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.
Optimization of evaporation and pyrolysis conditions for ethanol are important in carbon nanotube (CNT) synthesis. The activation enthalpy (ΔH(?)), the activation entropy (ΔS(?)), and the free energy barrier (ΔG(?)) to evaporation have been determined by measuring the molar coefficient of evaporation, k(evap), at nine different temperatures (30-70 °C) and four gas flow rates (25-200 mL/min) using nitrogen and argon as carrier gases. At 70 °C in argon, the effect of the gas flow rate on k(evap) and ΔG(?) is small. However, this is not true at temperatures as low as 30 °C, where the increase of the gas flow rate from 25 to 200 mL/min results in a nearly 6 times increase of k(evap) and decrease of ΔG(?) by ~5 kJ/mol. Therefore, at 30 °C, the effect of the gas flow rate on the ethanol evaporation rate is attributed to interactions of ethanol with argon molecules. This is supported by simultaneous infrared spectroscopic analysis of the evolved vapors, which demonstrates the presence of different amounts of linear and cyclic hydrogen bonded ethanol aggregates. While the amount of these aggregates at 30 °C depends upon the gas flow rate, no such dependence was observed during evaporation at 70 °C. When the evaporation was carried out in nitrogen, ΔG(?) was almost independent of the evaporation temperature (30-70 °C) and the gas flow rate (25-200 mL/min). Thus the evaporation of ethanol in a dynamic gas atmosphere at different temperatures may go via different mechanisms depending on the nature of the carrier gas.  相似文献   

2.
Thermodynamic and kinetics studies for adsorption of Pt ions complexes from the chloride solutions obtained by leaching chlorinated spent automotive catalysts on anionic exchange resin Diaion WA21J were carried out. It was found that only Si, Pt, Rh and Pd from the solution were selectively adsorbed on the resin Diaion WA21J more strongly. The adsorption equilibrium time for Pt ions was about 20 h. The isothermal adsorption of Pt ions was found to fit Langmuir, Freundlich and DKR models. The maximum monolayer adsorption capacities Q(max) and X(m) of Pt ions on the resin based on Langmuir and DKR model were 4.85, 5.36 and 5.69 mg/g as well as 5.01, 5.63 and 5.98 mg/g for temperatures 18°C, 28°C and 40°C, respectively. The apparent adsorption energy E(ad) based on DKR model were -11.79, -11.04 and -11.04 kJ/mol for the temperatures 18°C, 28°C and 40°C, respectively. Ion exchange was the mechanism involved in the adsorption process. The adsorption of Pt ions on the resin underwent pseudo-first-order kinetic process, and the apparent adsorption activation energy E(a,1) was 12.6 kJ/mol. The intraparticle diffusion of Pt ions was a main rate-controlling step in most of time of adsorption process.  相似文献   

3.
The rate of cyclohexane photo-catalytic oxidation to cyclohexanone over anatase TiO(2) was studied at temperatures between 23 and 60 °C by in situ ATR-FTIR spectroscopy, and the kinetic parameters were estimated using a microkinetic model. At low temperatures, surface cyclohexanone formation is limited by cyclohexane adsorption due to unfavorable desorption of H(2)O, rather than previously proposed slow desorption of the product cyclohexanone. Up to 50 °C, the activation energy for photocatalytic cyclohexanone formation is zero, while carboxylates are formed with an activation energy of 18.4 ± 3.3 kJ mol(-1). Above 50 °C, significant (thermal) oxidation of cyclohexanone contributes to carboxylate formation. The irreversibly adsorbed carboxylates lead to deactivation of the catalyst, and are most likely the predominant cause of the non-Arrhenius behavior at relatively high reaction temperatures, rather than cyclohexane adsorption limitations. The results imply that elevating the reaction temperature of photocatalytic cyclohexane oxidation reduces selectivity, and is not a means to suppress catalyst deactivation.  相似文献   

4.
We report a facile synthesis of ultrathin (2.5 nm) trimetallic FePtPd alloy nanowires (NWs) with tunable compositions and controlled length (<100 nm). The NWs were made by thermal decomposition of Fe(CO)(5) and sequential reduction of Pt(acac)(2) (acac = acetylacetonate) and Pd(acac)(2) at temperatures from 160 to 240 °C. These FePtPd NWs showed composition-dependent catalytic activity and stability for methanol oxidation reaction. Among FePtPd and FePt NWs as well as Pd, Pt, and PtPd nanoparticles (NPs) studied in 0.2 M methanol and 0.1 M HClO(4) solution, the Fe(28)Pt(38)Pd(34) NWs showed the highest activity, with their mass current density reaching 488.7 mA/mg Pt and peak potential for methanol oxidation decreasing to 0.614 V from 0.665 V (Pt NP catalyst). The NW catalysts were also more stable than the NP catalysts, with the Fe(28)Pt(38)Pd(34) NWs retaining the highest mass current density (98.1 mA/mg Pt) after a 2 h current-time test at 0.4 V. These trimetallic NWs are a promising new class of catalyst for methanol oxidation reaction and for direct methanol fuel cell applications.  相似文献   

5.
We synthesized superconducting fullerene nanowhiskers (C(60)NWs) by potassium (K) intercalation. They showed large superconducting volume fractions, as high as 80%. The superconducting transition temperature at 17 K was independent of the K content (x) in the range between 1.6 and 6.0 in K-doped C(60) nanowhiskers (K(x)C(60)NWs), while the superconducting volume fractions changed with x. The highest shielding fraction of a full shielding volume was observed in the material of K(3.3)C(60)NW by heating at 200 °C. On the other hand, that of a K-doped fullerene (K-C(60)) crystal was less than 1%. We report the superconducting behaviors of our newly synthesized K(x)C(60)NWs in comparison to those of K(x)C(60) crystals, which show superconductivity at 19 K in K(3)C(60). The lattice structures are also discussed, based on the x-ray diffraction (XRD) analyses.  相似文献   

6.
A comprehensive theoretical investigation of thermal rearrangements of 2-vinylmethylenecyclopropane and 3-vinylcyclobutene is carried out employing density functional theory and high level ab initio methods, such as the complete active space self-consistent field, multi-reference second-order M?ller-Plesset perturbation theory, and coupled-cluster singles and doubles with perturbative triples. In all computations, Pople's polarized triple-ζ split valence basis set, 6-311G(d,p), is utilized. The potential energy surface for the relevant system is explored to provide theoretical insights for the thermal aromatizations of 2-vinylmethylenecyclopropane and 3-vinylcyclobutene. The rate constant for each isomerization reaction is computed using the transition state theory. The simultaneous first-order ordinary-differential equations are solved numerically for the considered system to obtain time-dependent concentrations, hence the product distributions at a given temperature. Our results demonstrate that at high temperatures thermal aromatizations of 2-vinylmethylenecyclopropane (at 700 °C and higher) and 3-vinylcyclobutene (at 500 °C and higher) are feasible under appropriate experimental conditions. However, at low temperatures (at 500 °C and lower), 2-vinylmethylenecyclopropane yields 3-methylenecyclopentene as a unique product, kinetically, and the formation of benzene is not favorable. Similarly, at 300 °C and lower temperatures, 3-vinylcyclobutene can only yield trans-1,3,5-hexatriene (major) and cis-1,3,5-hexatriene (minor). At 300 < T < 500 °C, 3-vinylcyclobutene almost completely yields 1,3-cyclohexadiene. Hence, our computations provide a useful insight for the synthesis of substituted aromatic compounds. Further, calculated energy values (reaction energies and activation parameters) are in satisfactory agreement with the available experimental results.  相似文献   

7.
We investigated the effects of trehalose on thermal inactivation and aggregation of human brain-type creatine kinase (hBBCK) in this study. In the presence of 1.0 M trehalose, the midpoint temperature of thermal inactivation (T (m)) of hBBCK increased by 4.6 °C, and the activation energy (E (a)) for thermal inactivation increased from 29.7 to 41.1 kJ mol(-1). Intrinsic fluorescence spectra also showed an increase in the apparent transition temperature (T (1/2)) of hBBCK from 43.0 °C to 46.5 °C, 47.7 °C, and 49.9 °C in 0, 0.6, 0.8, and 1.2 M trehalose, respectively. In addition, trehalose significantly blocked the aggregation of hBBCK during thermal denaturation. Our results indicate that trehalose has potential applications as a thermal stabilizer and may aid in the folding of other enzymes in addition to hBBCK.  相似文献   

8.
传统的Shilov反应是以PtCl2作为催化剂在水溶液中实现甲烷转化的,该反应的条件温和,在低至80°C时即可将甲烷中非常稳定的C–H键活化.然而,如果将反应温度提高达100°C以上,催化剂Pt(II)则非常容易发生歧化反应转化为Pt(0)或者Pt(IV),其中Pt(0)将会以沉淀的形式存在于反应溶液中.所以该反应只能在较低的温度进行, Shilov体系也只能得到较低的甲烷转化率,因此如何避免高温时催化剂因沉淀失活成为了提高反应转化率的研究重点.本文重点考察了高温条件下Shilov体系的反应机理和反应动力学,从而寻求提高催化体系活性和稳定性的途径.我们在特殊设计的金管反应器中进行了一系列的H/D置换实验,通过GC根据产物不同的分子量来分析检测.实验中,利用特殊设计的金管反应器可将反应压力增加到25.5 MPa,此时甲烷的溶解度与常温条件下(~60°C)相比可被提高1000倍以上,因此甲烷的转化率大大提高.在高温(~200°C)条件下的Shilov体系的水溶液中添加了CD3COOD, F3COOD, D2SO4, DCl和一系列阳离子为[1mim]+的离子液体来考察它们对催化剂沉淀的抑制作用,结果发现,在140°C时添加30%CD3COOD可在少量催化剂存在的条件下就能够明显促进H/D交换,与Shilov的结论吻合.这可能是由于CD3COO基团的螯合作用造成的,但将反应温度升到150°C时则不可避免的生成了Pt(0)沉淀.而F3COOD却在较多催化剂的条件下仍未表现出明显作用,可能是因为F较强的亲电子性使得F3COO基团的螯合作用变弱所致.在140°C时, D2SO4和DCl均能有效抑制Pt(0)沉淀的生成,尤其是DCl,在185°C反应24 h后仍能够稳定水溶液中的Pt基催化剂,但是在该条件下D2SO4却并没有作用.我们还发现, Cl–的浓度与沉淀的抑制直接相关,浓度越高对Pt基催化剂的稳定作用越强,但质子浓度的增加则对沉淀现象没有太大影响,我们推断原因是大量的Cl-能够在[PtCl6]2–的共同作用下将Pt(0)重新转化为了[PtCl6]2–.在140°C进行反应时,各类离子液体的添加能够使Pt(0)沉淀得到抑制,但是对H/D交换率却没有影响,可能是因为离子液体与Pt基催化剂螯合形成了Pt-离子液复合物而削弱了催化活性.在此基础上,我们特别考察了Cl–浓度对催化剂沉淀的影响,发现在200°C时将Cl-浓度提高到一定程度,就能够完全抑制Pt(0)的生成,但Pt基催化剂的活性也会被同时削弱.由于高压金管反应器的应用和高浓度Cl–的添加,使得甲烷的转化率达到90%以上,因此,我们设计了H/D同位素交换实验来考察反应的活性和选择性,从而针对高温Shilov体系的反应动力学进行研究.反应在200°C时进行,催化剂为K2PtCl4,反应介质为30% CD3COOD和DCl的水溶液,实验产物中检测到了CH3D, CH2D2, CHD3和CD4四种甲烷的多重氘代同位素体,说明了交换反应中有多个C–H键被活化.在此基础上,为了对甲烷活化过程进行全面描述,我们建立了涵盖所有连锁反应在内的综合反应网络,其中包含了H/D交换过程中涉及到的一系列平行的一级反应,基于实验数据通过阿伦尼乌斯方程计算得到了全部反应的频率因子、活化能和化学计量系数等反应动力学参数.结果证明,由于甲烷中所有的C–H键均相同,因此多重氘代产物的生成在甲烷转化过程中是不可避免的.其中,甲烷的单一氘代反应活化能为29.9 kcal/mol,双重氘代反应活化能为29.8 kcal/mol,两者十分相近,因此甲烷活化后的单一氘代产物的选择性最高不会超过50%.  相似文献   

9.
Anhydrous proton transport at temperatures above 100 °C has attracted considerable attention in the development of fuel cells that operate at intermediate temperatures. Liquid-state imidazole (ImH) is known to be a fast anhydrous proton conductor above 100 °C; however, evaporation and severe conductivity drops above and below its melting point (~90 °C), respectively, are major drawbacks to ImH. In this paper, we report a novel solid-state anhydrous ImH-Al(H(2)PO(4))(3) (AlP) hybrid material prepared via a simple synthesis using mechanical milling. This solid-state hybrid exhibits relatively a high ionic conductivity of ~0.1 mS cm(-1) at 100 °C and remarkably a small activation energy of 0.23 eV. In addition, the ImH-AlP hybrid material provides a means of overcoming both temperature-dependent drawbacks to pure ImH: (1) the ImH-AlP hybrid is thermally stable up to 130 °C, and (2) the hybrid material maintains high ionic conductivity below the melting point of ImH.  相似文献   

10.
We investigated the dewetting of metastable poly(N-vinylpyrrolidone) (PNVP) thin films (45 nm) on top of polystyrene (PS) thin films (58 nm) as a function of annealing temperature and molecular weight of PS (96 and 6850 kg/mol). We focused on the competition between dewetting, occurring as a result of unfavorable intermolecular interactions at the PNVP/PS interface, and spontaneous cross-linking of PNVP, occurring during thermal annealing, as we recently reported (Telford, A. M.; James, M.; Meagher, L.; Neto, C. ACS Appl. Mater. Interfaces 2010, 2, 2399-2408). Using optical microscopy, we studied how the dewetting morphology and dynamics at different temperatures depended on the relative viscosity of the top PNVP film, which increased with cross-linking time, and of the bottom PS film. In the PNVP/PS96K system, cross-linking dominated over dewetting at temperatures below 180 °C, reducing drastically nucleated hole density and their maximum size, while above 180 °C the two processes reversed, with complete dewetting occurring at 200 °C. On the other hand, the PNVP/PS6850K system never achieved advanced dewetting stages as the dewetting was slower than cross-linking in the investigated temperature range. In both systems, dewetting of the PNVP films could be avoided altogether by thermally annealing the bilayers at temperatures where cross-linking dominated. The cross-linking was characterized quantitatively using neutron reflectometry, which indicated shrinkage and densification of the PNVP film, and qualitatively through selective removal of the bottom PS film. A simple model accounting for progressive cross-linking during the dewetting process predicted well the observed hole growth profiles and produced estimates of the PNVP cross-linking rate coefficients and of the activation energy of the process, in good agreement with literature values for similar systems.  相似文献   

11.
改善SBA-15介孔材料水热稳定性的简单溶剂热后处理方法   总被引:1,自引:0,他引:1  
提出了一种有效改善SBA-15介孔材料水热稳定性的简单溶剂热后处理方法. SBA-15材料经环己烷、甲苯和正丁醇等有机溶剂在157和190 oC密闭容器中分别处理6–24 h后,可呈现很好的水热稳定性.它们在800 oC经100%水蒸气处理12 h,依然能保持很好的有序介孔结构,比表面积可高达192–281 m2/g.其中,经环己烷190 oC溶剂热处理24 h的样品表现出最优的水热稳定性.溶剂热处理能显著提升材料孔壁中类似Si(OSi)2(OH)2和Si(OSi)3OH结构的Si–OH基间脱水,形成稳定的Si(OSi)4结构,从而有效减少了SBA-15材料孔壁的缺陷.由此,介孔材料的水热稳定性得到明显改善.溶剂热处理对SBA-15材料水热稳定性的这种提升作用与所用溶剂性质、处理温度以及SBA-15前驱体的类型密切相关.其中,以低沸点的非极性溶剂处理焙烧后的SBA-15材料表现出最好的稳定化效果.该方法具有简单、低能耗的特点,其在制备高水热稳定的有序硅基介孔材料上有很好的潜在应用价值.  相似文献   

12.
Silicon is an extremely important technological material, but its current industrial production by the carbothermic reduction of SiO2 is energy intensive and generates CO2 emissions. Herein, we developed a more sustainable method to produce silicon nanowires (Si NWs) in bulk quantities through the direct electrochemical reduction of CaSiO3, an abundant and inexpensive Si source soluble in molten salts, at a low temperature of 650 °C by using low‐melting‐point ternary molten salts CaCl2–MgCl2–NaCl, which still retains high CaSiO3 solubility, and a supporting electrolyte of CaO, which facilitates the transport of O2− anions, drastically improves the reaction kinetics, and enables the electrolysis at low temperatures. The Si nanowire product can be used as high‐capacity Li‐ion battery anode materials with excellent cycling performance. This environmentally friendly strategy for the practical production of Si at lower temperatures can be applied to other molten salt systems and is also promising for waste glass and coal ash recycling.  相似文献   

13.
Hematite (α-Fe(2)O(3)) was grown on vertically aligned Si nanowires (NWs) using atomic layer deposition to form a dual-absorber system. Si NWs absorb photons that are transparent to hematite (600 nm < λ < 1100 nm) and convert the energy into additional photovoltage to assist photoelectrochemical (PEC) water splitting by hematite. Compared with hematite-only photoelectrodes, those with Si NWs exhibited a photocurrent turn-on potential as low as 0.6 V vs RHE. This result represents one of the lowest turn-on potentials observed for hematite-based PEC water splitting systems. It addresses a critical challenge of using hematite for PEC water splitting, namely, the fact that the band-edge positions are too positive for high-efficiency water splitting.  相似文献   

14.
We investigated the structure and electronic properties of unidirectional Pt(2)Si nanowires (NWs) grown on a Si(100)-2 degrees off surface. We found that Pt(2)Si NWs were formed along the step edges of the Si(100)-2 degrees off surface with c(4x6) reconstructions that occurred on the terraces of Si(100) using scanning tunneling microscopy and the structure of formed NWs was found to be Pt(2)Si by core-level photoemission spectroscopy. Moreover, we confirmed that the electronic band structures of the NWs along the NW direction are different from those perpendicular to the NWs and the surface state induced by the Pt(2)Si NWs was observed with a small density of state using the angle-resolved photoemission spectra.  相似文献   

15.
A semiempirical nearest‐neighbor tight‐binding approach, that reproduces the indirect band gaps of elemental semiconductors, has been applied to study the electronic and optical properties of Si and Ge nanowires (NWs). The calculations show that Si‐NWs keep the indirect bandgap whereas Ge‐NWs changes into the direct bandgap when the wire cross section becomes smaller. Also, the band gap enhancement of Si‐NWs showing to quantum confinement effects is generally larger than that of similar‐sized Ge‐NWs, confirming the larger quantum confinement effects in Si than in Ge when they are confined in two dimensions. Finally, the dependence of the imaginary part of the dielectric function on the quantum confinement within two different schemes: intra‐atomic and interatomic optical matrix elements are applied. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem 110:2448–2454, 2010  相似文献   

16.
ZnO nanowires (NWs) were synthesized on Au-coated Si (100) substrates by vapor transport method. The effect of high temperature annealing on the structural and chemical composition as well as thermal stability was studied. The as-prepared ZnO NWs was nearly stoichiometric and identified as hexagonal ZnO phase. After annealing at 1,473 K, the atomic ratio of O/Zn, the intensity of the diffraction peaks, and the diameter of nanowires were increased. The ZnO NWs were fragmented into nanocrystals and the fragments coalesced with each other after annealing at 1,673 K. The thermal stability of ZnO NWs was studied by thermo-gravimetric (TG) analysis. A sharp increase in the TG curves was observed and can be attributed to the oxidation of some possibly presented Zn atoms. The activation energy of oxidation of Zn interstitial atoms was found to be 484.81 kJ mol?1. A mass gain peak was observed after annealing at 1,473 K, but it was completely eliminated after annealing at 1,673 K.  相似文献   

17.
The normal phase HPLC behavior of a bare zirconia column was studied at temperatures up to 200 °C using a hexane mobile phase. The use of elevated column temperatures significantly decreased the retention of twenty five aromatic model compounds according to the van't Hoff equation (>30-fold decrease for some compounds). Large improvements in peak shape, efficiency (>2.2-fold), aromatic group-type selectivity, and column re-equilibration times (>5-fold) were obtained at elevated temperatures. The thermal decomposition of two polar nitrogen compounds (indole and carbazole) was observed in a hexane/dichloromethane mobile phase at temperatures greater than 100 °C. The first order decomposition of carbazole was studied in further detail.  相似文献   

18.
Ti(2)(Ti(0.16)Ni(0.43)Al(0.41))(3) is a novel compound (labeled as τ(6)) in the Ti-rich region of the Ti-Ni-Al system in a limited temperature range 870 < T < 980 °C. The structure of τ(6)-Ti(2)(Ti,Ni,Al)(3) was solved from a combined analysis of X-ray single crystal and neutron powder diffracton data (space group C2/m, a = 1.85383(7) nm, b = 0.49970(2) nm, c = 0.81511(3) nm, and β = 99.597(3)°). τ(6)-Ti(2)(Ti,Ni,Al)(3) as a variant of the V(2)(Co(0.57)Si(0.43))(3)-type is a combination of slabs of the MgZn(2)-Laves type and slabs of the Zr(4)Al(3)-type forming a tetrahedrally close-packed Frank-Kasper structure with pentagon-triangle main layers. Titanium atoms occupy the vanadium sites, but Ti/Ni/Al atoms randomly share the (Co/Si) sites of V(2)(Co(0.57)Si(0.43))(3). Although τ(6) shows a random replacement on 6 of the 11 atom sites, it has no significant homogeneity range (~1 at. %). The composition of τ(6) changes slightly with temperature. DSC/DTA runs (1 K/min) were not sufficient to define proper reaction temperatures due to slow reaction kinetics. Therefore, phase equilibria related to τ(6) were derived from X-ray powder diffraction in combination with EPMA on alloys, which were annealed at carefully set temperatures and quenched. τ(6) forms from a peritectoid reaction η-(Ti,Al)(2)Ni + τ(3) + α(2) ? τ(6) at 980 °C and decomposes in a eutectoid reaction τ(6) ? η + τ(4) + α(2) at 870 °C. Both reactions involve the η-(Ti,Al)(2)Ni phase, for which the atom distribution was derived from X-ray single crystal intensity data, revealing Ti/Al randomly sharing the 48f- and 16c-positions in space group Fd3?m (Ti(2)Ni-type, a = 1.12543(3) nm). There was no residual electron density at the octahedral centers of the crystal structure ruling out impurity stabilization. Phase equilibria involving the τ(6) phase have been established for various temperatures (T = 865, 900, 925, 950, 975 °C, and subsolidus). The reaction isotherms concerning the τ(6) phase have been established and are summarized in a Schultz-Scheil diagram.  相似文献   

19.
In this paper we report the successful incorporation of silicon into SrMO(3) (M = Co, Mn) leading to a structural change from a hexagonal to a cubic perovskite. For M = Co, the cubic phase was observed for low doping levels (3%), and these doped phases showed very high conductivities (up to ≈350 Scm(-1) at room temperature). However, annealing studies at intermediate temperatures (700-800 °C), indicated that the cubic phase was metastable with a gradual transformation to a hexagonal cell on annealing. Further work showed that co-doping with Fe resulted in improved stability of the cubic phase; a composition SrCo(0.85)Fe(0.1)Si(0.05)O(3-y) displayed good stability at intermediate temperatures and a high conductivity (≈150 Scm(-1) at room temperature). For M = Mn, the work showed that higher substitution levels were required to form the cubic perovskite (≈15% Si doping), although in these cases the phases were shown to be stable to annealing at intermediate temperatures. Conductivity measurements again showed an enhancement in the conductivity on Si doping, although the conductivities were lower (≈0.3-14 Scm(-1) in the range 20-800 °C) than the cobalt containing systems. The conductivities of both systems suggest potential for use as cathode materials in solid oxide fuel cells.  相似文献   

20.
Si nanowires (NWs) were fabricated in a vacuum furnace using a Nd:YAG pulsed laser with the wavelength of 325 nm. Commercial p-type Si wafer is used for the target, and no catalytic materials are used. Scanning electron microscopy (SEM) images indicate that the diameters of Si NWs ranged from 10 to 150 nm. Si NWs have various sizes and shapes with a substrate position inside the furnace, and their morphologic construction is reproducible. The formation mechanism of the NWs is discussed.  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