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 109 毫秒
1.
《Tetrahedron: Asymmetry》2005,16(21):3512-3519
Stereoselective reductions of prochiral ketones were performed using a new thermophilic, NAD-dependent alcohol dehydrogenase from Thermus sp. (TADH). The enzyme was produced on 2L-scale from recombinant Escherichia coli and purified by a simple, one-step heat treatment procedure yielding 220 mg of pure enzyme. Regeneration of NADH was catalyzed by the organometallic complex [Cp*Rh(bpy)(H2O)]2+ using formate as a reducing agent. The catalytic performance of [Cp*Rh(bpy)(H2O)]2+ in terms of total number of catalytic cycles and number of catalytic cycles per hour achieved herein (up to 1500 and more than 400 h−1, respectively), are the highest reported for a non-enzymatic nicotinamide regeneration system so far. Chemoenzymatic reduction reactions in a two liquid phase setup were performed on a gramme-scale, for example, 1.3 g of enantiopure (1S,3S)-3-methylcyclohexanol was obtained after purification. The volumetric productivity reached up to 3.9 mM h−1 with an average of 2.6 mM h−1 (5.3 g L−1 d−1) over 10 h. In addition, chemoenzymatic oxidations utilizing the same catalyst set and molecular oxygen as a terminal electron acceptor were performed. Thus, the preparative value of chemoenzymatic transfer hydrogenations with [Cp*Rh(bpy)(H2O)]2+ as a regeneration catalyst coupled especially to thermophilic ADHs was demonstrated.  相似文献   

2.
High methanol electro-oxidation activity was obtained on novel PtRuFe/C (2:1:1 at.%) catalyst. Mass and specific activities were 5.67 A  g−1 catal. and 177 mA m−2 for the PtRuFe/C catalyst while those of the commercial PtRu/C catalyst were 2.28 A g−1 catal. and 87.7 mA m−2, respectively. CO stripping results showed that on-set voltage for CO electro-oxidation was lowered by incorporation of Fe. XRD and XPS results revealed that Fe2O3 was formed instead of Fe(0), which resulted in large electron deficiency in Pt and easy CO electro-oxidation. The electron deficiency of Pt was proved by XPS results of Pt4f peaks, which moved to higher binding energies in PtRuFe/C than PtRu/C.  相似文献   

3.
We show a great possibility of mediated enzymatic bioelectrocatalysis in the formate oxidation and the carbon dioxide (CO2) reduction at high current densities and low overpotentials. Tungsten-containing formate dehydrogenase (FoDH1) from Methylobacterium extorquens AM1 was used as a catalyst and immobilized on a Ketjen Black-modified electrode. For the formate oxidation, a high limiting current density (jlim) of ca. 24 mA cm 2 was realized with a half wave potential (E1/2) of only 0.12 V more positive than the formal potential of the formate/CO2 couple (E°′CO2) at 30 °C in the presence of methyl viologen (MV2 +) as a mediator, and jlim reached ca. 145 mA cm 2 at 60 °C. Even when a viologen-functionalized polymer was co-immobilized with FoDH1 on the porous electrode, jlim of ca. 30 mA cm 2 was attained at 60 °C with E1/2 = E°′CO2 + 0.13 V. On the other hand, the CO2 reduction was also realized with jlim  15 mA cm 2 and E1/2 = E°′CO2  0.04 V at pH 6.6 and 60 °C in the presence of MV2 +.  相似文献   

4.
The samples of dibarium magnesium orthoborate Ba2Mg(BO3)2 were synthesized by solid-state reaction. The X-ray diffraction (XRD) patterns and Raman spectra of the samples were collected. Electronic structure and vibrational spectroscopy of Ba2Mg(BO3)2 were systematically investigated by first principle calculation. A direct band gap of 4.4 eV was obtained from the calculated electronic structure results. The top valence band is constructed from O 2p states and the low conduction band mainly consists of Ba 5d states. Raman spectra for Ba2Mg(BO3)2 polycrystalline were obtained at ambient temperature. The factor group analysis results show the total lattice modes are 5Eu + 4A2u + 5Eg + 4A1g + 1A2g + 1A1u, of which 5Eg + 4A1g are Raman-active. Furthermore, we obtained the Raman active vibrational modes as well as their eigenfrequencies using first-principle calculation. With the assistance of the first-principle calculation and factor group analysis results, Raman bands of Ba2Mg(BO3)2 were assigned as Eg (42 cm−1), A1g (85 cm−1), Eg (156 cm−1), Eg (237 cm−1), A1g (286 cm−1), Eg (564 cm−1), A1g (761 cm−1), A1g (909 cm−1), Eg (1165 cm−1). The strongest band at 928 cm−1 in the experimental spectrum is assigned to totally symmetric stretching mode of the BO3 units.  相似文献   

5.
The reaction of [Cp1IrCl2]2 (Cp* = η5  C5Me5) with the tridentate 3-thiapentane-1,5-dithiolate ligand, S(CH2CH2S)2 (tpdt), led to the formation of [Cp1Ir(η3  tpdt)] (1) in 81% isolated yield. Subsequent reactions of 1 with [Cp1IrCl2]2 in 2:1 and 1:1 molar equiv ratios resulted in the formation of [Cp1Ir(μ  η2:η3  tpdt)Cp1IrCl][PF6] (2) and [Cp1Irμ  η2:η3  tpdt)Cp1IrCl][Cp1IrCl3] (3) in 86 and 79% yields, respectively, based on 1, whereas the reactions of 1 with [(COD)IrCl]2 (COD = 1,5-cyclooctadiene) in 2:1 and 1:1 molar equiv ratios resulted in the formation of the homo-bimetallic derivatives Cp1Ir(μ  η1:η3  tpdt)(COD)IrCl (4) (92% yield) and [Cp1Ir(μ  η2:η3  tpdt)(COD)Ir] [(COD)IrCl2] (5) (82% yield). Reactions between 1 and [(COD)RhCl]2, yielded the hetero-bimetallic derivatives Cp1Ir(μ  η1:η3  tpdt)(COD)RhCl (6) and [Cp1Ir(μ  η2:η3  tpdt)(COD)Rh][(COD)RhCl2] (7), in 92 and 93% yields, respectively. The reaction of 1 with methyl iodide gave mono-methylated derivative [Cp1Ir(η3-C4H8S3Me)]I (8) (93% yield). All these compounds have been comprehensively characterized.  相似文献   

6.
We investigate the nature of bonding and charge states in (U1−yCey)O2 (y = 0.0, 0.2, 0.4, 0.6, 0.8 and 1.0) by Raman spectroscopy. Raman spectrum of UO2 exhibits two prominent bands below 1000 cm−1, a F2g mode at 446 cm−1 and a F1u LO mode at 578 cm−1. As y is increased from 0 to 0.6, the F1u exhibits a large blue shift of 90 cm−1, and from y = 0.6 to 1.0, a red shift of 54 cm−1. We show that our results can be interpreted as arising from anisotropic compression/relaxation of the lattice under Ce substitution and this can give an indication of its charge states. Alternate interpretations have been given in the literature on the effect of substituents and dopants to the Raman spectra of UO2 and CeO2. The present interpretation of chemical stress effects can be taken as another plausible explanation.  相似文献   

7.
Photodegradation of 4-nitrophenol (4-Np) in the presence of zinc tetrasulfophthalocyanine (ZnPcS4), zinc octacarboxyphthalocyanine (ZnPc(COOH)8) and a sulfonated ZnPc containing a mixture of differently sulfonated derivatives (ZnPcSmix), as photocatalysts is reported. ZnPcSmix is the most effective catalyst in terms of a high quantum yield for 4-Np degradation and the stability of the catalyst. However ZnPc(COOH)8 degrades readily during the catalysis, but it has a higher quantum yield (Φ4-Np) for 4-Np degradation than the rest of the complexes. The Φ4-Np values were closely related to the singlet oxygen quantum yields ΦΔ and hence aggregation. The rate constants for the reaction with 4-Np were kr = 0.67 × 106 mol−1 dm3 s−1 for ZnPcSmix and 2.8 × 108 mol−1 dm3 s−1 for ZnPc(COOH)8.  相似文献   

8.
A perovskite-type oxide of Ba0.5Sr0.5Co0.8Fe0.2O3−δ (BSCFO) with mixed electronic and oxygen ionic conductivity at high temperatures was used as an oxygen-permeable membrane. A tubular membrane of BSCFO made by extrusion method has been used in the membrane reactor to exclusively transport oxygen for the partial oxidation of ethane (POE) to syngas with catalyst of LiLaNiO/γ-Al2O3 at temperatures of 800–900 °C. After only 30 min POE reaction in the membrane reactor, the oxygen permeation flux reached at 8.2 ml cm−2 min−1. After that, the oxygen permeation flux increased slowly and it took 12 h to reach at 11.0 ml cm−2 min−1. SEM and EDS analysis showed that Sr and Ba segregations occurred on the used membrane surface exposed to air while Co slightly enriched on the membrane surface exposed to ethane. The oxygen permeation flux increased with increasing of concentration of C2H6, which was attributed to increasing of the driving force resulting from the more reducing conditions produced with an increase of concentration of C2H6 in the feed gas. The tubular membrane reactor was successfully operated for POE reaction at 875 °C for more than 100 h without failure, with ethane conversion of ∼100%, CO selectivity of >91% and oxygen permeation fluxes of 10–11 ml cm−2 min−1.  相似文献   

9.
A systematic study on the electrocatalytic properties of Pt nanoparticles supported on nitrobenzene-modified graphene (Pt-NB/G) as catalyst for oxygen reduction reaction (ORR) in alkaline solution was performed. Graphene nanosheets were spontaneously grafted with nitrophenyl groups using 4-nitrobenzenediazonium salt. The electrocatalytic activity towards the ORR and stability of the prepared catalysts in 0.1 M KOH solution have been studied and compared with that of the commercial Pt/C catalyst. The results obtained show that the NB-modified graphene nanosheets can be good Pt catalyst support with high stability and excellent electrocatalytic properties. The specific activity of Pt-NB/G for O2 reduction was 0.184 mA cm−2, which is very close to that obtained for commercial 20 wt% Pt/C catalyst (0.214 mA cm−2) at 0.9 V vs. RHE. The Pt-NB/G hybrid material promotes a four-electron reduction of oxygen and can be used as a promising cathode catalyst in alkaline fuel cells.  相似文献   

10.
The molar enthalpies of reaction of metallic barium with 0.047 mol·dm−3 HClO4 as well as the molar enthalpies of dissolution of BaCl2 in 1.01 mol·dm−3 HCl and in water have been measured at T=298.15 K in a sealed swinging calorimeter with an isothermal jacket. From these results the standard molar enthalpy of formation of the barium ion in an aqueous solution at infinite dilution, as well as the enthalpies of formation of barium chloride and barium perchlorate, are calculated to be: ΔfH0m(Ba2+,aq)=−(535.83±1.25) kJ · mol−1; ΔfH0m(BaCl2,cr)=−(855.66±1.28) kJ · mol−1; and ΔfH0m(BaClO4,cr)=−(796.26±1.35) kJ · mol−1. The results obtained are discussed and compared with previous experimental values.  相似文献   

11.
12.
Standard values of Gibbs free energy, entropy, and enthalpy of Na2Ti6O13 and Na2Ti3O7 were determined by evaluating emf-measurements of thermodynamically defined solid state electrochemical cells based on a Na–β″-alumina electrolyte. The central part of the anodic half cell consisted of Na2CO3, while two appropriate coexisting phases of the ternary system Na–Ti–O are used as cathodic materials. The cell was placed in an atmosphere containing CO2 and O2. By combining the results of emf-measurements in the temperature range of 573⩽T/K⩽1023 and of adiabatic calorimetric measurements of the heat capacities in the low-temperature region 15⩽T/K⩽300, the thermodynamic data were determined for a wide temperature range of 15⩽T/K⩽1100. The standard molar enthalpy of formation and standard molar entropy at T=298.15 K as determined by emf-measurements are ΔfHm0=(−6277.9±6.5) kJ · mol−1 and Sm0=(404.6±5.3) J · mol−1 · K−1 for Na2Ti6O13 and ΔfHm0=(−3459.2±3.8) kJ · mol−1 and Sm0=(227.8±3.7) J · mol−1 · K−1 for Na2Ti3O7. The standard molar entropy at T=298.15 K obtained from low-temperature calorimetry is Sm0=399.7 J · mol−1 · K−1 and Sm0=229.4 J · mol−1 · K−1 for Na2Ti6O13 and Na2Ti3O7, respectively. The phase widths with respect to Na2O content were studied by using a Na2O-titration technique.  相似文献   

13.
Low-temperature calorimetric measurements have been performed on DyBr3(s) in the temperature range (5.5 to 420 K ) and on DyI3(s) from T=4 K to T=420 K. The data reveal enhanced heat capacities below T=10 K, consisting of a magnetic and an electronic contribution. From the experimental data on DyBr3(s) a C0p,m (298.15 K) of (102.2±0.2) J·K−1·mol−1 and a value for {S0m (298.15 K)  S0m (5.5 K)} of (205.5±0.5) J·K−1·mol−1, have been obtained. For DyI3(s), {S0m (298.15 K)  S0m (4 K)} and C0p,m (298.15 K) have been determined as (226.9±0.5) J·K−1·mol−1 and (103.4±0.2) J·K−1·mol−1, respectively. The values for {S0m (5.5 K)  S0m (0)} for DyBr3(s) and {S0m (4 K)  S0m (0)} for DyI3(s) have been calculated, giving S0m (298.15 K)=(212.3±0.9) J·K−1·mol−1 in case of DyBr3(s) and S0m (298.15 K) =(233.1±0.7) J·K−1·mol−1 for DyI3(s). The high-temperature enthalpy increment has been measured for DyBr3(s) in the temperature range (525 to 799 K) and for DyI3(s) in the temperature range (525 to 627 K). From the results obtained and enthalpies of formation from the literature, thermodynamic functions for DyBr3(s) and DyI3(s) have been calculated from T→0 to their melting temperatures at 1151.0 K and 1251.5 K, respectively.  相似文献   

14.
The paper reports the operation of a new-design microbial fuel cell using compost leachate as a substrate, oxygen/electrodeposited MnOx cathode and a new-anode concept with graphite modified by an iron/sulfur solid chemical catalyst which almost eliminates the starting delay time and gives very high current and power densities, I ~ 25 A m 3 at Pmax ~ 12 W m 3 or I ~ 3.8 A m 2 at Pmax ~ 1.8 W m 2.  相似文献   

15.
In this study, Nb2O5 nanobelts, with a ca. ∼15 nm in thickness, ca. ∼60 nm in width and several tens of mircrometers in length, have first been used as the electrode material for lithium intercalation over the potential window of 3.0–1.2 V (vs. Li+/Li). It delivers an initial intercalation capacity of 250 mA hg−1 at 0.1 Ag−1 current density, corresponding to x = 2.5 for LxNb2O5, and can still keep relative stable and reaches as large as 180 mA hg−1 after 50 cycles. Surprisingly, the electrodes composed of Nb2O5 nanobelts can work smoothly even at high current density of 10 Ag−1, and shows higher specific capacity and excellent cycling stable, as well as sloped feature in voltage profile. Cycling test indicates Nb2O5 nanobelts electrode shows a high reversible charge/discharge capacity, high rate capability with excellent cycling stability.  相似文献   

16.
《Chemical physics letters》2006,417(1-3):206-210
Two C–O stretching hot bands, (ν1 + 2ν3)  2ν3 and (2ν1 + ν3)  (ν1 + ν3), of the CCO radical in the ground electronic state were measured. These hot bands are red shifted by approximately 70 cm−1 compared to the C–O stretching fundamental. CCO was produced in a discharge through a flowing mixture of carbon suboxide and helium. The spectra were recorded using a diode laser spectrometer. The band origins were determined to be 1904.32512(62) and 1902.69130(56) cm−1 for (ν1 + 2ν3)  2ν3 and (2ν1 + ν3)  (ν1 + ν3), respectively. The measurements in this band together with previously reported frequencies in the C–C and C–O stretching regions were analysed to determine harmonic frequencies and anharmonicity constants.  相似文献   

17.
The kinetics of the diazotization reaction of procaine in the presence of anionic micelles of sodium dodecyl sulfate (SDS) and cationic micelles of cetyltrimethyl ammonium bromide (CTAB), dodecyltrimethyl ammonium bromide (DDTAB) and tetradecyltrimethyl ammonium bromide (TDTAB) were carried out spectrophotometrically at λmax = 289 nm. The values of the pseudo first order rate constant were found to be linearly dependent upon the [NaNO2] in the concentration range of 1.0 × 10−3 mol dm−3 to 12.0 × 10−3 mol dm−3 in the presence of 2.0 × 10−2 mol dm−3 acetic acid. The concentration of procaine was kept constant at 6.50 × 10−5 mol dm−3. The addition of the cationic surfactants increased the reaction rate and gave plateau like curve. The addition of SDS micelles to the reactants initially increased the rate of reaction and gave maximum like curve. The maximum value of the rate constant was found to be 9.44 × 10−3 s−1 at 2.00 × 10−3 mol dm−3 SDS concentration. The azo coupling of diazonium ion with β-naphthol (at λmax = 488) nm was found to linearly dependent upon [ProcN2+] in the presence of both the cationic micelles (CTAB, DDTAB and TDTAB) and anionic micelles (SDS). Both the cationic and anionic micelles inhibited the rate of reactions. The kinetic results in the presence of micelles are explained using the Berezin pseudophase model. This model was also used to determine the kinetic parameters e.g. km, Ks from the observed results of the variation of rate constant at different [surfactants].  相似文献   

18.
The stacks of multi-layer Ti3C2Tx and other types of MXene materials limit their electrochemical performance. Herein, we report a facile exfoliation technique to improve the exfoliation efficiency through Li-intercalation into Ti3C2Tx interlayers in isopropyl alcohol (IPA) with LiOH as intercalant. This de-intercalation method presented here not only effectively delaminates the stacked Ti3C2Tx multi-layers into separate few-layer MXene sheets, but also achieves high-rate supercapacitive performance of Ti3C2Tx electrode. The as-produced delaminated Ti3C2Tx shows highly improved electrochemical capacitive properties from 47 to 115 F g 1 at 200 mV s 1. Even at extremely high scan rate of 1000 mV s 1, a specific capacitance of 82 F g 1 is still obtained. The high-rate capability can be attributed to improved ions accessibility into the few-layer structures. This study offers a new and simple exfoliation pathway for MXenes materials to exploit their full potential in energy storage applications.  相似文献   

19.
We present a simple method for measuring Henry’s constant kHof ethanol using photoacoustic spectroscopy. At T =  298.1 K the measured value forkH is (0.877  ±  0.039)kPa · kg · mol  1. Our data show that Henry’s law is valid at ethanol molalities between 0.1mol · kg  1 and 1.4 mol · kg  1. The temperature dependence of Henry’s constant was carefully examined by measuring the ethanol vapour pressure of six different aqueous solutions between T =  273.1 K and T =  298.1 K. By analysing the gas phase concentration and applying Henry’s law, an ethanol molality of 0.864 mol · kg  1in the liquid phase can be measured with an error of  ± 0.038mol · kg  1. The detection limit of the photoacoustic sensor is a gaseous ethanol pressure of 10  3kPa. Ethanol molality changes as low as 1.10  3mol · kg  1can be measured.  相似文献   

20.
The molar heat capacities of GeCo2O4 and GeNi2O4, two geometrically frustrated spinels, have been measured in the temperature range from T=(0.5 to 400) K. Anomalies associated with magnetic ordering occur in the heat capacities of both compounds. The transition in GeCo2O4 occurs at T=20.6 K while two peaks are found in the heat capacity of GeNi2O4, both within the narrow temperature range between 11.4<(T/K)<12.2. Thermodynamic functions have been generated from smoothed fits of the experimental results. At T=298.15 K the standard molar heat capacities are (143.44 ± 0.14) J · K−1 · mol−1 for GeCo2O4 and (130.76 ± 0.13) J · K−1 · mol−1 for GeNi2O4. The standard molar entropies at T=298.15 K for GeCo2O4 and GeNi2O4 are (149.20 ± 0.60) J · K−1 · mol−1 and (131.80 ± 0.53) J · K−1 · mol−1 respectively. Above 100 K, the heat capacity of the cobalt compound is significantly higher than that of the nickel compound. The excess heat capacity can be reasonably modeled by the assumption of a Schottky contribution arising from the thermal excitation of electronic states associated with the CO2+ ion in a cubic crystal field. The splittings obtained, 230 cm−1 for the four-fold-degenerate first excited state and 610 cm−1 for the six-fold degenerate second excited state, are significantly lower than those observed in pure CoO.  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