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.
NiO(8 %)/Ni,H-ZSM-5 + Al2O3 (1:1) catalysts differing in metal-support interactions, which influenced the metal-to-acid ratios, were examined. The interactions were changed by modifying the method of zeolite and aluminium hydroxide combining and the method of Ni incorporation. The catalysts were characterised by ICP, XRD, N2 sorption, SEM, TEM, NH3-TPD, Py-IR, TPR, H2 chemisorption and XPS. The effect of metal-support interactions was determined during n-C6 conversion in a continuous system at H2:CH = 7:1 Nm3/m3, 0.1 MPa and LHSV = 1 h?1. It was found that over the catalysts with weaker Ni–alumina interactions (n Ni_a/n, 3.2 × 10?2 and 4.8 × 10?2), selectivity to isomerisation products was by 10–35 % higher, and selectivity to high boiling hydrocarbons by 10–30 % lower than over the catalysts with stronger Ni-support interactions (n Ni_a/n, 1.2 × 10?2 and 1.8 × 10?2).  相似文献   

2.
Two one-dimensional Cd(II) coordination polymers constructed by ribbon-like molecules, {[Cd(L1)2(NO3)]·NO3} n (1) (L1 = 1,3-bis(5,6-dimethylbenzimidazol-1-yl)benzene) and [Cd(L2)2(NO3)2] n (2) (L2 = 1,3-bis(benzimidazol-1-yl)benzene) have been obtained through assembly of the two structurally similar flexible bis(benzimidazole)-based ligands and cadmium nitrate tetrahydrate. The cadmium(II) centers display different coordination environments with trigonal-bipyramidal geometry in 1, and a octahedral geometry in 2. Weak C–H···O interactions in 1 result in a two-dimensional supramolecular layer; two π–π interactions are present in 2 forming a three-dimensional supramolecular network structure. The solid state fluorescence properties of 1 and 2 were investigated.  相似文献   

3.
A novel electrochemical sensor for the determination of p-nitrophenol (PNP) was fabricated with the nanoscaled composite oxide Mg(Ni)FeO-modified carbon paste electrode (CPE), and its electrocatalytic performances were investigated using the cyclic voltammogram and differential pulse voltammetry techniques. The influential factors were optimized such as the mass ratio of Mg(Ni)FeO to graphite, the pH value of buffer solution and the accumulation time at open circuit. The indirect oxidation peak current of PNP was found to be proportional to its concentration between 2.0 × 10?6 and 2.0 × 10?4 M on the proposed sensor Mg(Ni)FeO/CPE under the optimal condition (10 % Mg(Ni)FeO/graphite, pH 5.0 HAc–NaAc, 120 s quiescence). The sensor Mg(Ni)FeO/CPE exhibited a high sensitivity of 811 μA mM?1 cm?2 and a low detection limit of 0.2 μΜ (S/N– = 3) for PNP detection, and got satisfactory results when it was applied to determine PNP in real samples. The results demonstrate that Mg(Ni)FeO/CPE based on the nanomaterial Mg(Ni)FeO with high specific area and mesoporous structure could be employed as an electrochemical sensor for PNP determination with simplicity, low cost, good selectivity, repeatability, and stability.  相似文献   

4.
BACKGROUND: Ethylene oligomerization is the major industrial process to produce linear α‐olefins. Recently much work has been devoted to late transition metal catalysts used in this process, especially those with 2,6‐bis(imino)pyridyl dihalide ligands. Considering that most work has focused on simple modification to the substituents in imino‐aryl rings based on the symmetric bis(imino)pyridyl framework, here we expand this work to the asymmetric mono(imino)pyridyl ligands. RESULTS: The preparation, structure and ethylene polymerization/oligomerization behavior of series of mono(imino) pyridyl–MCl2 and bis(imino)pyridyl–MXn complexes are presented. The systematic studies were focused on the relationship between the catalytic behavior of these complexes for ethylene polymerization/oligomerization and reaction conditions, ligand structures, metal centers and counter‐anions. The influence of the coordination environment on catalyst behavior is also discussed. CONCLUSION: For mono(imino)pyridyl–Co(II) and ? Ni(II) catalysts bearing the Cl? counter‐anion, good activities ranging from 0.513 × 105 to 1.58 × 105 g polyethylene (mol metal)?1 h?1 atm?1 are afforded, and the most active catalysts are those with methyl in both ortho‐ and para‐positions of the imine N‐aryl ring. For bis(imino)pyridyl–Co(II) and ? Ni(II) catalysts bearing the SO42? and NO3? counter‐anions, the low activities for ethylene oligomerization are in sharp contrast to those of their chloride analogues. Copyright © 2009 Society of Chemical Industry  相似文献   

5.
Nitrous oxide is produced in soils and sediments essentially through the processes of nitrification and denitrification, although several rival processes could be competing. This study was undertaken in order to better understand the controlling factors of nitrification, denitrification and associated N2O production as well as the contribution of these two processes to the average N2O production by soils and sediments. With this aim, soil and sediment samples were taken in contrasting periods and different land use types, each time at different depths (upper and lower soil horizons). They were incubated in separate batches in specific conditions to promote denitrification and nitrification: (1) a complete anaerobic environment adding KNO3 for the denitrification assay and (2) an aerobic environment (21 % O2) with addition of NH4Cl for the nitrification assay. Potentials of nitrification and denitrification were determined by the rates of nitrate either reduced (for denitrification) or produced (for nitrification). Overall, denitrification potential varied from 70 to 2,540 ng NO 3 -N g?1 dry soil h?1 and nitrification potential from 30 to 1,150 ng NO3 ?-N g?1 dry soil h?1. Nitrous oxide production by denitrification was significantly (P < 0.05) greater in topsoils (10–30 cm) than in subsoils (90–110 cm), ranging, respectively, from 26 to 250 ng N2O-N g?1 dry soil h?1 versus 1.5 to 31 ng N2O-N g?1 dry soil h?1, i.e., a mean 19.5 versus. 6.0 % of the NO3 ? denitrified for the upper and lower horizons, respectively. Considering the N2O production in relation with the nitrate production (e.g., nitrification), no significant difference (P < 0.05) was found in the soil profile, which ranged from 0.03 to 6 ng N2O-N g?1 dry soil h?1. This production accounts for 0.21 and 0.16 % of the mean of the NO3 ? produced for the top and subsoils, respectively. On the basis of the average production by both top- and subsoils, N2O production by denitrification is clearly greater than by nitrification under the measurement conditions used in this study, from 30- to 100-fold higher. Such a high potential of N2O emission must be taken into account when reducing nitrate contamination by increasing denitrification is planned as a curative measure, e.g. in rehabilitation/construction of wetlands.  相似文献   

6.
In the Seine Basin, characterised by intensive arable crops, most of the surface and groundwater is contaminated by nitrate (NO3 ?). The goal of this study is to investigate nitrogen leaching on commercial arable crop farms in five organic and three conventional systems. In 2012–2013, a total of 37 fields are studied on eight arable crop rotations, for three different soil and climate conditions. Our results show a gradient of soil solution concentrations in function of crops, lower for alfalfa (mean 2.8 mg NO3-N l?1) and higher for crops fertilised after legumes (15 mg NO3-N l?1). Catch crops decrease nitrate soil solution concentrations, below 10 mg NO3-N l?1. For a full rotation, the estimated mean concentrations is lower for organic farming, 12 ± 5 mg NO3-N l?1 than for conventional farming 24 ± 11 mg NO3-N l?1, with however a large range of variability. Overall, organic farming shows lower leaching rates (14–50 kg NO3-N ha?1) than conventional farms (32–77 kg NO3-N ha?1). Taking into account the slightly lower productivity of organic systems, we show that yield-scaled leaching values are also lower for organic (0.2 ± 0.1 kg N kg?1 N year?1) than for conventional systems (0.3 ± 0.1 kg N kg?1 N year?1). Overall, we show that organic farming systems have lower impact than conventional farming on N leaching, although there is still room for progress in both systems in commercial farms.  相似文献   

7.
A polyaniline/titanium nitride (PANI/TiN) nanotube hybrid was prepared and used for an electrochemical supercapacitor application. Firstly, the well-aligned TiN nanotube array was prepared by anodization of titanium foil and subsequent nitridation through ammonia annealing. Then, PANI was deposited into TiN nanotube through the electrochemical polymerization process. The obtained PANI/TiN nanotube hybrid had an ordered porous structure. A high specific capacitance of 1,066 F g?1 was obtained at the charge–discharge current density of 1 A g?1 when only the mass of PANI was considered. The specific capacitance can even achieve 864 F g?1 at 10 A g?1 and still keep 93 % of the initial capacity after 200 cycles. An aqueous supercapacitor, consisting of two symmetric PANI/TiN nanotube hybrid electrodes and 1.0 M H2SO4 electrolyte solution, showed the specific capacitance of 194.8 F g?1, energy density of 9.74 Wh kg?1, and power density of 0.3 kW kg?1.  相似文献   

8.
The self assembly of [Ni(L)]Cl2·2H2O (L = 3,14-dimethyl-2,6,13,17-tetraazatricyclo[14,4,01.18,07.12]docosane) and 1,2,3,4-cyclobutanetetracarboxylic acid (H4cbtc) acid generates a 1D coordination and 2D hydrogen-bonded polymer [Ni(L)(H2cbtc)2·3H2O] n (1). Complex 1 is characterized by X-ray crystallography, spectroscopy and magnetic susceptibility. Each nickel(II) ion has a distorted octahedral coordination environment with the four secondary amines of the macrocycle in which two carboxylate anions of the H2cbtc2? ligand have assembled around each nickel center. The compound crystallizes in the triclinic system P-1 with a = 9.715(3) Å, b = 12.891(5) Å, c = 13.903(6) Å, α = 72.64(2)°, β = 75.70(3)°, γ = 73.27(3)°, V = 1566.6(10) Å3, Z = 2. The electronic spectrum of 1 indicates a high-spin octahedral environment. The magnetic behavior of 1 reveals a weak intramolecular antiferromagnetic interaction with J = ?1.23(1) cm?1.  相似文献   

9.
Nickel–copper metallic foams were electrodeposited from an acidic electrolyte, using hydrogen bubble evolution as a dynamic template. Their morphology and chemical composition was studied by scanning electron microscopy and related to the deposition parameters (applied current density and deposition time). For high currents densities (above 1 A cm?2) the nickel–copper deposits have a three-dimensional foam-like morphology with randomly distributed nearly-circular pores whose walls present an open dendritic structure. The nickel–copper foams are crystalline and composed of pure nickel and a copper-rich phase containing nickel in solid solution. The electrochemical behaviour of the material was studied by cyclic voltammetry and chronopotentiometry (charge–discharge curves) aiming at its application as a positive electrode for supercapacitors. Cyclic voltammograms showed that the Ni–Cu foams have a pseudocapacitive behaviour. The specific capacitance was calculated from charge–discharge data and the best value (105 F g?1 at 1 mA cm?2) was obtained for nickel–copper foams deposited at 1.8 A cm?2 for 180 s. Cycling stability of these foams was also assessed and they present a 90 % capacitance retention after 10,000 cycles at 10 mA cm?2.  相似文献   

10.
M. Zielinska 《Drying Technology》2016,34(10):1147-1161
The objective of this study was to dry–wet distillers grains and centrifuged solubles and to examine the effect of two different drying media, superheated steam and hot air, at different drying temperatures (110, 130, and 160°C), moisture contents (5–30% wb), and percentages of solubles’ presence (0 or 100%) on some thermophysical properties of laboratory-prepared corn/wheat dried distillers co-products, including geometric mean diameter (dg), particle density (ρp), bulk density (ρb), bulk porosity (?b), specific heat (C), effective thermal diffusivity (αeff), and bulk thermal conductivity (λb). The values of dg of corn/wheat dried distillers co-products ranged from 0.358 ± 0.001 to 0.449 ± 0.001 mm. Experimental values of ρp, ρb, and ?b varied from 1171 ± 6 to 1269 ± 3 kg m?3, from 359 ± 7 to 605 ± 5 kg m?3, and from 0.54 ± 0.01 to 0.71 ± 0.01 kg m?3, respectively. The values of αeff were between 0.58 × 10?7 and 0.93 × 10?7 m2 s?1. The calculated values of C ranged from 1887 ± 11 to 2599 ± 19 J kg?1 K?1, and the values of λb of corn/wheat dried distillers co-products ranged from 0.06 ± 0.01 to 0.09 ± 0.01 W m?1 K?1. Multiple linear regression prediction models were developed to predict the changes in dg, ρp, ρb, ?b, C, αeff, and λb of laboratory-prepared corn/wheat dried distillers co-products with different operational factors.  相似文献   

11.
The reaction of silver nitrate with (1e)-N′-hydroxy-2-(1H-1,2,4-triazol-1-yl) ethanimidamide (HTE) at room temperature affords a new 3D metal-organic frameworks, {[Ag3(HTE)2(NO3)2]·NO3}n (1). Single crystal X-ray diffraction analysis reveals that the three silver metal centers are bridged by two HTE ligands and Ag-Ag strong interactions in 1. The bridges provided through the nitrogen atoms on triazole-ring and the nitrogen atoms of the ethanimidamide resulted in a trinuclear units {Ag3(HTE)2} linked metal organic framework (MOF) structure. The MOF is displaying the trinuclear units that are connected via nitrates to give a 3D network. There are free nitrates incorporated in the channels of the MOF.  相似文献   

12.
Mesoporous nickel oxide (NiO) nanoparticles were synthesized by the thermal decomposition reaction of Ni(NO3)2·9H2O using oxalic acid dihydrate as the mesoporous template reagent. The pore structure of nanocrystals could be controlled by the precursor to oxalic acid dihydrate molar ratio, thermal decomposition temperature and thermal decomposition time. The structural characteristic and textural properties of resultant nickel oxide nanocrytals were characterized by X-ray diffraction (XRD), transmission electron microscopy (TEM), scanning electron microscopy (SEM), N2 adsorption–desorption isotherm and temperature programmed reduction. The results showed that the most excellently mesoporous nickel oxide particles (m-Ni-1-4) with developed wormlike pores were prepared under the conditions of the mixed equimolar precursor and oxalic acid and calcined for 4 h at 400 °C. The specific surface area and pore volume of m-Ni-1-4 are 236 m2 g?1 and 0.42 cm3 g?1, respectively. Over m-Ni-1-4 at space velocity = 20,000 mL g?1 h?1, the conversions of toluene and formaldehyde achieved 90 % at 242 and 160 °C, respectively. It is concluded that the reactant thermal decomposition with oxalic acid assist is a key step to improve the mesoporous quality of the nickel oxide materials, the developed mesoporous architecture, high surface area, low temperature reducibility and coexistence of multiple oxidation state nickel species for the excellent catalytic performance of m-Ni-1-4.  相似文献   

13.
Two type lanthanide complexes obtained by salen-type ligand N,N’-ethylenebis(salicylideneimine) (H2La) and N,N’-ethylene-bis(3-methoxysalicylideneimine) (H2Lb) have been isolated. They are 2D coordination polymer [Yb(H2La)1.5(NO3)3] n (1), mononuclear Ln(H2Lb)(NO3)3 [Ln = Sm (2), Ho (3)]. The structures of 13 were determined by single crystal X-ray crystallographic studies and their photophysical properties were investigated.  相似文献   

14.
The antecedent effect of lime on the gaseous products of denitrification (N2O and N2) was examined in a laboratory study using a clay loam soil (Soil 1) with a starting pH of 5.4, and a sandy loam soil (Soil 2) with a starting pH of 5.3. The soils were amended with 0, 2.3, 5.7 and 18.9 g CaCO3 kg?1, and were incubated for a period of 3 years at 4 °C during which time the soil equilibrated at pH values of 4.7, 5.8, 7.3 and 7.7 (Soil 1) and pH 4.7, 5.2, 6.6 and 7.6 (Soil 2). Ammonium nitrate, labelled with 15N (15NH4NO3, NH4 15NO3 and 15NH4 15NO3) was added to each incubation jar at a rate of 7.14 μmol N g?1 oven dried (OD) soil. Headspace gas samples were extracted daily over a 5 days incubation period at 20 °C. The amount of N2O and N2, and 15N enrichment of N2O-N in the headspace, was determined using continuous-flow isotope-ratio mass spectrometry. As pH increased, the quantity of N2 and N2O emitted significantly increased in both soils (P < 0.001), with a peak N2 flux of 0.179 μmol N g?1 OD soil h?1, and a peak N2O flux of 0.002 μmol N g?1 OD soil h?1 occurring at pH 7.6, 2 days after the addition of NH4NO3. The loss as N2 far exceeded the loss of N2O, which remained at less than 1 % of the total mineral N content of the soil. Lime generally lowered the N2O:N2 ratio, however the results from this study suggest that it is not a mitigation strategy for GHG emissions.  相似文献   

15.
Proton-conducting polymer blend electrolytes based on PVA–PVP–NH4NO3 were prepared for different compositions by solution cast technique. The prepared films are investigated by different techniques. The XRD study reveals the amorphous nature of the polymer electrolyte. The FTIR and laser Raman studies confirm the complex formation between the polymer and salt. DSC measurements show decrease in T g with increasing salt concentration. The ionic conductivity of the prepared polymer electrolyte was found by ac impedance spectroscopy analysis. The maximum ionic conductivity was found to be 1.41 × 10?3 S cm?1 at ambient temperature for the composition of 50PVA:50PVP:30 wt% NH4NO3 with low-activation energy 0.29 eV. The conductivity temperature plots are found to follow an Arrhenius nature. The dielectric behavior was analyzed using dielectric permittivity (ε*) and the relaxation frequency (τ) was calculated from the loss tangent spectra (tan δ). Using this maximum ionic conducting polymer blend electrolyte, the primary proton battery with configuration Zn + ZnSO4·7H2O/50PVA:50PVP:30 wt% NH4NO3/PbO2 + V2O5 was fabricated and their discharge characteristics studied.  相似文献   

16.
Pd nanoparticles have been synthesised using different reducing agents, including ethylene glycol (EG), formaldehyde and sodium borohydride and their activity for the oxygen reduction reaction (ORR) evaluated. The use of EG led to the best morphology for the ORR and this synthetic method was optimised by adjusting the system pH. Carbon-supported Pd nanoparticles of approximately 7 nm diameter were obtained when reduction took place in the alkaline region. Pd synthesised by EG reduction at pH 11 presented the highest mass activity 20 A g?2 and active surface area 15 m2 g?1. These synthetic conditions were used in further synthesis. The effect of heat treatment in H2 atmosphere was also studied; and increased size of the palladium nanoparticles was observed in every case. The Pd/C catalyst synthesised by reduction with EG at pH 11 was tested in a low temperature H2/O2 (air) PEMFC with a Nafion® 112 membrane, at 20 and 40 °C. Current densities at 0.5 V, with O2 fed to the cathode, at 40 °C were 1.40 A cm?2 and peak power densities 0.79 W cm?2, approximately; which compared with 1.74 A cm?2 and 0.91 W cm?2, respectively for a commercial Pt/C.  相似文献   

17.
Sodium-ion batteries (SIBs) is considered as a promising alternative to lithium-ion batteries. Supercapacitors (SCs) are receiving great attention for their significantly higher power density than batteries and prolonged cycle life. Herein, SIBs and SCs based on N-doped amorphous multi-size pores dominated polymeric frameworks were fabricated and examined. The enlarged interlayer spacing and multi-size-pore dominated interconnected architecture with high specific surface area, high pore volume and high N content optimize the electrochemical performance of N-PPF-20. As an anode material, N-PPF-20 exhibited a sodium ion storage capacity of 432.2 mAh g?1 at a current density of 0.05 A g?1, while maintaining a reversible capacity of 61.1 mAh g?1 at an ultrahigh current density of 20 A g?1. Additionally, a specific capacity of 158.3 mAh g?1 at 1 A g?1 was obtained after 1000 cycles, indicating an excellent cycling stability. When tested as an electrode material for SCs, N-PPF-20 delivered a high specific capacitance of 438.7 F g?1 at 0.1 A g?1, and a specific capacitance of 111.2 F g?1 was achieved even at a high current density of 10 A g?1. Meanwhile, a long-term cycling life test demonstrated a specific capacitance of 120 F g?1 at an ultrahigh current density of 10 A g?1 after 10,000 cycles.  相似文献   

18.
The hydrothermal reaction of AgNO3 and 4-amino-3,5-bis(3-pyridyl)-1,2,4-triazole (NH2-BPT) generated an unprecedented three-dimensional coordination polymer {[Ag2(BPT)] · NO3}n (1). The crystal structure reveals that two independent Ag(I) centers are linked via cisoid μ5-BPT anions to form a unique tetranuclear cluster, which is further interconnected to four neighbors to give a 3D cationic framework with dia-topology.  相似文献   

19.
The influence of ageing temperature, silver loading and type of reducing agent on the lean NO x reduction over silver–alumina catalysts was investigated with n-octane and bio-diesel (NExBTL) as reducing agent. The catalysts (2 and 6 wt% Ag–Al2O3) were prepared with a sol–gel method including freeze drying and the evaluation of NO x reduction and aging were performed using a synthetic gas-flow reactor. The results indicate a relatively high NO x reduction for both reducing agents. The hydrothermally treated 6 wt% Ag–Al2O3 sample displays a maximum NO x reduction of 78 % at 350 °C for n-octane as reductant and the corresponding value for NExBTL is 60 %. Furthermore, the catalysts show high durability and an increase in activity for NO x reduction after ageing at temperatures up to 650 °C, with n-octane as reducing agent.  相似文献   

20.
The biodegradable cholesteryl-(l-lactic acid) n (CPLA) was synthesized via ring opening polymerization of l-lactide in the presence of cholesterol as an initiator and the catalytic amount of Sn(Oct)2. The resulting monohydroxyl-terminated CPLA was subsequently converted to a bromine-ended macroinitiator (CPLA-Br) by esterification with 2-bromoisobutyryl bromide. Amphiphilic block-brush copolymers with different lengths of hydrophilic block (CPLA-b-P(PEGMA)4 and CPLA-b-P(PEGMA)12) were synthesized in a subsequent atom transfer radical polymerization of the poly(ethylene glycol)monomethyl ether methacrylate (PEGMA). The prepared polymers were characterized by FTIR, 1H NMR and GPC. The self-assembly of the copolymers into the micellar aggregates in aqueous media was followed with dynamic light scattering, transmission electron microscopy and fluorescence analysis. The CMC values of the CPLA-b-P(PEGMA)4 and CPLA-b-P(PEGMA)12 samples were estimated approximately 56 × 10?4 and 72 × 10?4 g/L in an aqueous solution by fluorescence probe technique, respectively. The hydrophobic/hydrophilic chain ratio of the amphiphilic copolymers could have demonstrated a correlation with micelle formation ability and inter-micellar aggregation in an aqueous solution. Using the naproxen as a hydrophobic model drug, the drug-loading efficiency and drug release properties of the CPLA–PEG nanoparticles were investigated. In vitro release study of the naproxen-loaded micelles with about 54–60 % loading efficiency and 11–12 % loading capacity was performed using dialysis method in phosphate-buffered solution at 37 °C. Accordingly, these polymeric micelles may be provided as an effective drug carrier for drug controlled release by modulating the copolymer composition and molecular weight of blocks.  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