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 484 毫秒
1.
Nucleophilic substitution of Pd(RaaiR′)Cl2 [(RaaiR′ = 1-alkyl-2-(arylazo)imidazole, p-R-C6H4-N=N-C3H2NN-1-R′; where R = H(a)/ Me(b)/ Cl(c) and R′ = Et(1)/Bz(2)] with 2-Mercaptopyridine (2-SH-Py) in acetonitrile (MeCN) at 298 K, to form [Pd2(2-S-Py)4], has been studied spectrophotometrically under pseudo-first-order conditions and the analyses support the nucleophilic association path. The reaction follows the rate law, Rate = {k 0 + k [2-SH-Py] 0 2 }[Pd(RaaiR′)Cl2]: first order in Pd(RaaiR′)Cl2 and second order in 2-SH-Py. The rate of the reaction follows the order: Pd(RaaiEt)Cl2 (1) < Pd(RaaiBz)Cl2 (2) and Pd(MeaaiR′)Cl2 (b) < Pd(HaaiR′)Cl2 (a) < Pd(ClaaiR′)Cl2 (c). External addition of Cl (LiCl) and HCl suppresses the rate (Rate ∝ 1/[Cl]0 & ∝1/[HCl]0). The reactions have been studied at different temperatures (293–308 K) and activation parameters (Δ H° and Δ S°) of the reactions were calculated from the Eyring plot and support the proposed mechanism.  相似文献   

2.
Nucleophilic substitutions of Pd(N,N)Cl2[(N,N = 1-methyl-2-(arylazo)imidazole (RaaiMe), p-RC6H4N=NC3H2NN-1-Me; 2-(arylazo)pyridine (Raap), p-RC6H4N=NC5H4N; 2-(arylazo)pyrimidine (Raapm), p-RC6H4N=NC4H3N2 where R = H (a), Me (b), Cl (c)] with 8-quinolinol (HQ) have been examined by spectrophotometry at 298 K in MeCN solution. The product, Pd(Q)2, has also been confirmed by independent synthesis from Na2[PdCl4] and HQ in EtOH. The kinetics of the reaction have been studied under pseudo-first-order conditions and the analyses support a nucleophilic association path. A single phase reaction has been observed and follows the rate law, rate = a + k [Pd(N,N)Cl2] [HQ]2. Thus, the reaction is first order in [Pd(N,N)Cl2] and second order in [HQ]. External addition of Cl(LiCl) suppresses the rate. The rate increases as follows: Pd(RaaiMe)Cl2 < Pd(Raap)Cl2 < Pd(Raapm)Cl2.  相似文献   

3.
Picolinic acid (picH) reacts with [Pd(α-/β-NaiR)Cl2] [α-/β-NaiR = 1-alkyl-2-(naphthyl-α-/β-azo)imidazole] in acetonitrile (MeCN) medium to give [Pd(α-/β-NaiR)(pic)](ClO4). The products are characterized by spectroscopic techniques (FT-IR, UV–Vis, NMR). The reaction kinetics show first order dependence of rate on each of the concentration of Pd(II) complex and picH. Addition of LiCl to the reaction decreases the rate of reaction and has proved the cleavage of Pd–Cl bond at the rate-determining step. Thermodynamic parameters (Δ and Δ) are determined from variable temperature kinetic studies. The magnitude of the second order rate constant, k2 increases as in the order: Pd(NaiEt)Cl2 < Pd(NaiMe)Cl2 <  Pd(NaiBz)Cl2 as well as Pd(β-NaiR)Cl2 <  Pd(α-NaiR)Cl2.  相似文献   

4.
The reaction of K2[PdCl4] with [(S,S)-H2(Et)2eddv]Cl2 diester (O,O′-diethyl-(S,S)-ethylenediamine-N,N′-di-2-(3-methyl)butanoate) (1) resulted in [PdCl2{(S,S)-(Et)eddv-κ2 N,N′O}] (2) complex with one hydrolyzed ester group. The compound was characterized by spectroscopic methods and it was found that the reaction is diastereoselective (1H and 13C NMR; one diastereoisomer of four possible). In addition, the structure of 2 was confirmed by X-ray diffraction analysis, indicating that the product is the (R,R)–N,N′-configured isomer. DFT calculations support the formation of one diastereoisomer of 2.  相似文献   

5.
Interaction of adenine (A) with dichloro-[1-alkyl-2-(α-naphthylazo)imidazole] palladium(II) [Pd(α-NaiR)Cl2], 1 and dichloro-[1-alkyl-2-(β-naphthylazo)imidazole] palladium(II) [Pd(β-NaiR)Cl2], 2 {where R=Me (a), Et (b) or Bz (c)} in MeCN-water (50% v/v) medium to yield [{1-alkyl-2-(α-naphthylazo)imidazole}(adenine)]palladium(II) perchlorates (3a, 3b, 3c) and [{1-alkyl-2-(β-naphthylazo)imidazole}(adenine)]palladium(II) perchlorates (4a, 4b, 4c) was studied. The products were characterized by physico-chemical and spectroscopic methods. The reaction kinetics were second order overall, being first order in both the Pd(II) complex and adenine. The effect of adding chloride was consistent with rate-limiting dissociation of chloride from the complex. Thermodynamic parameters were determined from temperature variation experiments. The second-order rate constant k 2 corroborates with the experimental ΔH° values, while the negative values of ΔS° indicate that the reaction proceeds through an associative inner sphere mechanism.  相似文献   

6.
Reaction of 2-(phenylazo)pyridine (pap) with [Ru(PPh3)3X2] (X = Cl, Br) in dichloromethane solution affords [Ru(PPh3)2(pap)X2]. These diamagnetic complexes exhibit a weakdd transition and two intense MLCT transitions in the visible region. In dichloromethane solution they display a one-electron reduction of pap near − 0.90 V vs SCE and a reversible ruthenium(II)-ruthenium(III) oxidation near 0.70 V vs SCE. The [RuIII(PPh3)2(pap)Cl2]+ complex cation, generated by coulometric oxidation of [Ru(PPh3)2(pap)Cl2], shows two intense LMCT transitions in the visible region. It oxidizes N,N-dimethylaniline and [RuII(bpy)2Cl2] (bpy = 2,2′-bipyridine) to produce N,N,N′,N′-tetramethylbenzidine and [RuIII(bpy)2Cl2]+ respectively. Reaction of [Ru(PPh3)2(pap)X2] with Ag+ in ethanol produces [Ru(PPh3)2(pap)(EtOH)2]2+ which upon further reaction with L (L = pap, bpy, acetylacetonate ion(acac) and oxalate ion (ox2−)) gives complexes of type [Ru(PPh3)2(pap)(L)]n+ (n = 0, 1, 2). All these diamagnetic complexes show a weakdd transition and several intense MLCT transitions in the visible region. The ruthenium(II)-ruthenium(III) oxidation potential decreases in the order (of L): pap > bpy > acac > ox2−. Reductions of the coordinated pap and bpy are also observed.  相似文献   

7.
Reaction of [Ni(dppe)Cl2/Br2] with AgOTf in CH2Cl2 medium following ligand addition leads to [Ni(dppe)(OSO2CF3)2] and then [Ni(dppe)(RaaiR)](OSO2CF3)2 [RaaiR′ = p–R–C6H4–N=N–C3H2–NN-1–R′,(1–3), abbreviated as N,N′-chelator, where N(imidazole) and N(azo) represent N and N′, respectively; R = H (a), Me (b), Cl (c) and R′ = Me (1), CH2CH3 (2), CH2Ph (3), OSO2CF3 is the triflate anion]. 31P{1H}-NMR confirm that stable bis-chelated square planar Ni(II) azoimine–dppe complex formation with one sharp peaks. The 1H NMR spectral measurements suggest azoimine link is present with lot of phenyl protons in the aromatic region. Considering all the moities there are a lot of different carbon atoms in the molecule which gives many different peaks in the 13C(1H)-NMR spectrum. In the 1H-1H COSY spectrum in the present complexes and contour peaks in the 1H-13C-HMQC spectrum in the present complexes, assign the solution structure and stereoretentive conformation in each complexes.  相似文献   

8.
Nucleophilic substitution of Pd(RaaiR′)Cl2 [RaaiR′=1-alkyl-2-(arylazo)imidazole, p-R—C6H4— N=N—C3H2NN-1-R′; where R= H(a)/Me(b)/Cl(c) and R′ = Et(1)/Bz(2)] with adenine (A) in MeCN–water (1:1) at 298 K, to form [Pd(A)2]Cl2, has been studied spectrophotometrically under pseudo-first-order conditions and the analyses support a nucleophilic association path. The reaction follows the rate law, rate = {a+k [A] 02[Pd(RaaiR′)Cl2]: first-order in Pd(RaaiR′)Cl2 and second-order in A. The rate increases as follows: Pd(RaaiEt)Cl2(1) < Pd(RaaiBz)Cl2(2) and Pd(MeaaiR′)Cl2(b) < Pd(HaaiR′)Cl2(a) < Pd(ClaaiR′)Cl2(c). External addition of Cl (LiCl) suppresses the rate (rate 1/[Cl]). The activation parameters, H0 and S0 of the reactions were calculated from the Eyring plot and support the proposed mechanism.  相似文献   

9.
Reaction of [Pd(dppe)Cl2/Br2] with AgOTf in a dichloromethane medium followed by ligand addition led to [Pd(dppe)(OSO2CF3)2] and then [Pd(dppe)(RaaiR)](OSO2CF3)2 [RaaiR′ = p-R-C6H4-N=N-C3H2-NN-1-R′, (1–3), abbreviated as a N,N′-chelator, where N(imidazole) and N(azo) are represented by N and N′, respectively; R = H (a), Me (b), Cl (c) and R′ = Me (1), CH2CH3 (2), CH2Ph (3), OSO2CF3 is the triflate anion, dppe = 1,2-bis-(diphenylphosphinoethane)]. 31P “1H” NMR confirmed that due to the two phosphorus atom interaction in the azoimine symmetrical environment one sharp peak was formed. The 1H NMR spectral measurements suggest that azo-imine link with lot of phenyl protons in the aromatic region. 13C (1H) NMR spectrum, 1H, 1H COSY and 1H, 13C HMQC spectrum assign the solution structure and stereo-retentive conformation in each complex.  相似文献   

10.
Two new chromium(III) complexes with picolinamide (pica) and oxalates, [Cr(C2O4)2(N,N′-pica)]2− and [Cr(C2O4)2(N,O-pica)], were obtained and the kinetics of their aquation in HClO4 solutions were studied. The aquation leads to pica liberation and proceeds in two stages: (i) the chelate-ring opening at the Cr–amide bond and (ii) the Cr–N-pyridine bond breaking, which gives free pica and cis-[Cr(C2O4)2(H2O2)2]. In the case of N,N′-bonded pica the kinetics of both stages was determined and in the case of the N,O-bonded pica only the second stage was investigated. The following rate laws were established: (k obs)1 = k 0 + k 1 Q 1[H+] and (k obs)2 = k 2 Q 2[H+], where k 0 and k 1 are the rate constants of the chelate-ring opening in the unprotonated and protonated starting complex, and k 2 is the rate constant of the pica liberation from the protonated intermediate. Kinetic parameters are calculated and the aquation mechanism is discussed.  相似文献   

11.
Oxidation of N-methylethylamine by bis(hydrogenperiodato)argentate(III) ([Ag(HIO6)2]5−) in alkaline medium results in demethylation, giving rise to formaldehyde and ethylamine as the oxidation products. The oxidation kinetics has been followed spectrophotometrically in the temperature range of 20.0–35.0 °C, and shows an overall second-order character: being first-order with respect to both Ag(III) and N-methylethylamine. The observed second-order rate constants k′ increase with increasing [OH] of the reaction medium, but decrease with increasing the total concentration of periodate. An empirical rate expression for k′ has been derived as: k′ = (k a + k b[OH])K 1/{f([OH])[IO4 ]tot + K 1}, where k a and k b are rate parameters, and K 1 is an equilibrium constant. These parameters have been evaluated at all the temperatures studied, enabling calculation of activation parameters. A reaction mechanism is suggested to involve two pre-equilibria, leading to formation of an intermediate Ag(III) complex, namely [Ag(HIO6)(OH)(MeNHEt)]2−. In the subsequent rate-determining steps, this intermediate undergoes inner-sphere electron transfer from the coordinated amine to the metal center via two distinct routes, one of which is spontaneous while the other is mediated by a hydroxide ion.  相似文献   

12.
Palladium(II) complexes of thiones having the general formula [Pd(L)4]Cl2, where L = thiourea (Tu), methylthiourea (Metu), N,N′-dimethylthiourea (Dmtu), and tetramethylthiourea (Tmtu) were prepared by reacting K2[PdCl4] with the corresponding thiones. The complexes have been characterized by elemental analysis, IR and NMR spectroscopy, and two of these, [Pd(Dmtu)4]Cl2 · 2H2O (1) and [Pd(Tmtu)4]Cl2 (2), by X-ray crystallography. An upfield shift in the >C=S resonance of thiones in 13C NMR and downfield shift in N–H resonance in 1H NMR are consistent in showing sulfur coordination with palladium(II). The crystal structures of the complexes show a square-planar coordination environment around the Pd(II) ions with the average cis and trans S–Pd–S bond angles of 89.64° and 173.48°, respectively. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users. An erratum to this article can be found at  相似文献   

13.
Summary Kinetics of formation of [PdCl4]2– from [Pd(ox)2]2– and [Pd(mal)2]2– has been studies in aqueous acid media in the presence of an excess of chloride ion by stopped-flow spectrophotometry. Both the complexes undergo the transformation in two well separated consecutive steps. In 0.02–0.05 M acid with 0.2 M Cl, Pd(AA)2– dissociates leading to the formation of [Pd(AA)Cl2]2– (where AA =ox2– or mal2–), which in 0.1–0.6 M acid and 1 M Cl forms [PdCl4]2– in a relatively slow step. For both steps kabs=k0+k2[H+][Cl]. Activation parameters corresponding to k0 and k2 have been determined. Results indicate that [Pd(mal)2]2– is much more labile to substitution than [Pd(ox)2]2– and for both the lability is far greater than that of [Pd(bigH)2]2+ and [Pt(ox)2]2– reported earlier.  相似文献   

14.
Summary Kinetic and mechanistic studies on the anation of cis-[Ru(tap)2(H2O)2]2+ (where tap = 2-(m-tolylazo)pyridine) by pyridine-2-aldoxime (LH) have been made spectrophotometrically at different temperatures (35–50° C) in aqueous medium and various EtOH-H2O mixtures. The following rate law has been established at pH 5.6: k obs = k 1 k 2[LH]/(k −1 + k 2[LH]) where k 1 is the H2O dissociation rate constant of the substrate complex and k -1 and k 2 are the aquation and ligand capturing rate constants of the pentacoordinate intermediate, [Ru(tap)2(H2O)]2+. The rate constants are independent of ionic strength. The reaction rate increases with increasing pH. Activation parameters (ΔH‡ and ΔS‡) have been calculated for media of four different dielectric constants and compared with other substitution reactions in aqueous medium. A dissociative mechanism is proposed.  相似文献   

15.
The compounds, 2,6-bis(3,5-dimethylpyrazol-1-ylmethyl)pyridine (MeNˆNˆN) (L1) and 2,6-bis(3,5-ditertbutylpyrazol-1-ylmethyl)pyridine (tBuNˆNˆN) (L2), react with either [Pd(NCMe)2Cl2] or [Pd(COD)ClMe] to form the mononuclear palladium complexes [Pd(MeNˆNˆN)Cl2] (1), [Pd(MeNˆNˆN)ClMe] (2), [Pd(tBuNˆNˆN)Cl2] (3) and [Pd(tBuNˆNˆN)ClMe] (4). Reactions of 1, 2 and 4 with the halide abstractor, NaBAr4 (Ar = 3,5-(CF3)2C6H3), led to the formation of stable tridentate cationic species [Pd(MeNˆNˆN)Cl]+(5), [Pd(MeNˆNˆN)Me]+ (6) and [Pd(tBuNˆNˆN)Cl]+ (7) respectively. The analogous carbonyl linker cationic species [Pd{(3,5-Me2pz-CO)2-py}Cl]+ (9) and [Pd{(3,5-tBu2pz-CO)2-py}Cl]+ (10), prepared by halide abstraction of the neutral complexes [Pd{(3,5-Me2pz-CO)2-py}Cl2] and [Pd{(3,5-tBu2pz-CO)2-py}Cl2] by NaBAr4, were however less stable with t1/2 of 14 and 2 days respectively. Attempts to crystallize 1 and 3 from the mother liquor resulted in the isolation of the salts [Pd(MeNˆNˆN)Cl]2[Pd2Cl6] (11) and [Pd(tBuNˆNˆN)Cl]2[Pd2Cl6] (12). Although when complexes 14 were reacted with modified methylaluminoxane (MMAO) or NaBAr4, no active catalysts for ethylene oligomerization or polymerization were formed, activation with silver triflate (AgOTf) produced active catalysts that oligomerized and polymerized phenylacetylene to a mixture of cis-transoidal and trans-cisoidal polyphenylacetylene.  相似文献   

16.
 The polymeric compound [Ru(cod)Cl2] x (cod = cyclooctadiene) reacts with 2 equivalents of tmeda (N,N,N′,N′-tetramethylethylenediamine) in refluxing MeOH to afford trans-[Ru(cod)(tmeda)(Cl)(H)] (1), which upon treatment with CHCl3 is readily converted to the dichloro complex trans-[Ru(cod)(tmeda)Cl2] (2). When [Ru(cod)Cl2] x is reacted with tmeda under an atmosphere of H2 (3 bar), the bis-tmeda complex trans-[Ru(tmeda)2Cl2] (3) is obtained in 80% yield. DFT calculations revealed that 3 is by 52 kJ/mol more stable than the corresponding cis isomer. Attempts to prepare the coordinatively unsaturated complex [Ru(tmeda)2Cl]+ by reacting 1 with TICF3SO3 were unsuccessful. According to DFT calculations, however, such a complex should be stable and, interestingly, should adopt a square pyramidal rather than a trigonal bipyramidal structure. If halide abstraction of 3 is performed in the presence of terminal alkynes HC*CR (R*t-Bu, n-Bu), the cationic vinylidene complexes [Ru(tmeda)2(Cl)(*C*CHR)]+ (4a,b) are obtained.  相似文献   

17.
The reaction of dichloro{1-methyl-2-(arylazo)imidazole}palladium(II), Pd(RaaiMe)Cl2 where RaaiMe = p-R–C6H4N=N–C3H2N2-1-Me; R = H(1), Me(2), Cl(3), with pyridine bases [RPY: R = H (a), 4-Me (b), 4-Cl (c), 2-Me (d), 2,6-Me2 (e), 2,4,6-Me3 (f)] has been studied spectrophotometrically in MeCN at 451 nm. The products (4) have been isolated and characterised as trans-Pd(RPy)2Cl2. The kinetics of the nucleophilic substitution has been examined under pseudo-first-order conditions at 298 K. A single phase reaction step has been observed for bases such as Hpy (a), 4-MePy (b) and 4-ClPy (c) and follows the rate law: rate = (a + k[RPy]2[Pd(RaaiMe)Cl2]). The bases 2-MePy (d), 2,6-Me2Py (e) and 2,4,6-Me3Py (f) exhibits a bi-phasic reaction and follows the rate laws: rate–1 = (a + k[RPy][Pd(RaaiMe)Cl2]) and rate–2 = (a + k[RPy][Pd(RaaiMe)-Cl2]), where k is the third-order rate constant; k is the second-order first phase rate constant, k is the second-order second phase rate constant and a/a/a correspond to the solvent dependent constant of the respective reaction path. The rate data supports a nucleophilic association path. External addition of Cl (LiCl) suppresses the rate, which follows the order: k/k/k (3) > k/k,k (1) > k/k,k (2). The k values are linearly related to the Hammett constants. The 2-substituted pyridines (d–f) remarkably reduce the rate and show a bi-phasic reaction behaviour as compared with 4-Rpy (a–c). This is attributed to the steric effect that destabilises the transition state. The rate decreases with increasing steric crowding at the ortho-position and follows the order: (d) > (f) > (e). The 4-substituted pyridines control the rate via an inductive effect and follow the order: (b) > (a) > (c).  相似文献   

18.
2-(Arylazo)pyrimidines (aapm) are N,N′-chelating ligands and synthesise orange-red complexes of composition [Pd(aapm)Cl2],1, with Pd(MeCN)2Cl2 in MeCN. The complex hascis-PdCl2 configuration [v(Pd-Cl): 340, 360 cm−1]. The treatment of Tollen’s reagent (‘AgOH’) leads to chelatative hydroxylation in the pendant aryl ring, affording a green phenolato complex, Pd(aapmO)Cl,5 (aapmO is deprotonated 2-((8-hydroxo)arylazo)pyrimidine). The reaction is also carried out by controlled addition of dilute sodium hydroxide in air or by the addition of PhIO/m-chloroperbenzoic acid to a MeCN suspension of the complex. A single Pd-Cl stretch at 360 cm−1 supports the composition of phenolato complex. Unlike Pd(aapm)Cl2 the hydroxylated product, Pd(aapmO)Cl, has a structured intense absorption in the visible region near 670 nm. The Pd-Cl bond in Pd(aapmO)Cl is highly sensitive to nucleophilic substitution and slowly hydrolyses in aqueous medium. For Parts I and II, see references 16 and 17; for Part III,Synth. React. Inorg. Met.- Org. Chem. (in press)  相似文献   

19.
The kinetics of the reductive stripping of plutonium(IV) by dihydroxyurea (DHU) in 30% TBP/kerosene-HNO3 system was studied with a constant interfacial area cell. The stripping rate of plutonium(IV) increases with the increase of the stirring speed of two phases and the interfacial area. The activation energy of this process is 28.4 kJ/mol. Under the given experimental conditions, the mass transfer of Pu is not controlled by redox reaction, but controlled by molecular diffusion from the organic phase to organic film layer and from the aqueous film layer to aqueous phase. The rate equation of reductive stripping (process is controlled by diffusion) was obtained as: r 0 = k′[Pu(IV)]0[DHU]a 0.16[HNO3]a −0.34. The rate constant k′ is (5.0±0.4)·10−2 (mol/L)0.18·min−1 at 18.0°C.  相似文献   

20.
Reaction of [Au(C6F5)(tht)2Cl](OTf) with RaaiR′ in CH2Cl2 medium leads to [Au(C6F5)(RaaiR′)Cl](OTf) [RaaiR′ = p-R–C6H4–N=N–C3H2–NN-1-R′, (1–3), abbreviated as N,N′-chelator, where N(imidazole) and N(azo) represent N and N′, respectively; R = H (a), Me (b), Cl (c) and R′ = Me (1), CH2CH3 (2), CH2Ph (3), tht is tetrahydrothiophen]. The maximum molecular peak of [Au(C6F5)(MeaaiMe)Cl] is observed at m/z 599.51 (100 %) in the FAB mass spectrum. Ir spectra of the complexes show –C=N– and –N=N– stretching near at 1590 and 1370 cm−1 and near at 1510, 955, 800 cm−1 due to the presence of pentafluorophenyl ring. The 1H-NMR spectral measurements suggest methylene, –CH2–, in RaaiEt gives a complex AB type multiplet while in RaaiCH2Ph shows AB type quartets. 13C-NMR spectrum of complexes confirm the molecular skeleton. In the 1H-1H-COSY spectrum as well as contour peaks in the 1H-13C HMQC spectrum for the present complexes, assign the solution structure and stereoretentive conformation. The electrochemistry gives the ligand reduction peaks.  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