,5-F,F (0.22) > 2,6-F,F (0.10). DSC curves showed that the copolymers were amorphous with high T g in comparison with that poly(4-fluorostyrene) indicating a substantial decrease in chain mobility of the copolymer due to the high dipolar character of the trisubstituted ethylene monomer units. From the thermogravimetric analysis, the copolymers began to degrade in the range 214–260°C. The copolymer of 4-fluorostyrene and 2-(2,4-difluorophenyl)-1,1-dicyanoethene and poly(4-fluorostyrene) were dielectrically characterized in the range 25–200°C. The dominating relaxation process detected in both materials was the α-relaxation, associated with the dynamic glass transition. The relationship polarity-permittivity was discussed.  相似文献   

3.
Novel trisubstituted ethylenes, halogen ring-substituted butyl 2-cyano-3-phenyl-2-propenoates, RPhCH ? C(CN)CO2C4H9 (where R 2-Br, 3-Br, 4-Br, 2-Cl, 3-Cl, 4-Cl, 2-F, 3-F, 4-F) were prepared and copolymerized with styrene. The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring-substituted benzaldehydes and butyl cyanoacetate, and characterized by CHN analysis, IR, 1H and 13C-NMR. All the ethylenes were copolymerized with styrene (M1) in solution with radical initiation (ABCN) at 70°C. The compositions of the copolymers were calculated from nitrogen analysis and the structures were analyzed by IR, 1H and 13C?NMR. The order of relative reactivity (1/r1) for the monomers is 3-F (22.97) > 3-Br (6.05) > 2-Cl (4.32) > 4-F (3.19) > 4-Br (2.47) > 4-Cl (1.92) > 2-F (1.74) > 3-Cl (1.10) > 2-Br (1.07). Decomposition of the copolymers in nitrogen occurred in two steps, first in the 200–500°C range with residue (2.3–5.9% wt.), which then decomposed in the 500–800°C range.  相似文献   

4.
Perfluoromethyl-Element-Ligands. XXXV. Reactivity of Metallated Phosphanes and Arsanes of the Type π-C5H5(CO)3MER2 (M ? Cr, Mo, W; E ? P, As; R ? CF3, CN) The influence of the complex fragments π-C5H5(CO)3M (M ? Cr, Mo, W) on the basicity of the metallated phosphanes or arsanes π-C5H5(CO)3MER2 (E ? P, As; R ? CF3, CN) has been investigated by reactions with sulfur, methyliodide, fluorotrichloromethane, and W(CO)5THF, respectively. π-C5H5(CO)3ME(CF3)2 (E ? P: 1a–c ; E ? As: 2a–c ) react with sulfur only for E ? P to give the complexes π-C5H5(CO)3P(S)(CF3)2 ( 5a–c ) in good yield. The attempted thermal transformation of the phosphane sulfides to η2 coordinated (CF3)2P?S complexes proves unsuccessful. The reactions of 1a–c, 2a–c and π-C5H5(CO)3MP(CN)2 ( 3a–c ) with CH3I or CCl3F do not lead to onium salts, but to cleavage of the M–E bonds forming π-C5H5(CO)3MX (X ? I, Cl) and CH3ER2 and R2ECCl2F, respectively. The reactivity depends on ER2 and M: P(CF3)2 > P(CN)2 > As(CF3)2; Cr > Mo > W. Due to the low donor ability of the complexes 1a–c, 2a–c and 3a–c binuclear compounds π-C5H5(CO)3MER2W(CO)5 (E ? As, R ? CF3: 11a–c ; E ? P, R ? CN: 12a–c ; ER2?P(CN)Ph: 13a, b ) are obtained only with the highly reactive W(CO)5THF. In case of the (CF3)2P bridged derivatives spontaneous CO-elimination leads to the threemembered ring systems ( 10a–c ).  相似文献   

5.
6.
Abstract

Full geometry optimizations were carried out on the singlet and triplet states of β-substituted divalent five-membered rings XC4H3M (X? ?NH2, ?OH, ?CH3 ?H, ?CH3, ?Br, ?Cl, ?F, ?CF3, and ?NO2; M?C, Si, and Ge) by the B3LYP method by using 6-311++G** basis set. The thermal energy gaps, ΔEt–s; enthalpy gaps, ΔHt–s; and Gibbs free energy gaps, ΔGt–s, between the singlet (s) and triplet (t) states of the above structures were calculated by using the GAUSSIAN 03 program. The ΔGt–s of XC4H3C was changed in the order: X? ?Cl > ?Br > ?CH3 > ?H > ?CF3 > ?F > ?NO2 > ?OH > ?NH2. The changes of ΔGt–s for XC4H3Si and XC4H3Ge were in the order: X? ?NH2 > OH > F > Cl > Br > CH3 > H > CF3 > NO2. The geometrical parameters, including bond lengths (R), bond angles (A), dihedral angles (D), natural bonding orbital (NBO) charge at atoms, HOMO and LUMO, and dipole moments, were presented and discussed.

Supplemental materials are available for this article. Go to the publisher's online edition of Phosphorus, Sulfur, and Silicon and the Related Elements to view the free supplemental file.

GRAPHICAL ABSTRACT   相似文献   

7.
Novel copolymers of trisubstituted ethylene monomers, halogen ring-disubstituted 2-phenyl-1,1-dicyanoethylenes, RC6H3CH= C(CN)2 (where R is 2,3-Cl2, 2,4-Cl2, 2,6-Cl2, 3,4-Cl2, 3,5-Cl2, 2-Cl-4-F, 2-Cl-6-F, 3-Cl-4-F) and 4-fluorostyrene were prepared at equimolar monomer feed composition by solution copolymerization in the presence of a radical initiator (ABCN) at 70°C. The composition of the copolymers was calculated from nitrogen analysis, and the structures were analyzed by IR, 1H and 13C-NMR. The order of relative reactivity (1/r 1) for the monomers is 2-Cl-4-F (2.42) > 3,4-Cl2(2.40) > 2,4-Cl2(1.97) > 2-Cl-6-F (1.86) > 3-Cl-4-F (1.68) > 2,3-Cl2 (0.89) > 3,5-Cl2 (0.70) > 2,6-Cl2 (0.47). High Tg of the copolymers, in comparison with that of poly(4-fluorostyrene) indicates a substantial decrease in chain mobility of the copolymer due to the high dipolar character of the trisubstituted ethylene monomer unit. Softening of the copolymers started in 194–216°C range. Decomposition of the copolymers in nitrogen occurred in two steps, first in the 290–400°C range with residue, which then decomposed in 400–800°C range.  相似文献   

8.
The electroreduction of the halofluoromethanes CF3Br, CF2Br2 and CF2BrCl has been studied in high‐pressure stainless steel autoclaves at different cathodes [Pt, steel (V2A, V4A), glassy carbon (GC)] and in various solvent‐supporting electrolyte systems (SSE), e.g. DMF/[Bu4N]Br, NMP/[Bu4N]BF4 etc. The reduction potentials for CF3Br increase from Pt (–1.6 V) < V2A (–1.8 V) < GC (–2.1 V) and are lower for CF2Br2 and CF2BrCl suggesting a reductive cleavage of C‐X bonds as the first step. CF2Br2 and CF2BrCl show a two‐step reduction in accord with the C–X bond energies (C–F > C–Cl > C–Br) and the “Perfluoro‐effect”. The electrolysis of CF3Br in different SSE‐systems with sacrificial zinc or cadmium anodes has been reinvestigated with our experimental set‐up to elucidate the influence of the experimental conditions on the type and ratio of the products. The observed products CF3MBr·42L and (CF3)2M·42L (M = Zn, Cd; L = DMF or AN) are the same as in the previous investigations, but are obtained in different ratios, as a rule caused by a parallel chemical corrosion of the respective anodes. By using aluminium as sacrificial anode no CF3Al compounds are formed. The CF3 species generated by electroreduction of CF3Br react with the solvents via hydrogen abstraction and formation of CF3H. The current yield with respect to the dissolution of the Al anode reaches 120 % indicating a considerable chemical corrosion in addition to the anodic oxidation. This result enabled a one‐pot trifluoromethylation reaction of NMP as organic carbonyl substrate and solvent with CF3Br and aluminium powder (ratio 3 : 2) at higher temperatures (> 70 °C). The complete reaction of CF3Br to give CF3H and 1‐methyl‐2‐trifluoromethyl‐4,5‐dihydropyrrol allowed the isolation of the latter by vacuum condensation and distillation in 45 % yield, rel. to the CF3Br used. Gallium and indium were also applied as sacrificial anodes in combination with CF3Br as substrate. In both cases, anodic current yields of about 280 % indicated an extreme chemical corrosion together with cathodic metal depositions corresponding to the cathodic current yield. These deposits – in contrast to those of Zn and Cd – do not react with CF3Br in Grignard‐type conversions to CF3Ga and CF3In compounds. So, the observed products (CF3)nMBr3–n·L (M = Ga, In; n 1‐3; L = DMF, NMP) are obviously formed by chemical corrosion of the electro‐activated anodes. Finally, electrochemical and chemical trifluoromethylations were successfully carried out, using R3SiCl (R = Me, Vi, Ph), Me3M′Cl (M′ = Ge, Sn) and aluminium anodes or Al‐powder. The products were characterized either after isolation or in the product solutions by NMR‐spectroscopic investigations.  相似文献   

9.
《中国化学会会志》2018,65(5):613-627
The general species (2,2′‐bpy)MX2 (M = Pd, Pt; X = Br, I) in a crystallization process results in an isomorphous convergence in P21/c. Yet, with polyfluorinated side chains, the general [5,5′‐(HCF2CF2CH2OCH2)2‐2,2′‐bpy]MX2 species proceeds to crystallize the isomorphous structures of 5 (M = Pt; X = I) and 6 (M = Pd; X = I) in P21/c only; structure 7 (M = Pt; X = Br) crystallizes in P21/c but is not isomorphous with 5 and 6 , and structure 8 (M = Pd; X = Br) forms differently in P–1. The causes making the system nonlinear are (1) the intramolecular CF2─HX(─M) hydrogen bonds found in 5–7 but not in 8, and (2) in response to the transition from I to Br, bifurcated [C─H]2 F ─C hydrogen bonds that are formed in 5 and 6 and bifurcated C─ H [F─C]2 hydrogen bonds in 7 . Additionally, the intramolecular CF2─HX(─M) hydrogen bonding from compounds 5–7 could be affirmed by the IR studies.  相似文献   

10.
Ab initio molecular orbital theory has been used to probe the effect of the substituent X on the structures, strain energies, stabilization energies, inversion barriers, and proton affinities of carbanions CH3CX CH and cis-C3H4X?, where X = H, F, CN, and NC. All geometries have been optimized with a 3-21G basis set, and the parent anions (X = H) were also optimized with the same basis set with a diffuse function added (i.e. the 3-21 + G basis set). The anions, with the exception of the α-cyanoisopropyl anion, are pyramidal. The out-of-plane angle, α, for the pyramidal anions decreases in the order F > H ≈ NC > CN, and the barriers to inversion follow the same order with the cyclopropyl anions consistently having higher barriers than the isopropyl anions. The substituents strongly stabilize the anions with the stabilization energy following the order CN > NC > F. The cyano group slightly reduces the strain energy of cyclopropane, but the isocyano and fluoro substituents are weakly and strongly destabilizing, respectively. The pyramidal cyclopropyl anions are less strained than the cyclopropanes except when the substituent is a cyano group where the strain energies are reversed but are very similar. The planar anions all have higher strain energies than the cyclopropanes.  相似文献   

11.
Electrophilic trisubstituted ethylenes, ring-disubstituted methyl 2-cyano-3-phenyl-2-propenoates, RPhCH?C(CN)CO2CH3, where R is 2,5-dichloro, 3,5-dichloro, 2,3-difluoro, 3-chloro-2-fluoro, 3-chloro-4-fluoro, 4-chloro-3-fluoro, 2-chloro-5-nitro, and 2-chloro-6-nitro were prepared and copolymerized with styrene. The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring-disubstituted benzaldehydes and methyl cyanoacetate, and characterized by CHN analysis, IR, 1H and 13C-NMR. All the ethylenes were copolymerized with styrene (M1) in solution with radical initiation (ABCN) at 70°C. The compositions of the copolymers were calculated from nitrogen analysis and the structures were analyzed by IR, 1H and 13C-NMR. The order of relative reactivity (1/r1) for the monomers is 4-Cl-3-F (4.87) > 2,3-F2 (4.49) > 3-Cl-4-F (3.50) > 3-Cl-2-F (2.96) > 2-Cl-5-NO2 (2.02) > 2,5-Cl2 (1.54) > 2-Cl-6-NO2 (1.00) > 3,5-Cl2 (0.41). Relatively high Tg of the copolymers in comparison with that of polystyrene indicates a decrease in chain mobility of the copolymer due to the high dipolar character of the trisubstituted ethylene monomer unit. Decomposition of the copolymers in nitrogen occurred in two steps, first in the 200–500ºC range with residue (1.5–34.5% wt), which then decomposed in the 500-800ºC range.  相似文献   

12.
The effects of several substituents (? BH2, ? BF2, ? AlH2, ? CH3, ? C6H5, ? CN, ? COCH3, ? CF3, ? SiH3, ? NH2, ? NH3+, ? NO2, ? PH2, ? OH, ? OH2+, ? SH, ? F, ? Cl, ? Br) on the Bergman cyclization of (Z)‐1,5‐hexadiyne‐3‐ene (enediyne, 3 ) were investigated at the Becke–Lee–Yang–Parr (BLYP) density functional (DFT) level employing a 6‐31G* basis set. Some of the substituents (? NH3+, ? NO2, ? OH, ? OH2+, ? F, ? Cl, ? Br) are able to lower the barrier (up to a minimum of 16.9 kcal mol?1 for difluoro‐enediyne 7rr ) and the reaction enthalpy (the cyclization is predicted to be exergonic for ? OH2+ and ? F) compared to the parent system giving rise to substituted 1,4‐dehydrobenzenes at physiological temperatures. © 2001 John Wiley & Sons, Inc. J Comput Chem 22: 1605–1614, 2001  相似文献   

13.
Henry's law constants of six kinds of hydrochlorofluorocarbons (HCFCs) were determined at 313–353 K by means of a phase‐ratio variation headspace method: and (KH353 in M atm?1, ΔHsol in kJ mol?1) = (0.0070 ± 0.0006, –23 ± 2), (0.0038 ± 0.0011, –22 ± 10), (0.0065 ± 0.0007, –21 ± 3), (0.0026 ± 0.0007, –23 ± 8), (0.0016 ± 0.0003, –30 ± 4), and (0.0022 ± 0.0003, –25 ± 4), respectively, for HCFC‐141b (CH3CCl2F), HCFC‐142b (CH3CClF2), HCFC‐123 (CF3CHCl2), HCFC‐124 (CF3CHClF), HCFC‐225ca (CF3CF2CHCl2), and HCFC‐225cb (CClF2CF2CHClF). Errors represent two standard deviations only for the fitting. The decay of headspace partial pressures of these HCFCs via hydrolysis was discerned only for CF3CHCl2 and CF3CF2CHCl2 under the experimental conditions examined. Rate constants (kOH– in M?1 s?1) for aqueous reactions of CF3CF2CHCl2 and CF3CHCl2 with OH? at 313–353 K were determined to be and , respectively, from monitoring changes in headspace partial pressure over prescribed concentrations of aqueous NaOH as a function of the headspace time duration and concentration of aqueous NaOH. The calculations performed included consideration of gas–water equilibrium and hydrolysis at both headspace and room temperatures. The calculation for CF3CHCl2 also included consideration of salting‐out effects: The salting coefficient of NaCl on a natural log basis was determined to be 0.36 ± 0.06 M?1, and this value was used for consideration of the salting‐out effect of NaOH. Whereas the activation energy for CF3CF2CHCl2 was greater than that for CF3CHCl2, the kOH– value at 353 K of CF3CF2CHCl2 was 103 times larger than that of CF3CHCl2, indicating that reaction mechanisms for these two HCFCs differed from each other. The aqueous reaction of CF3CF2CHCl2 with OH? was found to proceed through dehydrofluorination on the basis of detection of CF3CF?CCl2 as a primary degradation product of the reaction and proportionality of the rate constants to both concentrations of CF3CF2CHCl2 and OH?.  相似文献   

14.
A series of sterically encumbered [Pt( L )(σ‐acetylide)2] complexes were prepared in which L , a dendritic polyaromatic diimine ligand, was held constant ( L =1‐(2,2′‐bipyrid‐6‐yl)‐2,3,4,5‐tetrakis(4‐tert‐butylphenyl)benzene) and the cis ethynyl co‐ligands were varied. The optical properties of the complexes were tuned by changing the electronic character, extent of π conjugation and steric bulk of the ethynyl ligands. Replacing electron‐withdrawing phenyl‐CF3 substituents ( 4 ) with electron‐donating pyrenes ( 5 ) resulted in a red shift of both the lowest‐energy absorption (ΔE=3300 cm?1, 61 nm) and emission bands (ΔE=1930 cm?1, 64 nm). The emission, assigned in each case as phosphorescence on the basis of the excited‐state lifetimes, switched from being 3MMLL′CT‐derived (mixed metal–ligand‐to‐ligand charge transfer) when phenyl/polyphenylene substituents ( 3 , 4 , 6 ) were present, to ligand‐centred 3ππ* when the substituents were more conjugated aromatic platforms [pyrene ( 5 ) or hexa‐peri‐hexabenzocoronene ( 7 )]. The novel PtII acetylide complexes 5 and 7 absorb strongly in the visible region of the electromagnetic spectrum, which along with their long triplet excited‐state lifetimes suggested they would be good candidates for use as singlet‐oxygen photosensitisers. Determined by in situ photooxidation of 1,5‐dihydroxynaphthalene (DHN), the photooxidation rate with pyrenyl‐ 5 as sensitiser (kobs=39.3×10?3 min?1) was over half that of the known 1O2 sensitiser tetraphenylporphyrin (kobs=78.6×10?3 min?1) under the same conditions. Measured 1O2 quantum yields of complexes 5 and 7 were half and one‐third, respectively, of that of TPP, and thus reveal an efficient triplet–triplet energy‐transfer process in both cases.  相似文献   

15.
The crystal structure of the title compound, C9H6F3N, at 123 K contains mol­ecules linked together via several C—H?F and C—H?N contacts, the strongest of which are 2.58 and 2.65 Å, respectively. Apparently, an F atom in the CF3 group is able to compete with a cyano N atom for aromatic H atoms but is less prone to interact with the more acidic methyl­ene H atoms. The Ph–CH2CN torsion angle is ?6.4 (2)° and the planar phenyl ring exhibits a typical deformation of the endo angles at the ipso‐C atoms, due to the difference in the electron‐withdrawing power of the CF3 and CH2CN substituents.  相似文献   

16.
Density‐functional theory calculations of a series of organic biradicals on the basis of the N,N′‐dioxy‐2,6‐diazaadamantane core with different substituents at carbon atoms adjacent to the nitroxyl groups have been performed by the UB3LYP/6‐311++G(2d,2p) method. Using the breaking symmetry approach, the values of the exchange interaction parameter, J, between the radical centers are calculated. It is shown that the intramolecular exchange interaction for the most part is ferromagnetic in nature, but the J parameter gradually decreases, changing its sign to antiferromagnetic interaction for the last substituent in the following sequence: CF3(CH3)COH > CH2F(H)COH > CH2OH > H > CBr3 > CH2F > CCl3 > CF3 > CH2Br > CH2Cl > CH3 > C2H5 > C3H7 > i‐C4H9 > F > Br > OCH3 > Cl > CH2C6H5. The calculations at the UHSEH1PBE/6‐311++G(2d,2p) level with the most of substituents show nearly the same variation sequence for the J parameter. It is concluded that spin polarization effects in the diazaadamantane cage and a direct through‐space antiferromagnetic exchange interaction between the nitroxyl groups are the main mechanisms contributing to the exchange interaction parameter value in the studied series of compounds. The exchange coupling constant, J, depends on the electronic effects and geometry of the substituents, as well as on their specific interactions with the nitroxyl radical groups.  相似文献   

17.
Electrophilic trisubstituted ethylenes, dihalogen ring-substituted ethyl 2-cyano-1-oxo-3-phenyl-2-propenylcarbamates, RC6H3 CH = C(CN)CONHCO2C2H5(where R is 2,3-diCl, 2,4-diCl, 2,6-diCl, 3,4-diCl, 3,5-diCl, and 2-Cl-6-F, were prepared and copolymerized with styrene. The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of ring-substituted benzaldehydes and N-cyanoacetylurethane, and characterized by CHN analysis, IR, 1H- and 13C-NMR. All the ethylenes were copolymerized with styrene (M1) in solution with radical initiation (ABCN) at 70°C. The compositions of the copolymers were calculated from nitrogen analysis and the structures were analyzed by IR, 1H- and 13C-NMR. The order of relative reactivity (1/r 1) for the monomers 2,4-diCl (4.4) > 2,6-diCl (3.6) > 2,3-diCl (3.4) = 3,4-diCl (3.4) > 2-Cl-6-F (2.7) > 3,5-diCl (2.0). High T g of the copolymers in comparison with that of polystyrene indicates decrease in chain mobility of the copolymer due to the high dipolar character of the trisubstituted ethylene structural unit. Decomposition of the copolymers in nitrogen occurred in two steps, first in 270–420°C with residue (5–13% wt), which then decomposed in the 420–650°C range.  相似文献   

18.
Experimental redox potentials of 16 derivatives of tris(β-diketonato)iron(III) complexes (where β-diketonato(R1COCHCOR2), with substituents R1 and R2 in different combinations of H, C4H3S, C4H3O, CH3, Ph, CF3, or C (CH3)3), and 11 additional isomers, were studied theoretically in terms of the electronic properties, substituent effects, electron affinity, and molecular electrostatic potential (MESP) analysis, using density functional theory methods. The computational methods reproduced the experimental reduction potential to a very high level of accuracy, especially when the M062X functional was used (with mean absolute deviation [MAD] = 0.054 and 0.093 and correlation R2 = 0.978 and 0.981 obtained by application of two slightly different free energy cycles, respectively). The most negative computed reduction potential corresponds to the most negative reported experimental reductions, which is indicative of the least favorable reduction potential, also in most cases the most stable molecules energetically. The calculated reduction potentials of the fac isomers of the molecules were generally higher (less negative) than that of the mer isomers when one of the ligand substituents R1 or R2 was CF3 (M062X results), indicating better ease of reduction, although in many cases, the experimental reduction potential agreed better with the calculated reduction potential of the mer isomer instead. The calculated reduction potentials were also affected by the substituents in the order of CF3 > H > C4H3S > C4H3O > Ph > CH3 > C(CH3)3 (the most negative value). The stronger the electron withdrawing tendency of the substituent, the more favorable (less negative value) the reduction potential becomes. In relation to the CH3-substituted molecule 1 as a reference, the molecules with electron withdrawing substituents resulted in an electron-deficient MESP iso-surface, in both the neutral state and reduced state. All the molecules in their reduced state were characterized with an electron-deficient MESP iso-surface compared with the reduced CH3-substituted molecule 1, with the deficiency increasing in mer compared with fac, for both the neutral and reduced molecules. The relative MESP values of ΔVFe in the reduced state of the molecules were able to predict the corresponding experimental reduction potential to a significant level of accuracy (with MAD = 0.091).  相似文献   

19.
In this study, green phosphorescent Pt(II) complexes with N,N‐diphenyl‐6‐(1H‐pyrazol‐1‐yl)pyridin‐2‐amine (Ndpp) coordinated ligands, [Pt (Ndpp)Cl] 2a , [Pt (Ndpp)Pb, Pb = (prop‐1‐ynyl)benzene] 2b , and [Pt (Ndpp)CN] 2a? CN were theoretically investigated by means of density functional theory and time‐dependent density functional theory calculations to reveal their marked distinct phosphorescence quantum yields. These complexes exhibit evident absorption bands in the 200–450 nm region but emit strong green light with marked differences of phosphorescence quantum yields. Compared with the complex 2a , the complex 2b possesses large oscillator strengths of absorption spectra, strong spin‐orbit coupling, and transition electric dipole moment, as well as small singlet‐triplet splitting energies, which conduces to enhancing its radiative decay. To illustrate the nonradiative decay process, the transition state (TS) between the triplet metal‐centered (3MC) state and the excited state (T1) was optimized. The 3MC state is found to be the minimum energy crossing point (MECP) between the T1 state and the S0 state. Compared with the complex 2a , the complex 2b possesses a much larger energy barrier to the MECP state from the T1 state, so it is strongly emissive in the green region. Besides, the introduction of ? CN substitutions on 2a is useful for enhancing the energy barrier to the thermal deactivation pathway of 3MLCT → TS → MECP. These results demonstrate that the modification of metal–ligand conjugation is an effective way to develop high‐performance phosphorescent materials.  相似文献   

20.
Synthesis and Properties of (Acido)(nitrosyl)phthalocyaninato(2–)ruthenium (Acido)(nitrosyl)phthalocyaninato(2–)ruthenium, [Ru(X)(NO)pc2–] (X = F, Cl, Br, I, CN, NCO, NCS, NCSe, N3, NO2) is obtained by acidification of a solution of bis(tetra(n-butyl)ammonium) bis(nitro)phthalocyaninato(2–)ruthenate(II) in tetrahydrofurane with the corresponding conc. mineral acid or aqueous ammonium salt solution. The nitrite-nitrosyl conversion is reversal in basic media. The cyclic and differential pulse voltammograms show mainly three quasi-reversible one-electron processes at 1.05, –0.65 and –1.25 V, ascribed to the first ring oxidation and the stepwise reduction to the complexes of type {RuNO}7 and {RuNO}8, respectively. The B < Q < N regions in the electronic absorption spectra are still typical for the pc2– ligand, but are each split into two strong absorptions (14500/16500(B); 28000/30500(Q); 34500/37000 cm–1(N)), whose relative intensities strongly depend on the nature of the axial ligand X. In the IR spectra is active the N–O stretching vibration between 1827 (X = I) and 1856 cm–1 (F), the C–N stretching vibration at 2178 (X = NCO), 2072 (NCS), 2066 (NCSe), 2093 cm–1 (CN), the N–N stretching vibration of the azide ligand at 2045 cm–1, the fundamentals of the nitrito(O) ligand at 1501, 932, and 804 cm–1, and the Ru–X stretching vibration at 483 (F), 332 (Cl), 225 (Br), 183 (I), 395 (N3), 364 (ONO), 403 (CN), 263 (NCS), and 231 cm–1 (NCSe). In the resonance Raman spectra, excited in coincidence with the B region, the Ru–NO stretching vibration and the very intense Ru–N–O deformation vibration are selectively enhanced between 580 and 618 cm–1, and between 556 and 585 cm–1, respectively.  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