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 46 毫秒
1.
Interaction of N,N′,N″,N-tetramethyltetra-2,3-pyridinoporphyrazinatocopper(II), ([Cu(2,3-tmtppa)]4+) and N,N′,N″,N-tetramethyltetra-3,4-pyridinoporphyrazinatocopper(II), ([Cu(3,4-tmtppa)]4+) with calf thymus DNA was studied in 1 mM phosphate buffer and low ionic strength (5 mM NaCl) at various temperatures by UV-visible and circular dichroism (CD) spectroscopies and viscometric method. The binding constants were determined from the changes in the visible part of porphyrazine complexes spectra using SQUAD software. The values of K have been obtained (7.9±0.4)×104 and (2.2±0.1)×105 M−1 for [Cu(2,3-tmtppa)]4+ and [Cu(3,4-tmtppa)]4+, respectively at 27 °C. The higher affinity of 3,4-isomer of Cu complex towards DNA with respect to the 2,3-isomer was attributed to favorable external positioning of the cationic charges in former, which enables superior interaction with the DNA duplex. The thermodynamic parameters (ΔG°, ΔH°, ΔS°) were calculated by van't Hoff equation. The enthalpy and entropy changes were determined, +34.2±3.6 kJ mol−1 and +207.8±12.70 J mol−1 K−1 for [Cu(2,3-tmtppa)]4+ and +49.7±2.1 kJ mol−1 and +267.8±7.9 J mol−1 K−1 for [Cu(3,4-tmtppa)]4+. The existence of extensive hypochromicity, large red shift and negative CD in the visible part of [Cu(3,4-tmtppa)]4+ spectra suggested an intercalation binding mode. Analysis of the moderate hypochromicity, red shift and bisignate CD in the Q-band absorption region of [Cu(2,3-tmtppa)]4+ spectra possibly led us to the coexistence of intercalation and outside binding modes. The influence of the ionic strength on the binding parameters and binding modes was investigated. The results show that increase in ionic strength causes the decrease in the binding constants. It was also concluded that increase in ionic strength affects the binding characteristics of positively charged complexes with DNA.

The increase in DNA viscosity in the presence of Cu–tmtppa complexes is attributed to the lengthening of DNA helix due to the intercalation. This result is consistent with conclusions obtained from the spectroscopic techniques.  相似文献   


2.
Twenty-two isomers/conformers of C3H6S+√ radical cations have been identified and their heats of formation (ΔHf) at 0 and 298 K have been calculated using the Gaussian-3 (G3) method. Seven of these isomers are known and their ΔHf data are available in the literature for comparison. The least energy isomer is found to be the thioacetone radical cation (4+) with C2v symmetry. In contrast, the least energy C3H6O+√ isomer is the 1-propen-2-ol radical cation. The G3 ΔHf298 of 4+ is calculated to be 859.4 kJ mol−1, ca. 38 kJ mol−1 higher than the literature value, ≤821 kJ mol−1. For allyl mercaptan radical cation (7+), the G3 ΔHf298 is calculated to be 927.8 kJ mol−1, also not in good agreement with the experimental estimate, 956 kJ mol−1. Upon examining the experimental data and carrying out further calculations, it is shown that the G3 ΔHf298 values for 4+ and 7+ should be more reliable than the compiled values. For the five remaining cations with available experimental thermal data, the agreement between the experimental and G3 results ranges from fair to excellent.

Cation CH3CHSCH2+√ (10+) has the least energy among the eleven distonic radical cations identified. Their ΔHf298 values range from 918 to 1151 kJ mol−1. Nevertheless, only one of them, CH2=SCH2CH2+√ (12+), has been observed. Its G3 ΔHf298 value is 980.9 kJ mol−1, in fair agreement with the experimental result, 990 kJ mol−1.

A couple of reactions involving C3H6S+√ isomers CH2=SCH2CH2+√ (12+) and trimethylene sulfide radical cation (13+) have also been studied with the G3 method and the results are consistent with experimental findings.  相似文献   


3.
Amberlite XAD-16 resin has been functionalized using nitrosonaphthol as a ligand and characterized employing elemental, thermogravimetric analysis and FT-IR spectroscopy. The sorption of Ni(II) and Cu(II) ions onto this functionalized resin is investigated and optimized with respect to the sorptive medium (pH), shaking speed and equilibration time between liquid and solid phases. The monitoring of the influence of diverse ions on the sorption of metal ions has revealed that phosphate, bicarbonate and citrate reduce the sorption up to 10–14%. The sorption data followed Langmuir, Freundlich, and Dubinin–Radushkevich (D–R) isotherms. The Freundlich parameters computed are 1/n = 0.56 ± 0.03 and 0.49 ± 0.05, A = 9.54 ± 1.5 and 6.0 ± 0.5 mmol g−1 for Ni(II) and Cu(II) ions, respectively. D–R isotherm yields the values of Xm = 0.87 ± 0.07 and 0.35 ± 0.05 mmol g−1 and of E = 9.5 ± 0.23 and 12.3 ± 0.6 kJ mol−1 for Ni(II) and Cu(II) ions, respectively. Langmuir characteristic constants estimated are Q = 0.082 ± 0.005 and 0.063 ± 0.003 mmol g−1, b = (4.7 ± 0.2) × 104 and (7.31 ± 0.11) × 104 l mol−1 for Ni(II) and Cu(II) ions, respectively. The variation of sorption with temperature gives thermodynamic quantities of ΔH = −58.9 ± 0.12 and −40.38 ± 0.11 kJ mol−1, ΔS = −183 ± 10 and −130 ± 8 J mol−1 K−1 and ΔG = −4.4 ± 0.09 and −2.06 ± 0.08 kJ mol−1 at 298 K for Ni(II) and Cu(II) ions, respectively. Using kinetic equations, values of intraparticle transport and of first order rate constant have been computed for both the metal ions. The sorption procedure is utilized to preconcentrate these ions prior to their determination in tea, vegetable oil, hydrogenated oil (ghee) and palm oil by atomic absorption spectrometry using direct and standard addition methods.  相似文献   

4.
The enthalpy and entropy of sublimation of N-ethylthiourea were obtained from the temperature dependence of its vapour pressure measured by both the torsion–effusion and the Knudsen effusion method in the temperature range 360–380 K. The compound undergoes no solid-to-solid phase transition or decomposition below 380 K. The pressure against reciprocal temperature resulted in lg(p, kPa) = (13.40 ± 0.27) − (6067 ± 102) /T(K). The molar sublimation enthalpy and entropy at the mid interval temperature were ΔsubHm(370 K) = (116.1 ± 2.0) kJ mol−1 and ΔsubSm(370 K) = (218.0 ± 5.2) J mol−1 K−1, respectively. The same quantities derived at 298.15 K were (118.8 ± 2.1) kJ mol−1 and (226.1 ± 5.5) J mol−1 K−1, respectively.  相似文献   

5.
The receptor–ligand interaction between the cardiac glycoside Ouabain and purified, membrane-bound as well as micellar Na,K-ATPase is investigated. Calorimetric titrations are carried out with micromolar concentrations of the phosphorylated protein in the presence of Mg2+. The measured heat changes provide evidence for an exothermic, high affinity and specific receptor binding process as well as for a low affinity, nonspecific binding to the lipid part of the nanoparticulate membrane fragments. The degree of lipid binding markedly depends on the lipid composition of the tissue. The measured time course of the heat change resulting from specific binding to the receptor site is unusually slow and is limited by the binding kinetics of the ligand. A course estimation of the Ouabain binding kinetics leads to a rate constant around 104 mol−1 l s−1. Receptor binding is characterized by affinities ranging between 107 and 108 mol−1 l, ΔH values around −95 kJ mol−1 and ΔS values of about −130 J K−1 mol−1 at 25°C. The enthalpic contribution is assumed to be mainly due to hydrogen bond formations between the ligand and the receptor site whereas the large, negative entropy change may be attributed to an increased interaction between water and the protein as a consequence of a conformational transition. The evaluation of the titrations provides stoichiometric coefficients around 0.55, which implies that only about 50–60% of the Na,K-ATPase protomers are capable to bind the cardiotonic steroid. This result is consistent with radioactive phosphorylation studies and appears to be a typical feature of kidney-type Na,K-ATPase preparations. Possible implications of this finding are discussed. As a general result, this study demonstrates how simple and suitable calorimetric titrations with micromolar protein concentrations can be for the purpose of a quantitative characterization of a receptor in nanoparticulate membrane systems.  相似文献   

6.
The benzophenone-sensitised photodimerizations of N-acetyl- and N-propionyldibenz[e,f]azepine were investigated in acetone as the solvent. In both the systems, the 1H NMR analysis of the products revealed two isomeric photodimers differing in the chemical shifts and coupling constants of the cyclobutane protons, aromatic protons and the protons of the acetyl or propionyl group. Upon raising the temperature to ca. 70 °C the signals merge. The findings can be ascribed to a single thermally restricted conformational process such as the rotation about the C–N amide bond. The process exhibits free activation energies: ΔG#=(74±2) kJ mol−1 (N-acetyl) and ΔG#=(70±2) kJ mol−1 (N-propionyl).  相似文献   

7.
Ruthenium(III) acetylacetonate was employed for the first time as homogeneous catalyst in the hydrolysis of sodium borohydride. Ruthenium(III) acetylacetonate was not reduced by sodium borohydride under the experimental conditions and remains unchanged after the catalysis. Poisoning experiments with mercury and trimethylphosphite provide compelling evidence for the fact that ruthenium(III) acetylacetonate is indeed a homogenous catalyst in the hydrolysis of sodium borohydride. Kinetics of the ruthenium(III) acetylacetonate catalyzed hydrolysis of sodium borohydride was studied depending on the catalyst concentration, substrate concentration, and temperature. The hydrogen generation was found to be first order with respect to both the substrate concentration and catalyst concentration. The activation parameters of this reaction were also determined from the evaluation of the kinetic data: activation energy; Ea = 58.2 ± 2.6 kJ mol−1, the enthalpy of activation; ΔH# = 55.7 ± 2.5 kJ mol−1 and the entropy of activation ΔS# = 118 ± 5 J mol−1 K−1. Ruthenium(III) acetylacetonate was found to be highly active catalyst providing 1200 turnovers over 180 min in hydrogen generation from the hydrolysis of sodium borohydride before deactivation.  相似文献   

8.
The low temperature heat capacities of N-(2-cyanoethyl)aniline were measured with an automated adiabatic calorimeter over the temperature range from 83 to 353 K. The temperature corresponding to the maximum value of the apparent heat capacity in the fusion interval, molar enthalpy and entropy of fusion of this compound were determined to be 323.33 ± 0.13 K, 19.4 ± 0.1 kJ mol−1 and 60.1 ± 0.1 J K−1 mol−1, respectively. Using the fractional melting technique, the purity of the sample was determined to be 99.0 mol% and the melting temperature for the tested sample and the absolutely pure compound were determined to be 323.50 and 323.99 K, respectively. A solid-to-solid phase transition occurred at 310.63 ± 0.15 K. The molar enthalpy and molar entropy of the transition were determined to be 980 ± 5 J mol−1 and 3.16 ± 0.02 J K−1 mol−1, respectively. The thermodynamic functions of the compound [HT − H298.15] and [ST − S298.15] were calculated based on the heat capacity measurements in the temperature range of 83–353 K with an interval of 5 K.  相似文献   

9.
The enthalpies of reactions between alkaline-earth cuprates M2CuO3 (M = Ca, Sr) and hydrochloric acid were measured in a hermetic swinging calorimeter at 298.15 K. The M2CuO3 samples were prepared by solid-phase synthesis from calcium or strontium carbonate and copper oxide and characterized by X-ray powder diffraction, EDX and wet analysis. The standard enthalpies of formation obtained for the cuprates, −1431 ± 4 kJ mol−1 for Ca2CuO3 and −1374 ± 3 kJ mol−1 for Sr2CuO3, are discussed and compared with previous experimental and assessed values.  相似文献   

10.
The new strong anion exchanger (PUFIX) from polyurethane foam was prepared by coupling of the primary amine of the foam matrix with ethyl iodide. PUFIX was characterized using different tools (IR spectra, elemental analysis, density and thermal analysis). The sorption properties of the new anion exchanger (PUFIX) and chromatographic behaviour for separation and determination of palladium(II) ions at low concentrations from aqueous iodide or thiocyanate media were investigated by a batch and dynamic processes. The maximum sorption of Pd(II) was in the pH range of 0.3–2. The kinetics of sorption of the Pd(II) by the PUFIX was found to be fast with average values of half-life of sorption (t1/2) of 3.32 min. The variation of the sorption of Pd(II) with temperature gives average values of ΔH, ΔS, ΔG and ΔE to be −38.3 kJ mol−1, −100.7 J K−1 mol−1, −8.3 and 11.8 kJ mol−1, respectively. The sorption capacity of PUFIX was 1.69 mmol g−1 for Pd(II), preconcentration factors of values ≈250 and the recovery 99–100% were achieved (R.S.D. ≈ 1.24%). The lower detection limit, 1.28 ng mL−1 was evaluated using spectrophotometric method (R.S.D. ≈ 2.46%).  相似文献   

11.
The reactions of hydroxyl radical, hydrogen atom and hydrated electron intermediates of water radiolysis with N-isopropylacrylamide (NIPAAm) were studied by pulse radiolysis in dilute aqueous solutions. OH, H and eaq react with NIPAAm with rate coefficient of (6.9±1.2)×109, (6.6±1)×109, and (1.0±0.2)×1010 mol−1 dm3 s−1. In OH and H radical addition to the double bond mainly -carboxyalkyl type radicals form, (OHCH2CHC(N-i-C3H7)O and CH3CHC(N-i-C3H7)O). In reaction of eaq oxygen atom centered radical anion is produced (CH2CHC(N-i-C3H7)O), the anion undergoes reversible protonation with pKa=8.7. There is also an irreversible protonation on the β-carbon atom that produces the same radical as forms in H atom reaction (CH3CHC(N-i-C3H7)O). The -carboxyalkyl type radicals at low NIPAAm concentration (0.1–1 mmol dm−3) mainly disappear in self-termination reactions, 2kt,m=8.4×108 mol−1 dm3 s−1. At higher concentrations the decay curves reflect the competition of the self-termination and radical addition to monomer (propagation). The termination rate coefficient of oligomer radicals containing a few monomer units is 2kt≈2×108 mol−1 dm3 s1.  相似文献   

12.
Saran L  Cavalheiro E  Neves EA 《Talanta》1995,42(12):2027-2032
The highly neutralized ethylenediaminetetraacetate (EDTA) titrant (95–99% as Y4− anion) precipitates with Ag+ cations to form the Ag4Y species, in aqueous medium, which is well characterized from conductometric titration, thermal analysis and potentiometric titration of the silver content of the solid. The precipitate dissolves in excess Y4− to form a complex, AgY3−. Equilibrium studies at 25°C and ionic strength 0.50 M (NaNO3) have shown from solubility and potentiometric measurements that the formation constant (95% confidence level) β1 = (1.93 ± 0.07) × 105 M−1 and the solubility products are KS0 = [Ag +]4[Y4−] = (9.0 ± 0.4) × 10−18 M5 and KS1 = [Ag +]3[AgY3−] = (1.74 ± 0.08) × 10−12 M4. The presence of Na+, rather than ionic strength, markedly affects the equilibrium; the data at ionic strength 0.10 M are: β1 = (1.19 ± 0.03) × 106 M−1, KS0 = (1.6 ± 0.4) × 10−19 M5 and KS1 = (1.9 ± 0.5) × 10−13 M4; at ionic strength tending to zero; β1 = (1.82 ± 0.05) × 107 M−1, KS0 = (2.6 ± 0.8) × 10−22 M5 and KS1 = (5 ± 1) × 10−15 M4. The intrinsic solubility is 2.03 mM silver (I) in 0.50 M NaNO3. Well-defined potentiometric titration curves can be taken in the range 1–2 mM with the Ag indicator electrode. Thermal analysis revealed from differential scanning calorimetry a sharp exothermic peak at 142°C; thermal gravimetry/differential thermal gravimetry has shown mass loss due to silver formation and a brown residue, a water-soluble polymeric acid (decomposition range 135–157°C), tending to pure silver at 600°C, consistent with the original Ag4Y salt.  相似文献   

13.
The far-UV (193 nm) laser flash photolysis of nitrogen-saturated isooctane solutions of 1,1-dimethylsiletane allows the direct detection of 1,1-dimethylsilene as a transient species, which (at low laser intensities) decays with pseudo-first-order kinetics (τ 10 μs) and exhibits a UV absorption spectrum with λmax 255 nm. Characteristic rapid quenching is observed for the silene with methanol (kMcOH = (4.9 ± 0.2) × 109 M−1 s−1), tert-butanol (kBuOH = (1.8 ± 0.1) × 109 M−1 s−1) and oxygen (kO2 = (2.0 ± 0.5) × 108 M−1 s−1). The Arrhenius activation parameters for the reaction with methanol have been determined to be Ea = −2.6 ± 0.6 kcal mol−1 and log A = 7.7 ± 0.3.  相似文献   

14.
Excitation of solutions of Fe(bipy)2(CN)2 by a 266-nm laser pulse produces a hydrated electron and the oxidized complex, Fe(bipy)2 (CN)2+, in the primary photochemical step, in homogeneous aqueous solution as well as in aqueous solutions containing cetyltrimethylammonium bromide (CTAB) or sodium dodecyl sulfate (SDS) micelles. In all cases nascent hydrated electrons react with ground state Fe(bipy)2(CN)2 to form Fe(bipy)2(CN)2, and comparison of the decay constants in the three media (H2O: k = 2.8 × 1010 M−1 s−1; CTAB: k = 2.9 × 1010 M−1 s−1; SDS: k = 5.5 × 109 M−1 s−1), shows that the reaction is essentially unaffected by CTAB micelles but is much slower in SDS solution. Similar micellar effects were found for the back reaction between eaq and Fe(bpy)2(CN)2+. Rate constants for the scavenging of the photogenerated hydrated electrons by methyl viologen (MV2+) cations and NO3 anions were measured in the three systems, and the results indicate that for scavenging by MV2+ the rate constants are decreased in the micelle systems (k in H2O, 8.4 × 1010; CTAB, 3.5 × 1010 and SDS, 1.58 × 1010 M−1 s−1), whereas for NO3 the CTAB micelle decreases while the SDS micelle enhances the scavenging compared to water solution (k in H2O, 8.3 × 109; CTAB, 7 × 108; and SDS, 2.05 × 1010 M−1 s−1). For the comproportionation reaction between Fe(bipy)2(CN)2+ and Fe(bipy)2(CN)2 both micelles reduce the rate (k in H2O, 3.3 × 1010; CTAB, 2.3 × 1010; and SDS, 1.05 × 1010 M−1s−1), but while the reaction of Fe(bipy)2(CN)2+ with MV+ is increased in CTAB compared to water, it is slowed in SDS (k in H2O, 2.4 × 1010; CTAB, 8.9 × 1010; and SDS, 1.8 × 1010 M−1s−1). All effects observed in these microheterogeneous systems can be uniformly interpreted in terms of Coulombic interactions between the actual reactants and the charged surface of the micelles.  相似文献   

15.
Low-temperature heat capacities of the complex Zn(Thr)SO4·H2O (s) have been precisely measured with a small sample adiabatic calorimeter over the temperature range from 78 to 373 K. The initial dehydration temperature of the complex (Td=325.50 K) has been obtained by analysis of the heat-capacity curve. The experimental values of molar heat capacities have been fitted to a polynomial equation by least square method. The standard molar enthalpy of formation of the complex has been determined from the enthalpies of dissolution (ΔdHmΘ) of [ZnSO4·7H2O (s) +Thr (s)] and Zn(Thr)SO4·H2O (s) in 100 ml of 2 mol dm−3 HCl solvent as: ΔfHm,Zn(Thr)SO4·H2OΘ=−2111.7±3.4 kJ mol−1. These experiments were made by using an isoperibol solution calorimeter at 298.15 K.  相似文献   

16.
Gaussian-2 ab initio calculations were performed to examine the six modes of unimolecular dissociation of cis-CH3CHSH+ (1+), trans-CH3CHSH+ (2+), and CH3SCH2+ (3+): 1+→CH3++trans-HCSH (1); 1+→CH3+trans-HCSH+ (2); 1+→CH4+HCS+ (3); 1+→H2+c-CH2CHS+ (4); 2+→H2+CH3CS+ (5); and 3+→H2+c-CH2CHS+ (6). Reactions (1) and (2) have endothermicities of 584 and 496 kJ mol−1, respectively. Loss of CH4 from 1+ (reaction (3)) proceeds through proton transfer from the S atom to the methyl group, followed by cleavage of the C–C bond. The reaction pathway has an energy barrier of 292 kJ mol−1 and a transition state with a wide spectrum of nonclassical structures. Reaction (4) has a critical energy of 296 kJ mol−1 and it also proceeds through the same proton transfer step as reaction (3), followed by elimination of H2. Formation of CH3CS+ from 2+ (reaction (5)) by loss of H2 proceeds through protonation of the methine (CH) group, followed by dissociation of the H2 moiety. Its energy barrier is 276 kJ mol−1. On both the MP2/6-31G* and QCISD/6-31G* potential-energy surfaces, the H2 1,1-elimination from 3+ (reaction (6)) proceeds via a nonclassical intermediate resembling c-CH3SCH2+ and has a critical energy of 269 kJ mol−1.  相似文献   

17.
Hydrated strontium borate, SrB4O7·3H2O, has been synthesized and characterized by XRD, FT-IR, DTA-TG and chemical analysis. The molar enthalpy of solution of SrB4O7·3H2O in 1 mol dm−3 HCl(aq) was measured to be (21.15 ± 0.29) kJ mol−1. With incorporation of the previously determined enthalpies of solution of Sr(OH)2·8H2O(s) in [HCl(aq) + H3BO3(aq)] and H3BO3 in HCl(aq), and the enthalpies of formation of H2O(l), Sr(OH)2·8H2O(s) and H3BO3(s), the enthalpy of formation of SrB4O7·3H2O was found to be −(4286.7 ± 3.3) kJ mol−1.  相似文献   

18.
This study investigated the kinetics of a thermal dechlorination and oxidation of neodymium oxychloride (NdOCl) by using a non-isothermal thermogravimetric (TG) analysis under various oxygen partial pressures. The results of the isoconversional analysis of the TG data suggests that the dechlorination and oxidation of NdOCl follows a single step reaction and the observed activation energy was determined as 228.3 ± 6.1 kJ mol−1. A kinetic rate equation was derived for a conversion of the NdOCl into Nd2O3 with a power law model, f() = 2/3−1/2. The oxygen power dependency and the pre-exponential factor were determined as 0.315 and 3.55 × 105 s−1 Pa−0.315, respectively.  相似文献   

19.
CdII complexes with glycine (gly) and sarcosine (sar) were studied by glass electrode potentiometry, direct current polarography, virtual potentiometry, and molecular modelling. The electrochemically reversible CdII–glycine–OH labile system was best described by a model consisting of M(HL), ML, ML2, ML3, ML(OH) and ML2(OH) (M = CdII, L = gly) with the overall stability constants, as log β, determined to be 10.30 ± 0.05, 4.21 ± 0.03, 7.30 ± 0.05, 9.84 ± 0.04, 8.9 ± 0.1, and 10.75 ± 0.10, respectively. In case of the electrochemically quasi-reversible CdII–sarcosine–OH labile system, only ML, ML2 and ML3 (M = CdII, L = sar) were found and their stability constants, as log β, were determined to be 3.80 ± 0.03, 6.91 ± 0.07, and 8.9 ± 0.4, respectively. Stability constants for the ML complexes, the prime focus of this work, were thus established with an uncertainty smaller than 0.05 log units. The observed departure from electrochemical reversibility for the Cd–sarcosine–OH system was attributed mainly to the decrease in the transfer coefficient . The MM2 force field, supplemented by additional parameters, reproduced the reported crystal structures of diaqua-bis(glycinato-O,N)nickel(II) and fac-tri(glycinato)-nickelate(II) very well. These parameters were used to predict structures of all possible isomers of (i) [Ni(H2O)4(gly)]+ and [Ni(H2O)4(sar)]+; and (ii) [Ni(H2O)3(IDA)] and [Ni(H2O)3(MIDA)] (IDA = iminodiacetic acid, MIDA = N-methyl iminodiacetic acid) by molecular mechanics/simulated annealing methods. The change in strain energy, ΔUstr, that accompanies the substitution of one ligand by another (ML + L′ → ML′ + L), was computed and a strain energy ΔUstr = +0.28 kcal mol−1 for the reaction [Ni(H2O)4(gly)]+ + sar → [Ni(H2O)4(sar)]+ + gly was found. This predicts the monoglycine complex to be marginally more stable. By contrast, for the reaction [Ni(H2O)3IDA] + MIDA → [Ni(H2O)3MIDA] + IDA, ΔUstr = −0.64 kcal mol−1, and the monoMIDA complex is predicted to be more stable. This correlates well with (i) stability constants for Cd–gly and Cd–sar reported here; and (ii) known stability constants of ML complex for glycine, sarcosine, IDA, and MIDA.  相似文献   

20.
New compounds of phthalic acid, Cs(HPHT), and terephthalic acid, Cs2(TPA), have been synthesized. The enthalpy of solution of Cs(HPHT) in water was determined and combined with the standard molar enthalpies of formation of CsOH(aq), H2O(l) and phthalic acid(s) to calculate the standard molar enthalpy of formation of Cs(HPHT) of −(1035.6 ± 0.5) kJ mol−1. The enthalpies of solution of Cs2(TPA) and TPA in approximately 0.11 mol dm−3 CsOH were determined and combined with the standard molar enthalpies of formation of TPA(s), H2O(l) and CsOH(aq, 1:500) to calculate the standard molar enthalpy of formation of Cs2(TPA) of −(1266.2 ± 0.3) kJ mol−1.  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