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 31 毫秒
1.
The direct potentiometric determination of the normal potentials of Cu/Cu+, Cu+/Cu2+, Ag/Ag+, Hg/Hg2+2 couples in nitromethane and the comparison of these values with the normal potentials in aqueous solution makes possible the calculation of solvation coefficients γ(t)(Mn+) of the corresponding cations; the Strehlow hypothesis (γ(Fc) = γ(Fc+)) has been used. An indirect method using the determination of the solubility constants of alkaline chlorides and alkaline earth chlorides leads to the normal potentials of alkaline and Mg2+ cations and thus to the γ(t) values of these ions.The measures of the AgX-2 stability constants and of the AgX solubility constants lead to the γ(t) values of X? and AgX?2 anions (X? = Cl?, Br?, I?, Scn?, CH3CO?2, CN?, BrO?3, (C6H5)4B?). A comparison of γ(t)(H+) and γ(t)(Ag+) for 20 solvents is given.  相似文献   

2.
We examine the impact of the glass network-modifier cation field strength (CFS) on ion irradiation-induced mechanical property changes in borosilicate (BS) glasses for the ternary M2O–B2O3–SiO2 systems with M = {Na, K, Rb} and the quaternary [0.5M(2)O–0.5Na2O]–B2O3–SiO2 systems with M = {Li, Na, K, Rb Mg, Ca, Sr, Ba}. 11B nuclear magnetic resonance (NMR) experiments on the as-prepared BS glasses yielded the fractional population of four-coordinated B species (B[4]) out of all {B[3], B[4]} groups in the glass network, along with the fraction of B[4]–O–Si linkages out of all B[4]–O–Si/B bonds. Both parameters correlated linearly with the (average) CFS of the M+ and/or {M(2)+, Na+} cations. Both the nanoindentation-derived hardness and Young's modulus values of the glasses reduced upon their irradiation by Si2+ ions, with the property deterioration decreasing linearly with increasing Mz+ CFS, that is, for higher Mz+⋅⋅⋅O interaction strength. The irradiation damage of the glass network also increased linearly with the fraction of B[4]–O–Si linkages, which are the second weakest in the structure after the Mz+⋅⋅⋅O bonds. Our results underscore the advantages of employing BS glasses with high-CFS cations for enhancing the radiation resistance for nuclear waste storage.  相似文献   

3.
The few clusters [B?nA+n+1]+ (n = 0,1) with resolvable mobilities formed in electrosprays of large salts have been used for nanoparticle instrument testing and calibration at sizes smaller than 2 nm. Extensions of this modest size range by charge reduction with uncontrolled gas phase ions has resulted in impure singly charged clusters. Here, we combine two oppositely charged electrosprays of solutions of the same salt B?A+, including: (CnH2n+1)4N+Br? (n = 4,7,12,16), the large phosphonium cation (C6H13)3(C16H33)P+ paired with the anions Im? [(CF3SO2)2N?] or FAP? [(C2F5)3PF3?], and the asymmetric pair [1-methyl-3-pentylimidazolium+FAP?]. Both polarities are simultaneously produced by this source in comparable abundances, primarily as singly charged A+nB?n±1, with tiny contributions from higher charge states. Some but not all of these clusters produce narrow mobility peaks typical of pure ions, even beyond n = 43. Excellent independent stable control of the positive and the negative sprays brought very close to each other is achieved by isolating them electrostatically with a symmetrically interposed metallic screen. Two nanoDMAs covering the size range up to 30 nm (Halfmini and Herrmann DMAs, with classification lengths of 2 and 10 cm) are characterized with these standards, revealing resolving powers considerably higher than previously seen with unipolar electrospray sources. The bipolar source of pure and chemically homogeneous clusters described permits studying size and charge effects in a variety of aerosol instruments in the 1–4 nm size range.

Copyright © 2017 American Association for Aerosol Research  相似文献   

4.
The interactions of ions with molecules and the determination of their dissociation patterns are challenging endeavors of fundamental importance for theoretical and experimental science. In particular, the investigations on bond-breaking and new bond-forming processes triggered by the ionic impact may shed light on the stellar wind interaction with interstellar media, ionic beam irradiations of the living cells, ion-track nanotechnology, radiation hardness analysis of materials, and focused ion beam etching, deposition, and lithography. Due to its vital role in the natural environment, the pyridine molecule has become the subject of both basic and applied research in recent years. Therefore, dissociation of the gas phase pyridine (C5H5N) into neutral excited atomic and molecular fragments following protons (H+) and dihydrogen cations (H2+) impact has been investigated experimentally in the 5–1000 eV energy range. The collision-induced emission spectroscopy has been exploited to detect luminescence in the wavelength range from 190 to 520 nm at the different kinetic energies of both cations. High-resolution optical fragmentation spectra reveal emission bands due to the CH(A2Δ→X2Πr; B2Σ+→X2Πr; C2Σ+→X2Πr) and CN(B2Σ+→X2Σ+) transitions as well as atomic H and C lines. Their spectral line shapes and qualitative band intensities are examined in detail. The analysis shows that the H2+ irradiation enhances pyridine ring fragmentation and creates various fragments more pronounced than H+ cations. The plausible collisional processes and fragmentation pathways leading to the identified products are discussed and compared with the latest results obtained in cation-induced fragmentation of pyridine.  相似文献   

5.
Binary blend films from lactide-rich poly(D -lactide-co-glycolide) (PDLG) and poly(L -lactide-co-glycolide) (PLLG) were obtained by casting methylene chloride solutions of the two mixed copolymers with different D- and L-lactide contents (XDI and XLI ), and their crystallization was studied by differential scanning calorimetry (DSC). Four combinations were selected from the binary (A-B) blends: mixing of the same polymer [XDI (A) = XDI (B) or XLI (A) = XLI(B)], blending under XDI (B) = XLI (A), blending of a D -lactide homopolymer [XDI(B) = 1] with other PDLGs, and blending of a D -lactide homopolymer [XDI(B) = 1] with other PLLGs. Racemic crystallites were exclusively formed between PDLG and PLLG when they had high lactide unit contents. The melting point and enthalpy of fusion of the racemic crystallites decreased with a decrease in XDI of PDLG or XLI of PLLG, suggesting that glycolide units in the polymer disturbed the growth of the racemic crystallites. A similar behavior was also observed for the homocrystallization in nonblended copolymer films. Homocrystallites composed entirely either of D -lactide unit or L-lactide unit sequences were formed when one component was crystallizable and the other component had the same sign of optical rotation or very different lactide content. An interesting finding was that even nonhomocrystallizable lactide-poor PDLG and PLLG could form racemic crystallites when both were blended. © 1994 John Wiley & Sons, Inc.  相似文献   

6.
The crystal structure of Kuzel's salt has been successfully determined by synchrotron powder diffraction. It crystallizes in the rhombohedral R3? symmetry with a = 5.7508 (2) Å, c = 50.418 (3) Å, V = 1444.04 (11) Å3. Joint Rietveld refinement was realized using three X-ray powder patterns recorded with a unique wavelength and three different sample-to-detector distances. Kuzel's salt is the chloro-sulfoaluminate AFm phase and belongs to the layered double hydroxide (LDH) large family. Its structure is composed of positively charged main layer [Ca2Al(OH)6]+ and negatively charged interlayer [Cl0.50·(SO4)0.25·2.5H2O]. Chloride and sulfate anions are ordered into two independent crystallographic sites and fill successive interlayer leading to the formation of a second-stage compound. The two kinds of interlayer have the compositions [Cl·2H2O] and [(SO4)0.5·3H2O]. The crystal structure explains why chloride and sulfate anions are not substituted and why the formation of extended solid solution in the chloro-sulfate AFm system does not occur.  相似文献   

7.
The total cross section σB(E) for the Ca(1D) + HBr → CaBr(B2Σ+) + H reaction as a function of collision energy has been measured using crossed atomic and molecular beams. The maxima exhibited by σB(E) with increasing energy are attributed to the opening of successive bending vibrational reaction channels that proceed via a [Ca(1D) ⃛ Br ⃛ H] transition state. A dynamical model for the reaction may be constructed in terms of Landau-Zener probabilities for curve crossing at two locations on the reaction path, coupled with a preference for consumption of transition-state vibrational energy.  相似文献   

8.
By using CASSCF (for optimization of geometries) and MR-SDGI-CASSCF (for energies) methods we have studied and compared the mechanism of reaction MCH2+ + H2, as well as the electronic and geometrical structure of the MCH2+ complex, where M = Co, Rh, and Ir. It has been found that the mechanisms of reaction MCH2+ + H2 → M+ + CH4 (1) for M = Co and Rh are similar and follow the path: MCH2++H2 → (H2)MCH2+ → [TS1, H2-activation] → MCH4+ → M+ + CH4. The key step is activation of the H-H bond, which has a barrier about twice as high for M = Co as for M = Rh; reaction ( 1 ) occurs more easily for M = Rh than M = Co. M = Ir completely changes the mechanism of reaction ( 1 ), which now follows the path: IrCH2+(3A2) + H2 → (H2) IrCH2+(3A2) → [TS1, H2-activation] → (H)2 IrCH2+(1A') → [TS2, H-migration] → HIrCH3+(3A) → [TS3, CH4-elimination] → IrCH4+(3A2) → Ir+(5F, s1d7) + CH4. The reaction ( 1 ) is exothermic for M = Co and Rh, but endothermic for M = Ir. For M = Co and Rh, the reverse reaction M+ + CH4 can give only one product MCH4+ and does not proceed further easily; for M = Co, at elevated temperature CoCH4+may give CoH+ and CoCH3+. However, for M = Ir the reverse reaction can proceed further to give hydridomethyl HIrCH3+ and bishydrido (H)2CH2+ complexes, as well as IrCH4+.  相似文献   

9.
A theoretical analysis has been carried out of the structure of metalliferous epoxy–chelate polymers (MECP) based on diglycidyl ether of bisphenol-A (DGEBA) hardened with metal complexes of the formula [M(L)n(X)p], where M is the cation of the transition metal; R, a nitrogen-containing ligand; X, the anion of an organic acid; n, the number of the ligands in the complex molecule (n = 1 or 2), and p, the metal valency (p = 2 or 3). On the basis of the correlations between the tensile strength (σt) and tensile modulus (Et), and flexural strength (σf) and flexural modulus (Ef), of MECP, σt = f(Et)and σf = f(Ef), and supposing that when the condition $ \sigma _{_{t_A } }= \sigma _{_{t_B } } ,{\rm}\sigma _{_{f_A } }= \sigma _{f_B } ,{\rm}E_{t_A }= E_{t_B } ,{\rm}E_{f_A }= E_{f_B } $ is fulfilled, where A and B are complex hardeners of different structures but of the same class, the epoxy–chelate matrices have similar structures. The influence of the structural fragments of the hardener molecule (the metal, ligand, and anion) on the polymer properties was evaluated and it was found out that the biggest contribution to these properties belongs to the metal, the alteration of which changes the thermal stability (ΔM is the polymer mass loss after thermal treatment in air), deformability (ε), σf, Ef, and deflection temperature (DT) significantly. By this, the effect of the hardener structure change on the alteration of the MECP properties is maximal for ΔM, is minimal for the compressive strength (σc), and decreases in the series: ΔM > ε > DT > σf > Ef > σc. The type of the anion affects σc significantly, but the ligand type contributes the least to the polymer properties. The obtained dependencies of the MECP properties on the structural fragments of the complex hardeners allow preliminary evaluation of the structure of the chelates and epoxy–chelate compositions necessary to produce epoxy polymers with required properties. The new method of the theoretical investigation of the effect of the structural fragments (method of TIESF) of the polymer matrix on the polymer properties can be used to analyze the structures of the polymers of other classes and to predict the optimal structures, promising the production of the materials with the optimal properties. © 1993 John Wiley & Sons, Inc.  相似文献   

10.
Effects of the mixing ratio of poly(DL-lactide) (PDLLA) and poly (ε-caprolactone) (PCL) on the thermal and mechanical properties and morphologies of the solution-cast blends were investigated by differential scanning calorimetry (DSC), polarizing microscopy, tensile tests, and dynamic mechanical analysis. The presence of amorphous PDLLA did not disturb crystallization of PCL over the PDLLA content [XPDLLA = PDLLA/(PCL + PDLLA)] from 0.1 to 0.9 and allowed PCL to form spherulites over XPDLLA ranging from 0.1 to 0.6. The spherulite radius was larger for the blends than for the nonblended PCL. Phase separation occurred for the blends with XPDLLA between 0.1 and 0.9 Tm of PCL remained unchanged in the XPDLLA range up to 0.6 but decreased at XPDLLA above 0.6, whereas the crystallinity of PCL was constant around 60%, irrespective of XPDLLA. The tensile strength (σB, the yield stress (σY), the Young's modulus(E), and the storage modulus (G′) of the blends increased monotonously with XPDLLA if σB at XPDLLA = 0 and 0.6 and σY at XPDLLA = 0.6 were excluded. Elongation-at-break (εB) of PDLLA increased dramatically, while εB of PCL decreased remarkably when a small amount of the other component was added. Equations and parameters predicting σY, E, and G′ of the PCL-PDLLA blends were proposed as a function of XPDLLA. © 1996 John Wiley & Sons, Inc.  相似文献   

11.
Interaction in Crystals. 98. Protonated Hexamethylmelamin Salts with Different Anions: Monomeric Tetraphenylborate, Dimeric Trifluoracetate and Polymeric Chloride-Dihydrate Hexamethylmelamin (2,4,6-tris(dimethylamino)-1,3,5-triazine), on monoprotonation at one of the triazine nitrogens keeps its close to planar skeleton. Its salts with different anions reflect biochemically interesting hydrogen-bridge networks: Crystallization of the hydrochloride from isopropanol solution containing Li [B(C6H5)4], yields “naked” molecular cations, packed in herringbone fashion in between the lattice-dominating bulky, phenyl-shielded and non-protonable tetraphenylborate anions. From trifluoroacetic acid, crystals with sandwich-like subunits connected by hydrogen-bridged anions [F3CCOO…︁HOOCCF3] are obtained. The hydrochloride salt, prepared by adding aqueous HCl to a diethylether solution, crystallizes in stacks of triazinium cation dimers with an intermolecular bridge N-H…︁Cl…︁HN in between chloride-bydrate strands (…︁Cl…︁HOH…︁O(H)H…︁Cl…︁). Together, the structures of the three different hexamethyl-melaminium salts further illustrate the influence of anions on the crystallization of protonated nitrogen heterocycles and complement that of counter cation cation solvation in molecular anion salts.  相似文献   

12.
The synthesis of a new family of single‐ion conducting random copolymers bearing polyhedral boron anions is reported. For this purpose two novel ionic monomers, namely [B12H11(OCH2CH2)2OC(?O)C(CH3)?CH2]2?[(C4H9)4N+]2 and [8‐(OCH2CH2)2OC(?O)C(CH3)?CH2‐3,3′‐Co(1,2‐C2B9H10)(1′,2′‐C2B9H11)]?K+, having methacrylate function, diethylene glycol bridge and closo‐dodecaborate or cobalt bis(1,2‐dicarbollide) anions were designed. Such monomers differ from previously reported ones by (i) chemically attached highly delocalized boron anions, by (ii) valency of the anion (divalent anion and monovalent one) and by (iii) the presence of oxyethylene flexible spacer between the methacrylate group and bonded anion. Their free radical copolymerization with poly(ethylene glycol) methyl ether methacrylate and subsequent ion exchange provided lithium‐ion conducting polyelectrolytes showing low glass transition temperature (?53 to ?49 °C), ionic conductivity up to 9.1 × 10?7 S cm?1, lithium transference number up to 0.61 (70 °C) and electrochemical stability up to 4.1 V versus Li+/Li (70 °C). The incorporation of propylene carbonate (20–40 wt%) into the copolymers resulted in the enhancement of their ionic conductivity by one order of magnitude and significantly increased their electrochemical stability up to 4.7 V versus Li+/Li (70 °C). © 2019 Society of Chemical Industry  相似文献   

13.
《Ceramics International》2022,48(13):18094-18107
The impact of the cation field strength (CFS) of the glass network-modifier cations on the structure and properties of borosilicate glasses (BS) were examined for a large ensemble of mixed-cation (R/2)M(2)O–(R/2)Na2O–B2O3KSiO2 glasses with M+ ={Li+, Na+, K+, Rb+} and M2+ ={Mg2+, Ca2+, Sr2+, Ba2+} from four series of {K, R} combinations of K = n(SiO2)/n(B2O3) = {2.0, 4.0} and R =[n(M(2)O) ?+ ?n(Na2O)]/n(B2O3) = {0.75, 2.1}. Combined with results from La3+ bearing glasses enabled the probing of physical-property variations across a wide CFS range, encompassing the glass transition temperature (Tg), density, molar volume and compactness, as well as the hardness (H) and Young's modulus (E). We discuss the inferred composition–structure/CFS–property relationships. Each of Tg, H, and E revealed a non-linear dependence against the CFS and a strong Tg/H correlation, where each property is maximized for the largest alkaline-earth metal cations, i.e., Sr2+ and Ba2+, along with the high-CFS La3+ species. The 11B MAS NMR-derived fractional BO4 populations decreased linearly with the average Mz+/Na+ CFS within both K–0.75 glass branches, whereas the NBO-rich K–2.1 glasses manifested more complex trends. Comparisons with results from RM2O–B2O3KSiO2 glasses suggested no significant “mixed alkali effect”.  相似文献   

14.
Interactions in Crystals. 96. Preparation and Structures of Salts [RnNH…︁NRn][B(C6H5)4] with Prototype Hydrogen Bridges NH…︁N Straightforward crystallization of ammonium salts [RnNH] X added lithium tetraphenylborate, and amine RnN from acetone solution yields salts [RnNH…︁NRn][B(C6H5)4] with the (under the conditions) unprotonated anions and the cations with prototype hydrogen bridges NH…︁N. The structures of both identically substituted RnN (methylamine, trimethylamine, quinuclidine, diazabicyclooctane, and pyridine) as well as two-component cation species (quinuclidine…︁pyridine and diazabicyclooctane…︁pyridine) are reported and discussed. A Cambridge Structural Database search defines the area of charged NH…︁N interactions which can be correlated with both pK values and PM3 formation enthalpies. Additional information is provided by PM3 calculations based on the experimental structure coordinates. The charge distribution within the hydrogen bridges NH…︁N varies considerably with the individual proton donors N+H and proton acceptors N: Positive charges are highest at protonated quinuclidine and diazabicyclooctane centers and lowest at pyridine N acceptor centers.  相似文献   

15.
The protonation equilibria of weak bases (B) in solid acids (HClO4/SiO2, CF3SO3H/SiO2, H2SO4/SiO2) were studied by UV spectroscopy and the results were compared to those obtained for analogous compounds in concentrated aqueous solutions of strong acids (HClO4, CF3SO3H, H2SO4). The behaviour of B in liquid (L) and solid (S) phase was analysed by titration curves, log[BH+]/[B] ratios and thermodynamic pK BH+ values. It has been shown that the proton transfer process acid → base (i.e., from (H+A-)(L,S) to (BH+A-)(L,S)} can be described by the relationship observed between the activity coefficient terms that are to be taken into account for acid–base equilibria occurring in nonideal systems ( – log(f B f B+/f BH+)(L,S)= -n BA log(f Af H+/f HA)(L,S)) and can be estimated by the n BA values. Two “activity coefficient functions” (i.e., Mc(B) = – log(f B f B+/f BH+)and Mc(s) = – log(f Af H+/f HA)) were used to describe, respectively, the equilibria of B and the equilibria of the acids in concentrated aqueous solutions and the meaning of terms “activity coefficient function” and “protonating ability of an acid” were discussed. The difference between “acidity functions”, determined for solutes (Ac(i)) and solvents (Ac(s)) in aqueous acids, and the Hx acidity functions, the latter developed for solutes in analogous media by the Hammett procedure, was also shown. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

16.
From the radial stress σR and the normal stress σA, measured continuously during uniaxial loading and unloading on three compactable (sodium chloride, polyethylene and tartaric acid) and two non-compactable (polypropylene and polystyrene plastics) materials, characteristic compaction profiles of (σA ? σR) versus (σA + σR) can be observed. The uniaxial loading stress pathways for both compactable and non-compactable materials validated the assumption that the Coulomb yield criterion, which is usually applicable for the shear testing of soils, can be applied to the uniaxial compression of particulate materials. In addition, the unloading stress profiles for the compactable materials produced two characteristic parameters: a normal stress value at zero shear (σA + σR)o and a minimum shear stress value (σA ? σR)min. Correlation of (σA + σR)o and (σA ? σR)min values with either the tensile strength fc or the Vickers hardness number HV from the resultant compacts showed a linear logarithmic relationship. No such relationship was found, however, with non-compactable materials.  相似文献   

17.
The title compound, s-fac-[Co(dien)2][B7O9(OH)6]·9H2O (1) (dien = HN(CH2CH2NH2)2), has been prepared as a crystalline solid in moderate yield (35%) from the reaction of B(OH)3 with [Co(dien)2][OH]3 in aqueous solution (10:1 ratio). The structure contains a novel polyborate anion [B7O9(OH)6]3  which is structurally based on the known ‘ribbon’ isomer of [B7O9(OH)5]2 , with an additional [OH] group coordinated to a B atom in one of the outer boroxole rings. Compound 1 is formed by a self-assembly process in which the cation and anion mutually template themselves from equilibrium mixtures under reaction conditions. The [B7O9(OH)6]3  anions are H-bonded to each other in layers with ‘cavities’ suitable for the [Co(dien]2]3 + complex. Three [B7O9(OH)6]3  anions are in the secondary coordination sphere (via H-bonds) of each cation, with each anion H-bonded to three cations.  相似文献   

18.
A borate compound [Ni(en)3][B5O6(OH)4][CH3COO] (1) has been synthesized under mild hydrothermal conditions. The newly synthesized compound was characterized by single-crystal X-ray diffraction, elemental analysis and thermogravimetry. The three-dimensional (3D) supramolecular open-framework of compound 1 is constructed by inorganic [B5O6(OH)4] building units and organic [CH3COO] anions through hydrogen bonds, and the charge-balancing [Ni(en)3]2 + cations are located in 3D channels. It is very interesting that there exist two helical chains formed by [B5O6(OH)4] building units and organic [CH3COO] anions through hydrogen bonds in compound 1 along different axes. Magnetic measurement indicates that compound 1 exhibits paramagnetic behavior down to 4 K.  相似文献   

19.
Ion exchange equilibria of alkali metal ions (Li+, Na+, K+,, Rb+, and Cs+)H+, systems have been studied in MNO-j-HNOj media with ionic strength of 0.1 at 30, 45 and 60 °C on tin(IV) antimonate as a cation exchanger. The ion exchange isotherms have been measured for both forward and backward reactions by the batch technique. The isotherms showed S-shaped curves for all exchange systems studied. The selectivity coefficients (logarithmic scale) vary with the equivalent fraction XM of alkali metal ions in the exchanger and give two linear functions of XM with a break point (XM= 0.14, except 0.04 for Li+, /H+) indicating two different exchanging sites. The selectivity sequence, Na+, ? K+, ? Rb+, ? Cs+, ? ? Li, holds in the range of Xu= (0 - 0.04) and the sequence, Cs < Rb +, ? K +, ? Na +, < Li +, applies when XM is higher than 0.14.

Hypothetical thermodynamic data on “zero loading” of the ion exchange reaction was evaluated.  相似文献   

20.
The electrochemical behaviour of the series of ten [Rh(RCOCHCOR′)(P(OPh)3)2] complexes with R, R′ = CF3, CF3 (1), CF3, CH3 (2), CF3, Ph (C6H5) (3), CF3, Fc (ferrocenyl = (C5H5)Fe(C5H4)) (4), CH3, Ph (5), CH3, CH3 (6), Ph, Ph (7), Fc, CH3 (8), Fc, Ph (9) and Fc, Fc (10) were studied in acetonitrile containing 0.100 mol dm−3 tetra-n-butylammonium hexafluorophosphate as supporting electrolyte utilizing a glassy carbon working electrode. Results are consistent with Rh(I) being first oxidized in an electrochemically irreversible two-electron transfer process at peak anodic potentials ranging Epa(Rh) = 0.124–0.881 V vs. Fc/Fc+. For the ferrocene-containing complexes (4), and (8)–(10) the rhodium oxidation was followed by the electrochemically reversible oxidation of the ferrocenyl group in a one-electron transfer process at a slightly more positive potential. Relationships were established between the electrochemical quantity Epa(Rh) and kinetic parameter log k2 as well the sum of experimental group electronegativities (Gordy Scale) of the R and R′ groups (χR + χR′), the Hammett σ values (σR + σR′) and the Lever ligand parameter EL for the [Rh(RCOCHCOR′)(P(OPh)3)2] complexes: Epa(Rh) (vs. Fc/Fc+/V) = 0.31 (χR + χR′)–1.09 = 0.56 (σR + σR′) + 0.28 = SMEL) + (IM − 0.66 V) = −0.23 log k2 − 0.03 (k2 = second order rate constant for the oxidative addition of methyl iodide to rhodium). A profound shift of Epa(Rh) to a more positive potential was observed for Rh(I) substrates containing β-diketonato ligands with increasing electronegative substituents R and R′. An exponential dependence of Epa(Rh) on the pKa of the β-diketone was obtained.  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