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 kinetics of U(IV) produced by hydrazine reduction of U(VI) with platinum as a catalyst in nitric acid media was studied to reveal the reaction mechanism and optimize the reaction process. Electron spin resonance (ESR) was used to determine the influence of nitric acid oxidation. The effects of nitric acid, hydrazine, U(VI) concentration, catalyst dosage and temperature on the reaction rate were also studied. In addition, the simulation of the reaction process was performed using density functional theory. The results show that the influence of oxidation on the main reaction is limited when the concentration of nitric acid is below 0.5 mol/L. The reaction kinetics equation below the concentration of 0.5 mol/L is found as: -dc(UO22+)/dt)=kc0.5323(UO22+)c0.2074(N2H5+)c-0.2009(H+). When the temperature is 50 ℃, and the solid/liquid ratio r is 0.0667 g/mL, the reaction kinetics constant is k=0.00199 (mol/L)0.4712/min. Between 20 ℃ and 80 ℃, the reaction rate gradually increases with the increase of temperature, and changes from chemically controlled to diffusion-controlled. The simulations of density functional theory give further insight into the influence of various factors on the reaction process, with which the reaction mechanisms are determined according to the reaction kinetics and the simulation results.  相似文献   

2.
In this paper, a novel poly(aminosulfonic acid) modified glassy carbon electrode (PASA/GCE) for the determination of Sudan II was fabricated through electrochemical polymerizat ion. The electrochemical behavior of Sudan II at the modified electrode was studied by cyclic voltammetry. Results show that the modified electrode exhibits excellent electrocatalytic activity toward the electrochemical redox reaction of Sudan II. Under optimal experimental conditions, the oxidation peak current is linearly proportional to the concentration of Sudan II in the ranges of 4.0 × 10?8 to 1.0 × 10?6 mol L?1 and 1.0 × 10?6 to 1.2 × 10?5 mol L?1. The linear regression equations are i pa(A) = 2.87c + 3.74 × 10?6, r = 0.9977 and i pa(A) = 0.78c + 6.11 × 10?6, r = 0.9982, respectively, and the detection limit is 4.0 × 10?9 mol L?1. The novel method shows good recovery, reproducibility and sensitivity for the voltammetric determination of Sudan II in food samples.  相似文献   

3.
Spectrophotometric methods were used to investigate the rate of the reaction of Br2 with HCOOH in aqueous, acidic media. The reaction products are Br? and CO2. The kinetics of this reaction are complicated by both the formation of Br3? as Br? is formed and the dissociation of HCOOH into HCOO? and H+. Previous work on this reaction was carried out at acidities lower than the highest used here and led to the conclusion that only HCOO? reacts with Br2. It is agreed that this is by far the principal reaction. However, at the highest acidity experiments, an added small component of reaction was found, and it is suggested that it results from the direct reaction of Br2 with HCOOH itself. On this assumption, values of the rate constants for both reactions are derived here. The rate constant for the reaction of HCOO? with Br2 agrees with values previously reported, within a factor of 2 on the low side. The reaction involving HCOOH is more than 2000 times slower than the reaction involving HCOO?, but it does contribute to the overall rate as [H+] approaches 1M. These derived rate constants are able to simulate quantitatively the authors' absorbance-versus-time data, demonstrating the validity of their data treatment methods, if not mechanistic assignments. Finally, activation parameters were determined for both rate constants. The values obtained are: ΔE?(HCOOH + Br2) = 13.3 ± 1.1 kcal/mol, ΔS? (HCOOH + Br2) = ?28 ± 3 cal/deg mol, ΔE? (HCOO? + Br2) = 13.1 ± 0.9 kcal/mol, and ΔS?(HCOO? + Br2) = ?12 ± 1 cal/deg mol. That the activation energies of the two reactions turn out to be essentially identical does not support the authors' suggestion that both HCOOH and HCOO? react with Br2.  相似文献   

4.
The kinetics and mechanism for the reaction of NH2 with HONO have been investigated by ab initio calculations with rate constant prediction. The potential energy surface of this reaction has been computed by single‐point calculations at the CCSD(T)/6‐311+G(3df, 2p) level based on geometries optimized at the CCSD/6‐311++G(d, p) level. The reaction producing the primary products, NH3 + NO2, takes place via precomplexes, H2N???c‐HONO or H2N???t‐HONO with binding energies, 5.0 or 5.9 kcal/mol, respectively. The rate constants for the major reaction channels in the temperature range of 300–3000 K are predicted by variational transition state theory or Rice–Ramsperger–Kassel–Marcus theory depending on the mechanism involved. The total rate constant can be represented by ktotal = 1.69 × 10?20 × T2.34 exp(1612/T) cm3 molecule?1 s?1 at T = 300–650 K and 8.04 × 10?22 × T3.36 exp(2303/T) cm3 molecule?1 s?1 at T = 650–3000 K. The branching ratios of the major channels are predicted: k1 + k3 producing NH3 + NO2 accounts for 1.00–0.98 in the temperature range 300–3000 K and k2 producing OH + H2NNO accounts for 0.02 at T > 2500 K. The predicted rate constant for the reverse reaction, NH3 + NO2 → NH2 + HONO represented by 8.00 × 10?26 × T4.25 exp(?11,560/T) cm3 molecule?1 s?1, is in good agreement with the experimental data. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 678–688, 2009  相似文献   

5.
The kinetics of the aquation of (H2O)5Cr(O2CCCl3)2+ have been examined at 35–55°C and 1.00M ionic strength with [H+] = 0.01?1.00M. The reaction follows the rate equation -d ln [Crtotal]/dt = (a[H+]?1 + b + c[H+])/(1 + d[H+]), where [Crtotal] is the stoichiometric concentration of the complex. At 45°C a = (1.41 ± 0.03) × 10?7M/s, b = (1.66 ± 0.02) × 10?5 s?1, c = (7.0 ± 0.8) × 10?5M?1·S?1 and d = 2.3 ± 0.3M?1. Two mechanisms consistent with this rate law are discussed, with evidence being presented in favor of an ester hydrolysis mechanism involving steady-state intermediates. Equilibrium and activation parameters were determined.  相似文献   

6.
The dissolution of silver nanoparticles in their reaction with aqueous HNO3 solubilized to an reverse micelle solution of sodium bis(2-ethylhexyl)sulfosuccinate in decane is studied spectrophotometrically. A physicochemical model is advanced for quantifying the process kinetics on th basis of the following autocatalytic scheme: Ag0 + H+ + NO 3 ? → Ag+ + products (k 1), and Ag0 + Ag+ + NO 3 ? → 2Ag+ + products (k 2). The effective rate constant k 2 decreases with decreasing solubilization capacity V S/V O (where V S is the volume of the solubilized dispersed aqueous phase and V O is the volume of the micelle solution); the solubilization capacity determines the size of the micelle cavities in which the reaction between Ag0 and HNO3 occurs: k 2 = 74 (V S/V O) · 100% ≈ 3.8%), 41 (2.9), and 35 (2.0) L/(mol s). The effective constant k 1 is determined with a high uncertainty; the effect of V S/V O on k 1 has the opposite tendency.  相似文献   

7.
The complexation of 1-methyl-2-hydroxymethyl-imidazole (L) with Cu(I) and Cu(II) has been studied in aqueous acetonitrile (AN). Cu(I) forms three complexes, Cu(AN)L+, CuL2+, and Cu(AN)H?1L, with stability constants logK(Cu(AN)+ + L ? Cu(AN)L+) = 4.60 ± 0.02, logβ2 = 11.31 ± 0.04, and logK(Cu(AN)H?1L+H+ ? Cu(AN)L+) = 10.43 ± 0.08 in 0.15M AN. The main species for Cu(II) are CuL2+, CuH?1L+, CuH?1L2+, and CuH?2L2. The autoxidation of CuL2+ was followed with an oxygen sensor and spectrophotometrically. Competition between the formation of superoxide in a one-electron reduction of O2 and a path leading to H2O2 via binuclear (CuL2)2O was inferred from the rate law with ka = (2.31 ± 0.12) · 104M ?2S ?1, kb = (1.0 ± 0.2) · 103M ?1, kc = (2.85 ± 0.07) · 102M ?2S ?1, kd = 3.89 ± 0.14M ?1S ?1, ke = 0.112 ± 0.004, kf = (2.06 ± 0.24) · 10?10M S ?1, kg = (1.35 ± 0.07) · 10?7 S ?1, and kh = (6.8 ± 1.4) · 10?7M ?1 S ?1.  相似文献   

8.
Four Ag(I) complexes, [Ag(L1)2](NO3) (1), [Ag(L2)(NO3)] (2), [Ag(L3)3](NO3) (3), and [Ag(L4)2](NO3) (4), with ligands derived from halo-containing cyanoanilines (L1 = 4-amino-3fluorobenzonitrile, L2 = 4-amino-3-chlorobenzonitrile, L3 = 4-amino-3-bromobenzonitrile, L4 = 4-amino-2-bromobenzonitrile) were synthesized and characterized by C, H, and N elemental analysis, IR and 1H NMR spectroscopy and single crystal X-ray diffraction. Complexes 14 crystallized in the triclinic space group C2/c, P2(1)/n, P-1 and C2/c, respectively. In 1 and 4, Ag+ is four-coordinate with L1 or L4 to form 1-D {[Ag(L1/L4)2]+} polymeric cations. In 2, Ag+ is three-coordinate by two L2 ligands and one NO3? ligand to form a 1-D {[Ag(L2)(NO3)]} zigzag chain. In 3, Ag+ is four-coordinate by L3 to form a dinuclear [Ag(L3)3]+ cation. The NO3? is a 4-connector bridging group in 1 and 3 and a 5-connector bridging group in 2 and 4. The intermolecular hydrogen bonds and Ag?O weak interactions play important roles in forming 3-D networks of 14. The antibacterial activities for 14 were evaluated against Bacillus subtilis, Staphylococcus aureus and Escherichia coli with MTT method. The antibacterial results indicated that 2 showed the best inhibitory activity against the test bacterial strains, and was as potent as chloramphenicol.  相似文献   

9.
The reaction of peroxomonophosphoric acid and hydrazinium ion in acid perchlorate solutions occurs as per stoichiometry (i), and the rate law (ii) at large [N2H5 +], where K′d is the first acid dissociation constant of H3PO5 and k 1 and k 2 are rate constants found to be 2.6 × 10?4 s?1 and 5.0 × 10?2 M?1 s?1, respectively, at 35°. The reaction is greatly catalyzed by iodide ions. The mechanism involves a redox cycle I?/I2 and the rate is independent of [N2H5 +] in the presence of iodide ions. K′d was found to be 0.55 M?1 and independent of temperature.  相似文献   

10.
The oxidation of N,N-dimethylhydroxylamine (DMHAN) by nitrous acid is investigated in perchloric acid and nitric acid medium, respectively. The effects of H+, DMHAN, ionic strength and temperature on the reaction are studied. The rate equation in perchloric acid medium has been determined to be −d[HNO2]/dt = k[DMHAN][HNO2], where k = 12.8 ± 1.0 (mol/L)−1 min−1 when the temperature is 18.5 °C and the ionic strength is 0.73 mol/L with an activation energy about 41.5 kJ mol−1. The reaction becomes complicated when it is performed in nitric acid medium. When the molarity of HNO3 is higher than 1.0 mol/L, nitrous acid will be produced via the reaction between nitric acid and DMHAN. The reaction products are analyzed and the reaction mechanism is discussed in this paper.  相似文献   

11.
Poly(2-amino-5-(4-pyridinyl)-1,3,4-thiadiazole) (PAPT) modified glassy carbon electrode (GCE) was fabricated and used for the simultaneous determinations of dopamine (DA), uric acid (UA) and nitrite (NO2 ?) in 0.1 mol?L?1 phosphate buffer solution (PBS, pH 5.0) by using cyclic voltammetry and differential pulse voltammetry (DPV) techniques. The results showed that the PAPT modified GCE (PAPT/GCE) not only exhibited electrocatalytic activities towards the oxidation of DA, UA and NO2 ? but also could resolve the overlapped voltammetric signals of DA, UA and NO2 ? at bare GCE into three strong and well-defined oxidation peaks with enhanced current responses. The peak potential separations are 130 mV for DA–UA and 380 mV for UA–NO2 ? using DPV, which are large enough for the simultaneous determinations of DA, UA and NO2 ?. Under the optimal conditions, the anodic peak currents were correspondent linearly to the concentrations of DA, UA and NO2 ? in the ranges of 0.95–380 μmol?L?1, 2.0–1,000 μmol?L?1 and 2.0–1,200 μmol?L?1 for DA, UA and NO2 ?, respectively. The correlation coefficients were 0.9989, 0.9970 and 0.9968, and the detection limits were 0.2, 0.35 and 0.6 μmol?L?1 for DA, UA and NO2 ?, respectively. In 0.1 mol?L?1 PBS pH 5.0, the PAPT film exhibited good electrochemical activity, showing a surface-controlled electrode process with the apparent heterogeneous electron transfer rate constant (k s) of 25.9 s?1 and the charge–transfer coefficient (α) of 0.49, and thus displayed the features of an electrocatalyst. Due to its high sensitivity, good selectivity and stability, the modified electrode had been successfully applied to the determination of analytes in serum and urine samples.  相似文献   

12.
The kinetics and mechanism for the reaction of NH2 with HONO2 have been investigated by ab initio calculations with rate constant prediction. The potential energy surface of this reaction has been computed by single‐point calculations at the CCSD(T)/6‐311+G(3df, 2p) level based on geometries optimized at the B3LYP/6‐311+G(3df, 2p) level. The reaction producing the primary products, NH3 + NO3, takes place via a precursor complex, H2N…HONO2 with an 8.4‐kcal/mol binding energy. The rate constants for major product channels in the temperature range 200–3000 K are predicted by variational transition state or variational Rice–Ramsperger–Kassel–Marcus theory. The results show that the reaction has a noticeable pressure dependence at T < 900 K. The total rate constants at 760 Torr Ar‐pressure can be represented by ktotal = 1.71 × 10?3 × T?3.85 exp(?96/T) cm3 molecule?1 s?1 at T = 200–550 K, 5.11 × 10?23 × T+3.22 exp(70/T) cm3 molecule?1 s?1 at T = 550–3000 K. The branching ratios of primary channels at 760 Torr Ar‐pressure are predicted: k1 producing NH3 + NO3 accounts for 1.00–0.99 in the temperature range of 200–3000 K and k2 + k3 producing H2NO + HONO accounts for less than 0.01 when temperature is more than 2600 K. The reverse reaction, NH3 + NO3 → NH2 + HONO2 shows relatively weak pressure dependence at P < 100 Torr and T < 600 K due to its precursor complex, NH3…O3N with a lower binding energy of 1.8 kcal/mol. The predicted rate constants can be represented by k?1 = 6.70 × 10?24 × T+3.58 exp(?850/T) cm3 molecule?1 s?1 at T = 200–3000 K and 760 Torr N2 pressure, where the predicted rate at T = 298 K, 2.8 × 10?16 cm3 molecule?1 s?1 is in good agreement with the experimental data. The NH3 + NO3 formation rate constant was found to be a factor of 4 smaller than that of the reaction OH + HONO2 producing the H2O + NO3 because of the lower barrier for the transition state for the OH + HONO2. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 42: 69–78, 2010  相似文献   

13.
Polynuclear Pd(II) and Ni(II) complexes of macrocyclic polyamine 3,6,9,16,19,22‐hexaazatricyclo[22.2.2.211,14]‐triaconta 11,13,24,26(l),27,29‐hexaene (L) in solution were investigated by electrospray ionization mass spectrometry (ESIMS). For methanol solution of complexes M2LX4 (M = Pd(II) and Ni(II), X= Cl and I), two main clusters of peaks were observed which can be assigned to [M2LX3]+ and [M2LX2]2+. When Pd2LCl4 was treated with 2 or 4 mol of AgNO3, it gave rise formation of Pd2LCl2 (NO3)2 · H2O and [Pd2L(H2O)m(NO3)n](4‐n)+, respectively. ESMS spectra show that the dissociation of the former in the ionization process gave peaks of [Pd2LCl2]2+ and [(Pd2LCl2)NO3]+, while dissociation of the later gave the peaks of [Pd2L(CH3CO2)2]2+ and [Pd2L(CH3CO2)2](NO3) + in the presence of acetic acid. Similar species were observed for Pd2LI4 when treated with 4 mol of AgNO3. When [Pd2L · (H2O)m(NO3)n](4‐n)+ reacted with 2 mol of oxalate anions at 40°C, [Pd4L2(C2O4)2(NO3)2]2+ and [Pd4L2(C2O4)2 (NO3)]3+ were detected. This implies the formation of square‐planar molecular box Pd4L2(C2O4)2(NO3)4 in which C2O4? may act as bridging ligands as confirmed by crystal structure analysis. The dissociation form and the stability of complex cations in gaseous state are also discussed. This work provides an excellent example of the usefulness of ESIMS in the identification of metal complexes in solution.  相似文献   

14.
New hybrid ligands are reported that combine two types of popular donor groups within a single linear scaffold, viz., a central pyrazolate bridge and two appended bis(N‐heterocyclic carbene) units; the ligand strands thus provide two potentially tridentate {NCC} compartments. The pyrazole/tetraimidazolium proligands, [H5L1](PF6)4 and [H5L2](PF6)4 , were synthesized via multi‐step protocols, and the NH prototropy of [H5L1](PF6)4 was examined by variable temperature (VT) NMR spectroscopy, giving solvent dependent activation parameters (ΔH? = 27.6 kJ · mol–1, ΔS? = –125 J · mol–1 · K–1 in [D3]MeCN; ΔH? = 40.4 kJ · mol–1, ΔS? = –86.9 J · mol–1 · K–1 in [D6]DMSO) that are in the range typical for pyrazoles. Reaction of the proligands with Ag2O gave hexametallic complexes [Ag6(L1)2](PF6)4 and [Ag6(L2)2](PF6)4 that involve all six potential donor atoms of the ligands, viz. the four CNHC and two Npz donors, in metal coordination. X‐ray crystallography revealed a chair‐like central {Ag6} deck in both complexes but different arrangements of the ligand strands, which goes along with significantly different AgI ··· AgI distances that indicate more pronounced argentophilic interactions in case of [Ag6(L1)2]4 +.  相似文献   

15.
The SCN Ion as an Ambidentate Ligand – Synthesis and Crystal Structures of (Bu4N)4[Ag2Fe2(SCN)12] and (Et4N)2 [Ag2Fe(SCN)6] In (Bu4N)4[Ag2Fe2(SCN)12] · 2 CH3NO2 ( 1 ) and (Et4N)2[Ag2Fe(SCN)6] ( 2 ) the ambidentate SCN anions link Ag+ with Fe3+ and Fe2+ centers, respectively. The tetranuclear anions in 1 are built from [Fe(NCS)6]3– groups connected by Ag+ ions. In 2 the same bridging pattern leads to polymeric anionic chains containing [Fe(NCS)6]4– groups linked by Ag+ ions. (Bu4N)4[Ag2Fe2(SCN)12] · 2 CH3NO2 ( 1 ): a = 1184.10(10), b = 1370.80(10), c = 1776.5(2) pm, α = 99.090(10), β = 102.100(10), γ = 100.360(10)°, V = 2715.5(4) · 106 pm3, space group P1; (Et4N)2[Ag2Fe(SCN)6] ( 2 ): a = 1607.0(2), b = 1006.92(9), c = 1096.13(9) pm, V = 1773.7(3) · 106 pm3, space group Pnnm.  相似文献   

16.
Four coordination polymers, [Ag(L1)](m-Hbdc) (1), [Ag(L1)]2(p-bdc)?·?8H2O (2), [Ag(Hbtc)(L1)][Ag(L1)]?·?2H2O (3) and [Ag2(L2)2](OH-bdc)2?·?4H2O (4), where L1?=?1,1′-(1,4-butanediyl)bis(imidazole), L2?=?1,2-bis(imidazol-1-ylmethyl)benzene, m-H2bdc?=?1,3-benzenedicarboxylic acid, p-H2bdc?=?1,4-benzenedicarboxylic acid, H3btc?=?1,3,5-benzenetricarboxylic acid, and OH–H2bdc?=?5-hydroxisophthalic acid, were synthesized under hydrothermal conditions. Compound 1 contains a–Ag-L1–Ag-L1–chain and a hydrogen-bonding interaction induced–(m-Hbdc)-(m-Hbdc)–chain. Compound 2 consists of two independent–Ag-L1–Ag-L1–chains. P-bdc anions are not coordinated. Hydrogen bonds form a 3D supramolecular structure. A novel (H2O)16 cluster is formed by lattice water molecules in 2. Compound 3 contains a–Ag-L1–Ag-L1–and a–Ag(Hbtc)-L1–Ag(Hbtc)-L1–chain. The packing diagram shows a 2D criss-cross supramolecular structure, with?π?···?π?and C–H ···?π?interactions stabilizing the framework. Compound 4 contains a [Ag2(L2)2]2+ dimer with hydrogen-bonding,?π?··· π, and Ag ··· O interactions forming a 3D supramolecular framework. The luminescent properties for these compounds in the solid state are discussed.  相似文献   

17.
The gas‐phase kinetics of CHBr2 + NO2 and CH3CHBr + NO2 reactions have been studied in direct time resolved measurements using a tubular flow reactor coupled to a photoionization mass spectrometer. The radicals were generated by pulsed laser photolysis of bromoform and 1,1‐dibromoethane at 248 nm. The subsequent decays of the radical concentrations were monitored as a function of [NO2] under pseudo–first‐order conditions. The rate coefficients of both reactions are independent of bath gas (He) pressure and display negative temperature dependence under the conditions of 2–6 Torr pressure (He) and 250–480 K. The obtained bimolecular rate coefficients are k(CHBr2 + NO2) = (9.8 ± 0.4) × 10?12 (T/300 K)?1.65 ± 0.18 cm3 s?1 (288–483 K) and k(CH3CHBr + NO2) = (2.27 ± 0.06) × 10?11 (T/300 K)?1.28 ± 0.11 cm3 s?1 (250–483 K), with the uncertainties given as one standard error. Estimated overall uncertainties in the measured bimolecular reaction rate coefficients are ±25%. The reaction products identified were CBr2O for the CHBr2 + NO2 reaction and CHBrO and CH3CHO with minor amounts of CH3 for the CH3CHBr + NO2 reaction, respectively. © 2012 Wiley Periodicals, Inc. Int J Chem Kinet 44: 767–777, 2012  相似文献   

18.
A new voltammetric sensor, based on a new p-tert-butylcalix[4]arene derivative (TCAD) modified glassy carbon electrode (GCE) using Langmuir–Blodgett (LB) technique, was designed successfully and used for recognition and determination of Ag+. The π?-?A isotherms suggested that the monolayer of TCAD can coordinate with Ag+ at the air–water surface. Under the optimum experimental conditions, this voltammetric sensor shows a linear voltammetric response for Ag+ in the range of 1.0?×?10?8?~?6.0?×?10?6?mol?L?1 with detection limit 5.0?×?10?9?mol?L?1. The high sensitivity, selectivity, and stability of this LB film modified electrode also demonstrate its practical application for a simple, rapid and economical determination of Ag+ in water sample.  相似文献   

19.
The kinetics of reactions of HCCl with NO and NO2 were investigated over the temperature ranges 298–572 k and 298–476 k, respectively, using laser‐induced fluorescence spectroscopy to measure total rate constants and time‐resolved infrared diode laser absorption spectroscopy to probe reaction products. Both reactions are fast, with k(HCCl + NO) = (2.75 ± 0.2) × 10?11 cm3 molecule?1 s?1 and k(HCCl + NO2) = (1.10 ± 0.2) × 10?10 cm3 molecule?1 s?1 at 296 K. Both rate constants displayed only a slight temperature dependence. Detection of products in the HCCl + NO reaction at 296 K indicates that HCNO + Cl is the major product with a branching ratio of ? = 0.68 ± 0.06, and NCO + HCl is a minor channel with ? = 0.24 ± 0.04. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 34: 12–17, 2002  相似文献   

20.
The composition and the stability constants of the complexes formed between Ag(I) and ligands of the type where n,m = 2,2; 2,3; 2,4; 3,3; 3,4; 4,4, have been determined by a pH/pM-metric method at 25°C in a 0.5 M (K)NO3 medium. The determination of the stability constants proceeded by a combination of a graphical method and a least squares minimization procedure. The complex formation is discussed in comparison with silver complexes formed with S-containing mono-amines and in terms of the Taft σ*-parameters for the substituents. Below pH = 4, the protonated complexes AgLH23+ and AgL2H45+ are formed and the thioethergroup is the only coordinating centre. In neutral and alkaline medium there was evidence for the species AgLH2+, Ag2L2H24+, Ag2L2H3+, Ag2L22+, AgL2H34+, AgL2H2+, AgL2 + and Ag2L2+ in which Ag+? S and Ag+? NH2 bonds are involved. It is shown that in the Ag2L22+ and Ag2L2+ complexes the ligands effectively coordinate through all available donor centres.  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