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 503 毫秒
1.
We report the complete ethanolysis of Kraft lignin over an α‐MoC1?x/AC catalyst in pure ethanol at 280 °C to give high‐value chemicals of low molecular weight with a maximum overall yield of the 25 most abundant liquid products (LP25) of 1.64 g per gram of lignin. The LP25 products consisted of C6–C10 esters, alcohols, arenes, phenols, and benzyl alcohols with an overall heating value of 36.5 MJ kg?1. C6 alcohols and C8 esters predominated and accounted for 82 wt % of the LP25 products. No oligomers or char were formed in the process. With our catalyst, ethanol is the only effective solvent for the reaction. Supercritical ethanol on its own degrades Kraft lignin into a mixture of small molecules and molecular fragments of intermediate size with molecular weights in the range 700–1400, differing in steps of 58 units, which is the weight of the branched‐chain linkage C3H6O in lignin. Hydrogen was found to have a negative effect on the formation of the low‐molecular‐weight products.  相似文献   

2.
In this work, the effect of Fento’s reagent on the degradation of residual Kraft black liquor was investigated. The effect of Fenton’s reagent on the black liquor degradation was dependent on the concentration of H2O2. At low concentrations (5 and 15 mM) of H2O2, Fenton’s reagent caused the degradation of phenolic groups (6.8 and 44.8%, respectively), the reduction of reaction medium pH (18.2%), and the polymerization of black liquor lignin. At a high concentration (60 mM) of H2O2, Fenton’s reagent induced an extensive degradation of lignin (95–100%) and discoloration of the black liquor. In the presence of traces of iron, the addition of H2O2 alone induced mainly lignin fragmentation. In conclusion, Fenton’s reagent and H2O2 alone can degrade residual Kraft black liquor under acidic conditions at room temperature.  相似文献   

3.
Developing organic compounds with multifunctional groups to be used as electrode materials for rechargeable sodium‐ion batteries is very important. The organic tetrasodium salt of 2,5‐dihydroxyterephthalic acid (Na4DHTPA; Na4C8H2O6), which was prepared through a green one‐pot method, was investigated at potential windows of 1.6–2.8 V as the positive electrode or 0.1–1.8 V as the negative electrode (vs. Na+/Na), each delivering compatible and stable capacities of ca. 180 mAh g?1 with excellent cycling. A combination of electrochemical, spectroscopic and computational studies revealed that reversible uptake/removal of two Na+ ions is associated with the enolate groups at 1.6–2.8 V (Na2C8H2O6/Na4C8H2O6) and the carboxylate groups at 0.1–1.8 V (Na4C8H2O6/Na6C8H2O6). The use of Na4C8H2O6 as the initial active materials for both electrodes provided the first example of all‐organic rocking‐chair SIBs with an average operation voltage of 1.8 V and a practical energy density of about 65 Wh kg?1.  相似文献   

4.
《中国化学快报》2023,34(8):108090
Electrochemical oxidation of aqueous tris(1,3-dichloro-2-propyl) phosphate (TDCPP) by using Ti/SnO2-Sb/La-PbO2 as anode was investigated for the first time, and the degradation mechanisms and toxicity changes of the degradation intermediates were further determined. Results suggested that electrochemical degradation of TDCPP followed pseudo-first-order kinetics, and the reaction rate constant (k) was 0.0332 min−1 at the applied current density of 10 mA/cm2 and Na2SO4 concentration of 10 mmol/L. There was better TDCPP degradation performance at higher current density. Free hydroxy radical (OH) was proved to play dominant role in TDCPP oxidation via quenching experiment, with a relative contribution rate of 60.1%. A total of five intermediates (M1, C6H11Cl4O4P; M2, C3H7Cl2O4P; M3, C9H16Cl5O5P; M4, C9H14Cl5O6P; M5, C6H10Cl3O6P) were identified, and the intermediates were further degraded prolonging with the reaction time. Flow cytometer results suggested that the toxicity of TDCPP and degradation intermediates significantly reduced, and the detoxification efficiency was achieved at 78.1% at 180 min. ECOSAR predictive model was used to assess the relative toxicity of TDCPP and the degradation intermediates. The EC50 to green algae was 3.59 mg/L for TDCPP, and the values raised to 84, 574, 54.6, 391, and 8920 mg/L for M1, M2, M3, M4, and M5, respectively, indicating that the degradation intermediates are less toxic or not toxic. Electrochemical advanced oxidation process is a valid technology to degrade TDCPP and pose a good detoxification effect.  相似文献   

5.
All‐solid‐state sodium batteries (ASSSBs) with nonflammable electrolytes and ubiquitous sodium resource are a promising solution to the safety and cost concerns for lithium‐ion batteries. However, the intrinsic mismatch between low anodic decomposition potential of superionic sulfide electrolytes and high operating potentials of sodium‐ion cathodes leads to a volatile cathode–electrolyte interface and undesirable cell performance. Here we report a high‐capacity organic cathode, Na4C6O6, that is chemically and electrochemically compatible with sulfide electrolytes. A bulk‐type ASSSB shows high specific capacity (184 mAh g?1) and one of the highest specific energies (395 Wh kg?1) among intercalation compound‐based ASSSBs. The capacity retentions of 76 % after 100 cycles at 0.1 C and 70 % after 400 cycles at 0.2 C represent the record stability for ASSSBs. Additionally, Na4C6O6 functions as a capable anode material, enabling a symmetric all‐organic ASSSB with Na4C6O6 as both cathode and anode materials.  相似文献   

6.
The structure of the hemihydrate of sodium phenoxy­acetate, Na+·C8H7O3·0.5H2O, has been redetermined at low temperature (160 K). The structure consists of ribbons containing octahedral NaO6 units, and half of the Na2O2 squares within the ribbon are bridged by water mol­ecules which lie across twofold rotation axes in C2/c. The phenyl substituents lie on the outside of the ribbon.  相似文献   

7.
The title acid salt, Na+·C4H5O2?·2C4H6O2, contains finite anions in which two cyclo­propanoic acid mol­ecules are hydrogen bonded to a cyclo­propanoate residue. Each such anion interacts with four different Na+ cations.  相似文献   

8.
Sodium citrate dihydrate doped with Mn3+ ions, namely trisodium(I) managnese(III) citrate(3−) dihydrate, [Na3Mn0.011(C6H5O7)(H2O)2]n, was obtained during attempts to prepare some complex MnIII citrates from a concentrated strong alkaline solution containing Na+, Mn3+ and citrate ions. The compound is isostructural with the recently described Na3(C6H5O7)·2H2O [Fischer & Palladino (2003). Acta Cryst. E 59 , m1080–m1082]. The essential difference between these two structures is the presence of a very small proportion (0.205 wt%) of Mn3+ ions, which are positioned at the special 4e Wyckoff position in C2/c, where they are in a highly distorted octahedral environment of O atoms from two citrate anions.  相似文献   

9.
Thermal studies on various oxalato complexes have been of immense interest as they yield finely divided, highly reactive oxides which are usually obtained at a much lower temperature than that required in the conventional method of preparation, i.e., heating a mixture of two or more constituents [1]. A survey of the literature reveals that the compounds having the general formula A2[Mo2O5(C2O4)2(H2O)2], where A = K+, NH+4[2] and A = Cs+ [3], have been prepared and their thermal decomposition is studied, but no such information is available regarding the preparation and characterisation of Na2[Mo2O5(C2O4)2(H2O)2] (SMO), which forms the subject of study of this paper. Sodium dimolybdate (Na2Mo2O7), the decomposition product of SMO, is obtained at 280°C, a temperature much lower than that required in the conventional method of preparation of heating a mixture of Na2MoO4 and MoO3 [4].  相似文献   

10.
The preparations of some bisoxalatobisfluoroaluminates having the general formula M3[Al(C2O4)2F2.3H2O], where M=K+, Na+ and [Co(NH3)6]3+, and a bisoxalatobisfluorogallate, [Co(NH3)6] [Ga(C2O4)2F2].3H2O, are described. The compounds are characterised by chemical analyses, TGA, IR spectroscopy and X-ray powder photography. IR spectra support the presence of chelating oxalate ligands in these compounds. On isothermal heating at 100–130°C the compounds yield their respective anhydrous products.  相似文献   

11.
Two crystal samples, sodium 5-methylisophthalic acid monohydrate (C9H6O4Na2·H2O, s) and sodium isophthalic acid hemihydrate (C8H4O4Na2·1/2H2O, s), were prepared from water solution. Low-temperature heat capacities of the solid samples for sodium 5-methylisophthalic acid monohydrate (C9H6O4Na2·H2O, s) and sodium isophthalic acid hemihydrate (C8H4O4Na2·1/2H2O, s) were measured by a precision automated adiabatic calorimeter over the temperature range from 78 to 379 K. The experimental values of the molar heat capacities in the measured temperature region were fitted to a polynomial equation on molar heat capacities (C p,m) with the reduced temperatures (X), [X = f(T)], by a least-squares method. Thermodynamic functions of the compounds (C9H6O4Na2·H2O, s) and (C8H4O4Na2·1/2H2O, s) were calculated based on the fitted polynomial equation. The constant-volume energies of combustion of the compounds at T = 298.15 K were measured by a precise rotating-bomb combustion calorimeter to be Δc U(C9H6O4Na2·H2O, s) = −15428.49 ± 4.86 J g−1 and Δc U(C8H4O4Na2·1/2H2O, s) = −13484.25 ± 5.56 J g−1. The standard molar enthalpies of formation of the compounds were calculated to be Δ f H m θ (C9H6O4Na2·H2O, s) = −1458.740 ± 1.668 kJ mol−1 and Δ f H m θ (C8H4O4Na2·1/2H2O, s) = −2078.392 ± 1.605 kJ mol−1 in accordance with Hess’ law. The standard molar enthalpies of solution of the compounds, Δ sol H m θ (C9H6O4Na2·H2O, s) and Δ sol H m θ (C8H4O4Na2·1/2H2O, s), have been determined as being −11.917 ± 0.055 and −29.078 ± 0.069 kJ mol−1 by an RD496-2000 type microcalorimeter. In addition, the standard molar enthalpies of hydrated anion of the compounds were determined as being Δ f H m θ (C9H6O4 2−, aq) = −704.227 ± 1.674 kJ mol−1 and Δ f H m θ (C8H4O4Na2 2−, aq) = −1483.955 ± 1.612 kJ mol−1, from the standard molar enthalpies of solution and other auxiliary thermodynamic data through a thermochemical cycle.  相似文献   

12.
Upon collisional activation, gaseous metal adducts of lithium, sodium and potassium oxalate salts undergo an expulsion of CO2, followed by an ejection of CO to generate a product ion that retains all three metals atoms of the precursor. Spectra recorded even at very low collision energies (2 eV) showed peaks for a 44‐Da neutral fragment loss. Density functional theory calculations predicted that the ejection of CO2 requires less energy than an expulsion of a Na+ and that the [Na3CO2]+ product ion formed in this way bears a planar geometry. Furthermore, spectra of [Na3C2O4]+ and [39K3C2O4]+ recorded at higher collision energies showed additional peaks at m/z 90 and m/z 122 for the radical cations [Na2CO2]+? and [K2CO2]+?, respectively, which represented a loss of an M? from the precursor ions. Moreover, [Na3CO2]+, [39K3CO2]+ and [Li3CO2]+ ions also undergo a CO loss to form [M3O]+. Furthermore, product‐ion spectra for [Na3C2O4]+ and [39K3C2O4]+ recorded at low collision energies showed an unexpected peak at m/z 63 for [Na2OH]+ and m/z 95 for [39K2OH]+, respectively. An additional peak observed at m/z 65 for [Na218OH] + in the spectrum recorded for [Na3C2O4]+, after the addition of some H218O to the collision gas, confirmed that the [Na2OH] + ion is formed by an ion–molecule reaction with residual water in the collision cell. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

13.
Enzymatic lignin activation may be an environmentally friendly alternative to the use of chemicals in the production of wood fibers composites. Most studies on enzymatic activation of lignin for improving the adhesion of lignocellulosic products have been carried out using laccases. In this work, the use of a versatile peroxidase (VP) from the white-rot fungus Bjerkandera sp. (anamorph R1) for activating Kraft lignin was studied. The effect of enzyme dosage, incubation time, and H2O2 addition profile on lignin activation was evaluated by quantifying the phenoxy radicals formed using electron paramagnetic resonance (EPR) spectroscopy. Two alternative enzymatic systems based on the use of VP (a two-stage and an enzymatic cascade system) were also assayed. At optimal conditions (dose of 15 U?g?1 and continuous addition of H2O2 (5.24 μmol?h?1) during 1 h) the content of phenoxy radicals was doubled as compared with an untreated control. Moreover, using the two-stage VP system, a lignin activation similar to that found at optimal conditions could be reached in a shorter time.  相似文献   

14.
Polyol Metal Complexes. XIII. Na2[Be(C4H6O3)2] · 5H2O and Na2[Pb(C4H6O3)2] · 3H2O – Two Homoleptic Bis Polyolato Metallates with Beryllium and with Lead Na2[Be(C4H6O3)2] · 5H2O ( 1 ) and Na2[Pb(C4H6O3)2] · 3H2O ( 2 ) crystallize from concentrated, alkaline aqueous solutions. The polyol anhydroerythritol is deprotonated twice in the mononuclear, homoleptic complex anions. The preference of beryllium for the binding of cis-furanoid diols is shown. In 2 , a stereochemically active lone pair at the central atom is the reason for the construction of low dimensional aggregates from three plumbate and three sodium ions.  相似文献   

15.
The preparation of the η4-4-2,3,5,6-tetramethyl-1,4-benzoquinonecomplex [CO(C5Me5)(C10H12O2)] (I) is reported. Complex I undergoesreversible protonation to yield the 2-6-η-4-hydroxy-1-oxo-2,3,5,6-tetramethylcyclohexadienyl complex [Co(C5Me5)(C10H13O2)BF4 (II) and diprotonation to yield the η6-6-1,4-dihydroxy-2,3,5,6-tetramethylbenzene complex [Co(C5Me5)(C10H14O2)] (BF4)2 (III). Methylation of complex I with MeI/AgPF6 gives the 26-η-4-methoxy-1-oxo-2,3,5,6-tetramethylcyclohexadienyl complex [Co(C5Me5)(C11H15O2])PF6 (IV). In trifluoroacetic acid solution complex IV is protonated to form the η6-1-hydroxy-4-methoxy-2,3,5,6-tetramethylbenzene cation [Co(C5Me5)-(C11H16O2)]2+  相似文献   

16.
The oxidative coupling of methane (OCM) is an attractive route to convert natural gas directly into value-added chemical products (C2+). This work comparatively investigated SiO2- or La2O3-supported Na2WO4-MnxOy (denoted as NWM) catalysts in powder and fiber forms. The powder catalysts were prepared using a co-impregnation method and the fiber catalysts were prepared successfully using an electrospinning technique. The NWM/La2O3 fiber catalysts were activated at low temperature (500 °C) and had a 4.7% C2+ yield, with the maximum C2+ yield of 9.6% at 650 °C, while the NWM/SiO2 fiber catalyst was activated at 650 °C and had a maximum C2+ yield of 20.4% at 700 °C. The XPS results in the O 1s region indicated that NWM/La2O3 had a lower binding energy than NWM/SiO2, suggesting that the lattice oxygen species is easily released from the catalyst surface and creates vacancy sites that enhance performance. The stability test of the catalysts indicated that the La2O3-containing catalysts had excellent activity and high thermal stability, while the SiO2-containing catalysts had a higher C2+ yield when the prepared catalysts were compared at 700 °C. Considering the same component catalysts, the fiber catalysts achieved higher performance because their heat and mass transfer properties were enhanced.  相似文献   

17.
Efficient conversion of lignin to aromatic hydrocarbons via depolymerization and subsequent hydrodeoxygenation is important. Previously, we found that NbOx species played a key role in the activation and cleavage of C–O bonds in lignin and its model compounds. In this study, commercial niobic acid (HY-340), niobium phosphate (NbPO-CBMM) and lab-made layered niobium oxide (Nb2O5-Layer) were chosen as supports to study the effect of Brönsted and Lewis acids on the activation of C–O bonds in lignin conversion. A variety of Ru-loaded, Nb-based catalysts with different Ru particle sizes were prepared and applied to the conversion of p-cresol. The results show that all the Ru/Nb-based catalysts produce high mole yields of C7–C9 hydrocarbons (82.3–99.1%). What's more, Ru/Nb2O5-Layer affords the best mole yield of C7–C9 hydrocarbons and selectivity for C7–C9 aromatic hydrocarbons, of up to 99.1% and 88.0%, respectively. Moreover, it was found that Lewis acid sites play important roles in the depolymerization of enzymatic lignin into phenolic monomers and the cleavage of the C–O bond of phenols. Additionally, the electronic state and particle size of Ru are significant factors which influence the selectivity for aromatic hydrocarbons. A partial positive charge on the metallic Ru surface and a smaller Ru particle size are beneficial in improving the selectivity for aromatic hydrocarbons.  相似文献   

18.
The structure of the title compound, Na2[Zn(C6H11O2)4], consists of two‐dimensional polymeric sheets. The Zn2+ ions are approximately tetrahedrally coordinated by O atoms from different hexanoate anions. Both Na+ ions are six‐coordinated by carboxyl­ate O atoms. One of the hexanoate O atoms is attached to one Zn2+ ion and one Na+ ion, and the remaining O atom is attached to two Na+ ions.  相似文献   

19.
Catalytic oxidative coupling of methane over perovskite CaTiO3 prepared by modified ceramic method has been studied. The C2 yield was 13% at 830 °C and promotion with Na4P2O7 did not improve considerably the catalyst peformance. Regarding reactor test and mechanism studies it is believed that charge deficient, oxygen O generated by transforming oxygen adsorbed on surface defects and dissolved into bulk vacancies is responsible for activating methane (active site). Adsorbed oxygen generates the oxidizing sites for actived methane. Strict conditions are required for generation and regeneration of the activating sites in lattice.  相似文献   

20.
《中国化学快报》2023,34(1):107298
Photocatalytic selective transform native lignin into valuable chemicals is an attractive but challenging task. Herein, we report a mesoporous sulfur-doped carbon nitride (MSCN-0.5) which is prepared by a facile one-step thermal condensation strategy. It is highly active and selective for the cleavage Cα?Cβ bond in β?O?4 lignin model compound under visible light radiation at room temperature, achieving 99% substrate conversion and 98% Cα?Cβ bond cleavage selectivity. Mechanistic studies revealed that the Cβ?H bond of lignin model compounds activated by holes and generate key Cβ radical intermediates, further induced the Cα?Cβ bond cleavage by superoxide anion radicals (?O2?) to produce aromatic oxygenates. Waste Camellia oleifera shell (WCOS) was taken as a representative to further understand the reaction mechanisms on native lignin. 33.2 mg of monophenolic compounds (Vanillin accounted for 22% and Syringaldehyde for 34%) can be obtained by each gram of WCOS lignin, which is 2.5 times as that of the pristine carbon nitride. The present work offers useful guidance for designing metal-free heterogeneous photocatalysts for Cα?Cβ bond cleavage to harvest monophenolic compounds.  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