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.
Thermodynamic properties of γ-aminobutyric acid were studied in the temperature interval from 5.7 to 300 K using a vacuum adiabatic calorimeter. The curve C p (T) in the mentioned temperature interval is S-shaped without any anomalies. Based on the smoothed values of heat capacity, the calorimetric entropy $ S_{m}^{0} (T) - S_{m}^{0} (0) $ and the difference in the enthalpies $ H_{m}^{0} (T) - H_{m}^{0} (0) $ were calculated and tabulated. At the standard temperature 298.15 K, these values are equal to 158.1 ± 0.3 J K?1 mol?1 and 23020 ± 50 J mol?1, respectively. At temperatures from 5 to 10 K, the function C p (T) was found to obey the Debye law C = AT 3. Contrary to what has been supposed previously, the empirical Parks–Huffman rule for estimating entropy in the homologous series was shown to be not valid for the series glycine–β-alanine–γ-aminobutyric acid.  相似文献   

2.
Thermal and thermochemical investigations of natural hydroxyl-bearing copper sulfate Cu3SO4(OH)4??antlerite have been carried out. The stages of its thermal decomposition have been studied employing the Fourier-transform IR spectroscopy. The enthalpy of formation of antlerite from the elements ??f H m o (298.15?K)?=?(?1750?±?10)?kJ?mol?1 has been determined by the method of oxide melt solution calorimetry. Using value of S m o (298.15?K), equal to (263.46?±?0.47)?J?K?1?mol?1, obtained earlier by the method of adiabatic calorimetry, the Gibbs energy value of ??f G m o (298.15?K)?=?(?1467?±?10)?kJ?mol?1 has been calculated.  相似文献   

3.
Potassium ferrate, K2FeO4(cr), has numerous promising environmental applications. An aqueous thermodynamic model applicable to high ionic strengths is essential for guiding its applications. In this study, a thermodynamic model is developed for the solubility of K2FeO4(cr) in aqueous alkali metal hydroxide solutions, from 283.15 to 333.15 K to high ionic strengths, up to saturation of KOH and NaOH, based on the Pitzer activity coefficient model for aqueous species. The solubility products for K2FeO4(cr) at infinite dilution in the temperature range from 283.15 to 333.15 K were obtained. Based on the thermodynamic solubility product of K2FeO4(cr) at 298.15 and its temperature dependence, in combination with thermodynamic properties for $ {\text{FeO}}_{4}^{2 - } $ FeO 4 2 ? and K+ from the literature, standard thermodynamic properties of K2FeO4(cr) at 298.15 K and 0.1 MPa (1 bar) are derived for the first time as follows: Δf G 0 = ?(896 ± 8) kJ·mol?1, Δf H 0 = ?(1026 ± 4) kJ·mol?1, and S 0 = (130 ± 17) J·mol?1·K?1. Using the above thermodynamic properties for K2FeO4(cr), the potential presence or preservation of K2FeO4(cr) in the Martian soils under the conditions relevant to Mars were quantitatively evaluated. Thermodynamic calculations pertaining to the Martian conditions indicate that the presence or preservation of K2FeO4(cr) as a strong oxidant in the Martian soils can be supported.  相似文献   

4.
In an effort to understand the reactions of antibiotics hydrolysis with metallo-β-lactamases (MβLs), the thermokinetic parameters of cefazolin hydrolysis with B1 subclass MβL CcrA from Bacteroides fragilis were determined by microcalorimetric method. The values of activation free energy $ \Updelta G_{ \ne }^{\theta } $ are 88.032 ± 0.038, 89.075 ± 0.025, 90.095 ± 0.034, and 91.261 ± 0.044 kJ mol?1 at 293.15, 298.15, 303.15, and 308.15 K, respectively, the activation enthalpy $ \Updelta H_{ \ne }^{\theta } $ is 25.278 ± 0.005 kJ mol?1, the activation entropy $ \Updelta S_{ \ne }^{\theta } $ is ?213.99 ± 0.14 J mol?1 K?1, the apparent activation energy E is 27.776 kJ mol?1, and the reaction order is 1.4. The results indicated that the cefazolin hydrolysis with CcrA is an exothermic and spontaneous reaction. An association between the thermokinetic and kinetic parameters was revealed, which is that the catalytic constant K cat increase with increase in $ \Updelta H_{ \ne }^{\theta } $ .  相似文献   

5.
We report molar heat capacities for racemic sec-butylcyclohexane (s-BCH) measured by the adiabatic calorimetric method within the temperature range from 14 to 200 K. In the crystalline state, we identified the presence of a second-order phase transition and a glass transition phenomenon, both originating from the same configurational degree of freedom. The phase and glass transition temperatures T trs and T g were determined to be (136.7 ± 1.0) and (100 ± 1) K, respectively. The entropy of phase transition was estimated to be Δtrs S m = 1.4 J K?1 mol?1 ≈ (1/4) R ln 2. The phase transition was judged to be of an order–disorder type on the basis of the fact that the glass transition occurred in the low-temperature heat-capacity tail. The entropy was interpreted to suggest that four molecules in the crystalline state constitute a unit for producing two distinct configurations in the coexistence of (d)- and (l)-s-BCH.  相似文献   

6.
The aqueous chemistry of phosphorus is dominated by P(V), which under typical environmental conditions (and depending on pH and concentration) can be present as the orthophosphate species H3PO 4 0 (aq),H2PO 4 ? (aq),HPO 4 2? (aq) or PO 4 3? (aq). Many divalent, trivalent and tetravalent metal ions form sparingly soluble orthophosphate phases that, depending on the solution pH and concentrations of phosphate and metal ions, can be solubility limiting phases. Geochemical and chemical engineering modeling of solubilities and speciation require comprehensive thermodynamic databases that include the standard thermodynamic properties for the aqueous species and solid compounds. The most widely used sources for standard thermodynamic properties are the NBS (now NIST) Tables (from 1982 and earlier, with a 1989 erratum) and the final CODATA evaluation (1989). However, a comparison of the reported enthalpies of formation and Gibbs energies of formation for key phosphate compounds and aqueous species, especially H2PO 4 ? (aq) and HPO 4 2? (aq), shows a systematic and nearly constant difference of 6.3 to 6.9 kJ?mol?1 per phosphorus atom between these two evaluations. The existing literature contains numerous studies (including major data summaries) that are based on one or the other of these evaluations. In this report we examine and identify the origin of this difference and conclude that the CODATA evaluation is more reliable. Values of the standard entropies of the H2PO 4 ? (aq) and HPO 4 2? (aq) ions at 298.15 K and p?° =1 bar were re-examined in the light of more recent information and data not considered in the CODATA review, and a slightly different value of S m o (H2PO 4 ? , aq, 298.15 K) = (90.6±1.5) J?K?1?mol?1 was obtained.  相似文献   

7.
Novel anilino-pyrimidine fungicides, pyrimethanil maleic salt, and pyrimethanil fumaric salt (C28H30N6O4) were synthesized by a chemical reaction of pyrimethanil with maleic acid/fumaric acid. The low-temperature heat capacities of the two compounds were measured with an adiabatic calorimeter from 80 to 350 K. The heat capacities of pyrimethanil fumaric salt are bigger than that of pyrimethanil maleic salt in the measurement temperature range. The thermodynamic function data relative to 298.15 K were calculated based on the heat capacity-fitted curves. The melting points, the molar enthalpies (Δfus H m), and entropies (Δfus S m) of fusion of pyrimethanil maleic salt and pyrimethanil fumaric salt were determined from their DSC curves. The values indicate that pyrimethanil fumaric salt was more thermostable than pyrimethanil maleic salt. The constant-volume energies of combustion (Δc U m o ) of pyrimethanil maleic salt and pyrimethanil fumaric salt were measured using an isoperibol oxygen bomb combustion calorimeter at T = (298.15 ± 0.001) K. From the Hess thermochemical cycle, the standard molar enthalpies of formation of the two compounds were derived and determined to be Δf H m o (pyrimethanil maleic salt) = ?459.3 ± 4.9 kJ mol?1 and Δf H m o (pyrimethanil fumaric salt) = ?557.2 ± 4.8 kJ mol?1, respectively. The results suggest that pyrimethanil fumaric salt is more chemically stable than pyrimethanil maleic salt.  相似文献   

8.
The kinetics of the oxidation of tris(2,2′-bipyridyl)iron(II) and tris(1,10-phenanthroline)iron(II) complexes ([Fe(LL)3]2+, LL = bipy, phen) by nitropentacyanocobaltate(III) complex [Co(CN)5NO2]3? was investigated in acidic aqueous solutions at ionic strength of I = 0.1 mol dm?3 (HCl/NaCl). The reactions were carried out at fixed acid concentration ([H+] = 0.01 mol dm?3) and the temperature maintained at 35.0 ± 0.1 °C. Spectroscopic evidence is presented for the protonated oxidant. Protonation constants of 360.43 and 563.82 dm3 mol?1 were obtained for the monoprotonated and diprotonated Co(III) complexes respectively. Electron transfer rates were generally faster for [Fe(bipy)3]2+ than [Fe(phen)3]2+. The redox complexes formed ion-pairs with the oxidant with increasing concentration of the oxidant over that of the reductant. Ion-pair constants for these reaction were 160.31 and 131.9 dm3 mol?1 for [Fe(bipy)3]2+ and [Fe(phen)3]2+, respectively. The activation parameters measured for these systems have values as follows: ?H (kJ K?1 mol?1) = +113.4 ± 0.4 and +119 ± 0.3; ?S (J K?1) = +107.6 ± 1.3 and 125.0 ± 1.6; ?G (kJ K?1) = +81 ± 0.4 and +82.4 ± 0.4; and E a (kJ mol?1) = 115.9 ± 0.5 and 122.3 ± 0.6 for LL = bipy and phen, respectively. Effect of added anions (Cl?, $ {\text{SO}}_{4}^{2 - } $ and $ {\text{ClO}}_{4}^{ - } $ ) on the systems showed decrease in the electron transfer rate constant. An outer-sphere mechanism is proposed for the reaction.  相似文献   

9.
A new class of surfactant–cobalt(III) complexes of the type trans-[Co(DH)2(OA)X], where DH = dimethylglyoxime, OA = octadecylamine, X = Cl?, Br?, I?, N3 ?, NO2 ?, SCN? or OA, were synthesized and characterized by physicochemical and spectroscopic methods. The critical micelle concentration (CMC) values of these surfactant–cobalt(III) complexes in ethanol solution were obtained by measuring absorption at ~250 nm. Specific conductivity data (at 303–313 K) served for the evaluation of the temperature-dependent CMC and the thermodynamics of micellization (ΔG m 0 , ΔH m 0 and ΔS m 0 ). Steady-state photolysis and cyclic voltammetry of the complexes were studied. The surfactant–cobalt(III) complexes were screened for their antibacterial and antifungal activities against various microorganisms.  相似文献   

10.
The product from reaction of samarium chloride hexahydrate with salicylic acid and Thioproline, [Sm(C7H5O3)2·(C4H6NO2S)]·2H2O, was synthesized and characterized by IR, elemental analysis, molar conductance, and thermogravimetric analysis. The standard molar enthalpies of solution of [SmCl3·6H2O(s)], [2C7H6O3(s)], [C4H7NO2S(s)] and [Sm(C7H5O3)2·(C4H7NO2S)·H2O(s)] in a mixed solvent of absolute ethyl alcohol, dimethyl sulfoxide(DMSO) and 3 mol L?1 HCl were determined by calorimetry to be Δs H m Φ [SmCl3 δ6H2O (s), 298.15 K]= ?46.68±0.15 kJ mol?1 Δs H m Φ [2C7H6O3 (s), 298.15 K]= 25.19±0.02 kJ mol?1, Δs H m Φ [C4H7NO2S (s), 298.15 K]=16.20±0.17 kJ mol?1 and Δs H m Φ [Sm(C7H5O3)2·(C4H6NO2S)]·2H2O (s), 298.15 K]= ?81.24±0.67 kJ mol?1. The enthalpy change of the reaction (1) $$ SmCl_3 \cdot 6H_2 O(s) + 2C_7 H_6 O_3 (s) + C_4 H_7 NO_2 S(s) = Sm(C_7 H_5 O_3 )_2 \cdot (C_4 H_6 NO_2 S) \cdot 2H_2 O(s) + 3HCl(g) + 4H_2 O(1) $$ was determined to be Δs H m Φ =123.45±0.71 kJ mol?1. From date in the literature, through Hess’ law, the standard molar enthalpy of formation of Sm(C7H5O3)2(C4H6NO2S)δ2H2O(s) was estimated to be Δs H m Φ [Sm(C7H5O3)2·(C4H6NO2S)]·2H2O(s), 298.15 K]= ?2912.03±3.10 kJ mol?1.  相似文献   

11.
A precision rotating-bomb combustion calorimeter (thermistor of which was constructed in the laboratory) was calibrated using benzoic acid with purity of 99.999 %. The combustion energy of phenanthroline monohydrate (phen·H2O) at 298.15 K was determined to be Δc U m θ  = ?(5,757.45 ± 2.53) kJ mol?1. Then, the standard enthalpy of combustion and the standard enthalpy of formation of phen·H2O were calculated to be Δc H m θ  = ?(5,759.93 ± 2.53) kJ mol?1 and Δf H m θ  = ?(391.34 ± 2.98) kJ mo1?1, respectively. Particularly, the effect of phen·H2O on growth and metabolism of Escherichia coli (E. coli) was also determined by a TAM air isothermal calorimeter at 37 °C. The thermokinetic parameters, including maximum heat output power (P max), growth rate constant (κ), generation times (t G), inhibitive rate (I), and half inhibition concentration (C I,50), were obtained. The results showed that phen·H2O possessed the bi-directional biological effect and Hormesis effect, which stimulated the growth of E. coli at lower concentration, but inhibited the growth at higher concentration. The half inhibition concentration C I,50 of phen·H2O was found to be 7.31 mg L?1.  相似文献   

12.
The study detailed in this paper is about the determination of the physical–chemical parameters of water, after keeping it in prolonged contact with the Nafion polymer. The parameters under study are: electrical conductivity, χ (μS cm?1); heat of mixing with acid (HCl), ΔQ mix HCl (J kg?1) or basic (NaOH) solutions, ΔQ mix NaOH  (J kg?1), and pH. χ increases of up to two orders of magnitude, ΔQ mix NaOH  (J kg?1) is exothermic and increases as the electrical conductivity increases, with a roughly linear trend, up to one order of magnitude. The analogous ΔQ mix HCl  (J kg?1), on the contrary, is found to be null. The pH is quite acid and shows a very good linear correlation with log χ. The linear correlations hint at a single cause for the variation of the three very different physical–chemical parameters. This complex and hard to rationalize phenomenology, finds a good theoretical support in the work hypothesis of the formation of dissipative structures within the liquid. These are far-from-equilibrium systems outside the paradigm of classical thermodynamics. The work hypothesis of the formation of molecular aggregates of water molecules (dissipative structures, aqueous nanostructures, clusters, coherence domains, etc.) is shared with two other aqueous systems obtained with different preparation protocols, so we briefly recall them here: (1) EDS (extremely diluted solutions): obtained through an iterative process of successive dilutions and agitations. (2) IFW (iteratively filtered water): obtained through an iterative process of successive filtrations through sintered glass filters. (3) INW (iteratively nafionized water): obtained through an iterative process of successive drying and wetting of the Nafion polymer. Each protocol produces water exhibiting its own peculiarities, to the point that they can be considered different, albeit with the common element of a variation of the super-molecular structure of the water solvent. The physical–chemical properties of these perturbed waters cannot be framed by the paradigm of classical thermodynamics, but rather require the use of the thermodynamics of systems far from the equilibrium and of irreversible processes.  相似文献   

13.
The major objective of this study, based on critical review and experimental studies, was to develop a reliable thermodynamic model for the Nd–F system at 25 °C. The SIT model was used to convert concentration constants reported in the literature to constants at zero ionic strengths for cross comparison and selection of reliable values. The critically evaluated thermodynamic constants for the formation of NdF2+ and NdF 2 + were then used to interpret the extensive NdF3(cr) solubility data in NaF and NH4F solutions, ranging in concentrations from extremely low values to as high as 1.0 mol·kg?1, equilibrated for different periods ranging up to as long as 72 days. These efforts have resulted in $ \log_{10} \beta_{n}^{0} $ log 10 β n 0 for the reaction [Nd3+ + nF? ? NdF n 3?n ] of (3.81 ± 0.10), (5.89 ± 0.77), and <12.48 for n values of 1–3, respectively. The $ \log_{10} K_{\text{sp}}^{0} $ log 10 K sp 0 for the solubility of NdF3(cr) (NdF3(cr) ? Nd3+ + 3F?) was determined to be (?20.49 ± 0.37). Because (1) Nd is an excellent analog for trivalent actinides—An(III) (i.e., Pu(III), Am(III), and Cm(III)), and (2) the available data for the An(III)–F system, especially the solubility products of AnF3(cr), are of extremely poor quality, the critical literature review in combination with the experimental Nd–F system data have been used to assign thermodynamic constants for the An(III)–F reactions until good quality specific data for them becomes available.  相似文献   

14.
The thermal and kinetic studies of epoxidized natural rubber (ENR) and its polymer electrolytes, LiX/ENR PEs, (where X = ClO 4 ? , CF3SO 3 ? , COOCF 3 ? , I?, and BF 4 ? ) were carried out using thermogravimetric analysis at different heating rates. The thermal behaviors for LiX/ENR PEs are closely related to the morphology and interactions between the LiX and ENR chains. The LiCF3SO3, LiCOOCF3, and LiI form pseudo-crosslinking within the ENR; their thermal behavior resembled purified ENR. The LiClO4 tends to form aggregates within the ENR. This phenomenon has promoted a much earlier decomposition of epoxide in the ENR. The occurrence of ring-opening and complexation or cross-linking reactions in and between the ENR chains in the LiBF4/ENR has produced a thermally stable macrostructure. The activation energy for the thermal degradation (E d) of purified ENR was 239.8 and 239.9 kJ mol?1 using Kissinger and FWO methods, respectively. According to the Coats–Redfern method, the degradation mechanism of purified ENR follows the F1 type model, while the Criado method revealed that the degradation starts with F1 followed by D3 type models. The E d for LiX/ENR (X = COOCF 3 ? , CF3SO 3 ? , I?, and BF 4 ? ) PE’s obtained via the Kissinger method are 258.5, 257.0, 251.0, and 198.9 kJ mol?1, respectively, and the corresponding E d values obtained by FWO are 236.0, 223.6, 349.7, and 206.6 kJ mol?1, respectively. The degradation of ENR in these PEs followed the D3 type model. However, for LiClO4/ENR, the presence of two distinct degradations of ENR gave two E d values. These are 174.5 and 234.7 kJ mol?1 using Kissinger and 117.8 and 293.6 kJ mol?1 using FWO method. The degradation mechanism of ENR in the LiClO4/ENR PE was similar to purified ENR that is F1 followed by D3 type models.  相似文献   

15.
Conductivities, densities and ultrasonic speeds measurements of hexadecyltrimethylammonium bromide (HTAB) in aqueous solutions of glycine (Gly) and glycylglycine (Gly-Gly) have been made at various temperatures. The critical micelle concentration (CMC), the degree of ionization (??) of the micelles, standard free energy, enthalpy, and entropy of the micellization process (??G m ° , ??H m ° , and ??S m ° ) for the present systems were estimated at different temperatures. The CMC values of HTAB in aqueous Gly and Gly-Gly were also evaluated by density and ultrasonic speed measurements. Apparent molar volumes, (V ?), apparent molar volumes at infinite dilution, (V ? ° ), apparent molar compressibilities, (K ?), of HTAB in the pre- and post-micellar regions, and volume change on micellization (??V ? m ) were also estimated. Large positive values of T??S m ° and small negative values of ??H m ° suggest that micellization process is driven primarily by entropy increase. The increase in ??V ? m and K ? with rise in temperature is indicative of less compact micellar structure of HTAB in presence of amino acid additives. These data suggest that amino acids are solubilised probably in the palisade layer of the micelle.  相似文献   

16.
The complexation of the diflunisal anion (DF) by γ-cyclodextrin (γCD) and permethylatedβ-cyclodextrin (βPCD) in aqueous solution at pH 7.00 at 298.2 K, has been studied by UV-visible and19F NMR spectroscopy. The formation of 1∶1 and 1∶2 γCD inclusion complexes proceeds through the two equilibria: (K1) $${\text{DF + }}\gamma {\text{CD}} \rightleftharpoons {\text{DF}} \cdot \gamma {\text{CD}}$$ (K2) $${\text{DF}} \cdot \gamma {\text{CD + }}\gamma {\text{CD }} \rightleftharpoons {\text{ DF}} \cdot {\text{(}}\gamma {\text{CD)}}_{\text{2}} {\text{ }}$$ characterised byK 1=(5.5±0.2)×104 dm3 mol?1 andK 2=(2.3±0.2)×104 dm3 mol?1 derived from UV-visible spectrophotometric data. The analogous βPCD complexes are characterised byK 1=(6.86±0.02)×104 dm3 mol?1 andK 2=(8.75±2.7)×101 dm3 mol?1. The variation of the19F chemical shift of DF on inclusion is consistent with the formation of 1∶1 and 1∶2 complexes also. Comparisons with related systems are made.  相似文献   

17.
The heat capacities of Ln(Me2dtc)3(C12H8N2) (Ln = La, Pr, Nd, Sm, Me2dtc = dimethyldithiocarbamate) have been measured by the adiabatic method within the temperature range 78–404 K. The temperature dependencies of the heat capacities, C p,m [La(Me2dtc)3(C12H8N2)] = 542.097 + 229.576 X ? 27.169 X 2 + 14.596 X 3 ? 7.135 X 4 (J K?1 mol?1), C p,m [Pr(Me2dtc)3(C12H8N2)] = 500.252 + 314.114 X ? 17.596 X 2 ? 0.131 X 3 + 16.627 X 4 (J K?1 mol?1), C p,m [Nd(Me2dtc)3(C12H8N2)] = 543.586 + 213.876 X ? 68.040 X 2 + 1.173 X 3 + 2.563 X 4 (J K?1 mol?1) and C p,m [Sm(Me2dtc)3(C12H8N2)] = 528.650 + 216.408 X ? 16.492 X 2 + 12.076 X 3 + 4.912 X 4 (J K?1 mol?1), were derived by the least-squares method from the experimental data. The heat capacities of Ce(Me2dtc)3(C12H8N2) and Pm(Me2dtc)3(C12H8N2) at 298.15 K were evaluated to be 617.99 and 610.09 J K?1 mol?1, respectively. Furthermore, the thermodynamic functions (entropy, enthalpy and Gibbs free energy) have been calculated using the obtained experimental heat capacity data.  相似文献   

18.
The conformational preference of the methyl group of 1-methyl-1-germacyclohexane was studied experimentally in solution (low-temperature 13C NMR) and by quantum chemical calculations (CCSD(T), MP2 and DFT methods). The NMR experiment resulted in an axial/equatorial ratio of 44/56 mol% at 114 K corresponding to an A value (A = G ax G eq) of 0.06 kcal mol?1. An average value for ΔG e→a #  = 5.0 ± 0.1 kcal mol?1 was obtained for the temperature range 106–134 K. The experimental results are very well reproduced by the calculations. CCSD(T)/CBS calculations + thermal corrections resulted in an A value of 0.02 kcal mol?1, whereas a ΔE value of ?0.01 kcal mol?1 at 0 K was obtained.  相似文献   

19.
A density-functional theory investigation on the interactions between C2H radical and small gold clusters Au n 0/? (n = 1–4) has been performed. The calculated results predict that C2H radical inclines to interact with small gold clusters Au n 0/? (n = 1–4) as an integrity in the most stable structures of C2HAu n 0/? (n = 1–4). The Au n 0/? (n = 1–4) clusters retain their structural integrity as units in the ground states of C2HAu n 0/? (n = 1–4). The stretching vibrational frequencies of C≡C and C–H in the ground states of C2HAu n ? (n = 1–4) are decreased compared with those of the C2H radical due to the interaction between the Au n 0/? clusters and C2H radical. Smaller red shifts in the C≡C and C–H stretching bands of C2HAu n ? occur with an increase in n. The photoelectron spectra of the most stable structures of C2HAu n ? (n = 1–4) have been simulated to aid their future experimental characterizations. The current study provides further insight into the interaction between C2H radicals and gold clusters, which may lead to exploitation of the high activity of gold nanocrystals.  相似文献   

20.
In an effort to probe the reaction of antibiotic hydrolysis catalyzed by B3 metallo-??-lactamase (M??L), the thermodynamic parameters of penicillin G hydrolysis catalyzed by M??L L1 from Stenotrophomonas maltophilia were determined by microcalorimetric method. The values of activation free energy ??G ?? ?? are 88.26, 89.44, 90.49, and 91.57?kJ?mol?1 at 293.15, 298.15, 303.15, and 308.15?K, respectively, activation enthalpy ??H ?? ?? is 24.02?kJ?mol?1, activation entropy ??S ?? ?? is ?219.2511?J?mol?1?K?1, apparent activation energy E is 26.5183?kJ?mol?1, and the reaction order is 1.0. The thermodynamic parameters reveal that the penicillin G hydrolysis catalyzed by M??L L1 is an exothermic and spontaneous reaction.  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