首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 156 毫秒
1.
以Co(COOH)_2、FeCl_3和PdCl_2为原料,柠檬酸为稳定剂,乙醇为加速剂,采用超声辅助制备Co-Fe-Pd金属纳米粒子,并评估其氧还原反应(ORR)电催化性能。研究结果表明,Co-Fe-Pd金属纳米粒子平均粒径约3~5 nm,由于Co、Fe固溶于Pd晶格,使Co-Pd、Fe-Pd和Co-Fe-Pd纳米粒子仅显示Pd衍射峰,且伴有不同程度的宽化;相比于Co-Fe、Fe-Pd或Co-Pd纳米粒子,三元Co-Fe-Pd晶格压缩更为明显,晶格缺陷诱使的活性位点增加,氧还原催化能力增强;其氧还原起峰电位为1.03 V (vs RHE),Tafel斜率为-87 mV/dec,可与商用Pt/C催化剂相媲美;氧还原过程中电子转移数为3.80±0.04,说明其主导四电子转移路径;此外,RRDE结果显示氧还原过程中的中间产物H_2O_2含量约10%。  相似文献   

2.
马翔宇  金长春  董如林 《化工进展》2015,34(4):1019-1022,1073
以氧化石墨(GO)和Pd(NO3)2为原料,通过化学还原法制备Pd纳米粒子-石墨烯(Pd/G)纳米复合材料,然后以H2PtCl6作为Pt前体,在Pd纳米粒子的表面恒电位沉积Pt,制备不同Pt负载量的Pd/G(Pt-Pd/G)电极.利用场发射扫描电镜(FE-SEM)、透射电镜(TEM)和X射线能谱仪(EDX)对材料的微观结构进行了表征和分析.结果显示石墨烯上的金属粒子分散均匀,平均粒径约7.2nm.电化学测试结果显示Pt-Pd/G电极对乙二醇电化学氧化反应具有良好的催化性能.当纳米粒子的Pt:Pd原子百分比为1:42时,其反应峰电流密度分别为Pd/G和Pt/G电极的3.0倍和2.7倍.少量的Pt沉淀可显著改进Pd/G电极的催化活性.本研究采用的修饰方法简单,修饰效果明显,可应用于其他金属纳米复合材料的异金属修饰.  相似文献   

3.
李静  王乐  孙维周  左志军  黄伟 《化工进展》2015,34(3):738-744,757
以钴、钯为活性金属, 分别采用浸渍法和溶胶-凝胶法制备了Co-Pd/TiO2催化剂, 考察了不同制备方法制备的Co-Pd/TiO2催化剂对CH4-CO2梯阶转化直接合成C2含氧化合物的影响。利用XRD、XPS和N2-吸附-脱附对催化剂进行了表征。结果表明:两种方法制备的催化剂反应前与反应后表面织构都存在较大变化, 且催化剂中均存在CoTiO3物种, 这是活性金属Co与载体TiO2之间发生强相互作用, CO2+替代TiO2晶格中的Ti4+的结果;CoO和金属Pd可能是该反应的活性中心;反应前与反应后溶胶-凝胶法制备的催化剂的表面Co含量均低于浸渍法制备的催化剂, 而表面Pd含量则均高于浸渍法制备的催化剂, 且溶胶-凝胶法制备的催化剂各种产物的生成速率均高于浸渍法制备的催化剂, 因此, 与浸渍法制备的催化剂相比, 溶胶-凝胶法制备的催化剂具有更好的催化活性。  相似文献   

4.
采用电化学方法在玻碳(GC)表面电沉积CoNi合金纳米粒子,成功制得碳载CoNi合金纳米电极(CoNi/GC)。SEM结果显示,CoNi粒子呈十八面体结构,粒径约100 nm,分布较均匀。选区电子衍射(SAED)结果显示,CoNi合金纳米粒子为单晶结构。XPS结果显示,金属态的Co(0)和Ni(0)占主导地位。性能测试结果表明:CoNi/GC不但对亚硝酸钠具有较好的催化性能,相对于本体Co和本体Ni,CoNi/GC的起始还原电位(Ei)正移约90 mV,还原峰电流(jp)增大6~14倍。而且对氧还原亦有较好的电催化活性,CoNi/GC的峰电流密度(jp)和动力电流密度(jk)分别是GC电极的1.7和 5.2倍。  相似文献   

5.
以Pd Cl2和Co(NO3)2为原料,采用分步乙二醇还原法制备了多壁碳纳米管负载Pd-Co复合纳米催化剂Pd-Co/CNT。利用TEM、XRD和XPS对催化剂的结构进行了表征,考察了其甲醇电氧化性能。结果显示,Co的引入使Pd催化剂的分散性得到改善,其电化学表面积可达39.7 m2/g。循环伏安测试表明,当Pd∶Co物质的量比为1∶0.2时,Pd-Co/CNT的甲醇氧化峰电流密度约为Pd/CNT的2.7倍。计时电流结果表明,Co的添加使催化剂的活性衰减比例由Pd/CNT的63.8%降至54.2%,显示出较强的抗中毒能力。Pd-Co复合催化剂性能的改善归因于Pd与Co之间的协同相互作用。  相似文献   

6.
分别以蔗糖和硅酸钠为碳源和硅源,硫酸为催化剂,采用溶胶-凝胶法和碳热还原法合成SiC;以其为载体,Pd为活性组分,CeO2为助剂,采用浸渍法制备Pd-CeO2/SiC催化剂。考察催化剂制备方法、助剂CeO2添加量和还原温度等对催化剂在CO和C3H6氧化反应中活性的影响。结果表明, 采用分步浸渍法制备的催化剂具有较高活性, CeO2添加质量分数2%时制备的0.5%Pd-2%CeO2/SiC催化剂经200 ℃预还原后,对CO和C3H6氧的最低完全转化温度分别为210 ℃和215 ℃,CeO2助剂可显著提高金属Pd在催化剂表面分散度,与Pd产生相互作用,使Pd还原峰向低温移动,提高了催化剂活性。  相似文献   

7.
以镍(Ni)为金属节点,腺嘌呤(A)和柠檬酸(CA)为有机配体,采用溶剂热法制备了非晶态金属有机配合物Ni-A-CA。将Pd Cl2溶液浸渍于载体Ni-A-CA后用NaBH4还原制得高分散的Pd纳米粒子(Pd NPs)催化剂Pd/Ni-A-CA。通过SEM、TEM、XRD、FTIR、XPS、N2吸附-脱附对载体Ni-A-CA[n(Ni)∶n(A)∶n(CA)=2∶1∶1.5]和催化剂3%Pd/Ni-A-CA(3%为Pd的理论负载量,以Ni-A-CA的质量计,下同)进行了表征。结果表明,Pd NPs高度分散在载体Ni-A-CA上,其粒径为(2.2±0.3) nm,且载体与Pd NPs之间存在的强相互作用增强了催化剂的催化性能。在90℃、2 MPa H2条件下,3%Pd/Ni-A-CA催化喹啉加氢反应70 min,喹啉转化率为99.0%,生成1,2,3,4-四氢喹啉选择性>99%。  相似文献   

8.
张申  郭玉玉  李星颖  李哲 《化工进展》2019,38(2):885-891
采用浸渍法制备了一系列具有不同CuO含量的Pd-CuO/Al2O3催化剂,并将其用于乙醇氧化反应,其结构与性质通过XRD、H2-TPR和NH3-TPD等手段进行分析。结果发现,催化剂的活性并不是随着CuO含量的增加而增强,Pd-1.0%CuO/Al2O3催化剂表现出最佳的活性,其点火温度和完全转化温度比Pd/Al2O3催化剂至少降低了50℃。与Pd/Al2O3催化剂相比,含CuO催化剂增强的衍射峰强度以及氢化钯分解峰的消失,说明Pd-Cu合金结构的形成有利于Pd、Cu物种之间的协同作用。对于Pd-1.0%CuO/Al2O3催化剂来说,还原峰向低温的移动以及还原峰面积的增大说明该催化剂上氧化性物质更易被还原且数量在增加,这对于氧化反应是十分有利的,新出现的还原峰表示Pd、Cu的相互作用生成了新物种。NH3-TPD结果中更高含量的低温酸有利于高活性,而且新出现的脱附峰说明形成了新的酸性位点。  相似文献   

9.
采用高温固相烧结法制备了粉体材料LaFeO3、LaFe0.75Co0.25O3、LaFe0.75Cr0.25O3和LaFe0.75Mn0.25O3,通过XRD、FTIR、SEM、XPS等检测手段对材料进行表征,同时运用CASTEP模块模拟计算了材料的电子能带结构和光学性质。实验结果表明:掺入Mn/Cr/Co离子后晶格发生畸变,晶格对称性降低。Mn/Cr/Co掺杂后的粉体材料在近中红外波段发射率排序为:掺Mn>掺Co>掺Cr>纯LaFeO3,其中掺Mn在0.2~2.5μm波段为0.8722,2.5~5μm波段为0.6755,远大于纯LaFeO3的发射率(近中红外波段发射率均为0.5左右)。发射率提升的机理在于:Mn掺杂后,引入了Mn3+杂质能级,产生了激活能小的Mn3+?Mn4+跳跃小极化子,电子-氧空位载流子吸收亦增强,同时体系的晶格畸变导致振动吸收加剧。第一性原理计算结果表明掺杂Mn/Cr/Co材料的禁带宽度分别0.793eV、2.406eV、1.722eV均小于纯LaFeO3的3.817eV,结合态密度计算结果分析其原因,主要Mn3d、Cr3d、Co3d轨道与O2p轨道杂化形成杂质能级,同时Mn3d、Cr3d、Co3d在导带也存在态密度峰,且峰的位置都比LaFeO3峰更靠近费米能级,作为新的导带底相当于缩短了价带顶到导带底之间的间隙宽度。LaFe0.75Mn0.25O3材料在近中红外波段的优异辐射性能表现,可作为耐高温抗氧化高发射率材料在高温热工炉窑具有潜在应用前景。  相似文献   

10.
在还原过程中,非负载型钴基催化剂堆积孔结构容易坍塌,从而使金属钴的比表面积大幅降低,活性中心暴露的数量减少。作者采用简单的水热法制备了Co(OH)2/Co3O4混合物相的非负载型钴纳米颗粒催化剂,用于费托合成反应性能研究。结果表明:相比于单一物相的Co(OH)2或Co3O4催化剂,混合物相的催化剂显示出更高的费托合成反应活性。XRD、TEM、BET、H2-TPR等表征方法揭示出Co(OH)2与Co3O4具有不同的还原性质,两者紧密结合有利于催化剂在还原后维持更大的比表面积,进而有利于更多活性位点暴露,显著提高催化剂的反应活性。  相似文献   

11.
In order to improve a “Three Function Catalysts Model”, the present paper deals with alumina based catalysts containing cobalt and palladium for the NO reduction by methane.

The deNOx temperature window was estimated by adsorption and subsequent desorption of NO in lean conditions. Two NOx desorption peaks were detected for both catalysts. For Pd(0.63)Co(0.58)/Al2O3, the two desorption peaks appeared at 205 and 423 °C, whereas for Pd(0.14)Co(0.57)/Al2O3, the maxima desorption temperature peaks were at 205 and 487 °C. In addition, NO oxidation was also studied to evaluate the catalyst first function. It was found that, the oxidation begins on Co–Pd/Al2O3 around 250 °C. On Pd(0.63)Co(0.58)/Al2O3, 8% of deNOx were found in the range of the second NOx desorption peak temperature (410 °C). During TPSR, CxHyOz species such as formaldehyde were detected. These oxygenate species are the reactive intermediate for deNOx by methane.  相似文献   


12.
The effect of high temperature reduction (HTR) in hydrogen (up to 1180 K) on the microstructure of 9 wt.-% Pd/CeO2 catalyst was studied by HRTEM and XRD methods. Reduction of the catalyst at or above 973 K caused severe recrystallization of CeO2 and Pd with simultaneous strong interaction between the two components appearing as three phenomena: epitaxial growth of small Pd particles on CeO2 (most frequently with [111]Pd[111]CeO2); decoration of large Pd particles with ordered CeO2 overlayer and expansion of the lattice parameter of Pd (by 2.1%). The origin of the Pd lattice expansion is discussed and diffusion of Ce species into the Pd lattice seems to be the most probable one. HTR caused also phase transformations in the ceria support. At 973 K and 1100 K, whole CeO2 was transformed into oxygen deficient CeOx phase exhibiting the same or similar structure but with expanded lattice parameter (by 2.8%). At 1180 K most ceria was transformed into hexagonal A-Ce23. The CeOx phase appeared to be stable in hydrogen and in vacuum at room temperature, but upon exposure to air at room temperature it rapidly reoxidised to CeO2. Ce2O3 also reoxidised to CeO2 but much slower. Another consequence of HTR at or above 773 K was formation of pits in CeO2 crystallites, mainly on (112)-type crystal faces. The pits (1–10 nm) exhibited well defined walls parallel to CeO2 lattice fringes and they could possibly constitute nucleation sites for strongly bonded, epitaxial oriented Pd particles.  相似文献   

13.
The solubility of Pd(NO3)2 in water is moderate whereas it is completely soluble in diluted HNO3 solution. Pd/MIL-101(Cr) and Pd/MIL-101-NH2(Cr) were synthesized by aqueous solution of Pd(NO3)2 and Pd(NO3)2 solution in dilute HNO3 and used for CO oxidation reaction. The catalysts synthesized with Pd(NO3)2 solution in dilute HNO3 showed lower activity. The aqueous solution of Pd(NO3)2 was used for synthesis of mono-metal Ni, Pd and bimetallic PdNi nanoparticles with various molar ratios supported on MOF. Pd70Ni30/MIL-101(Cr) catalyst showed higher activity than monometallic counterparts and Pd+ Ni physical mixture due to the strong synergistic effect of PdNi nanoparticles, high distribution of PdNi nanoparticles, and lower dissociation and desorption barriers. Comparison of the catalysts synthesized by MIL-101(Cr) and MIL-101-NH2(Cr) as the supports of metals showed that Pd/MIL-101-NH2(Cr) outperforms Pd/MIL-101-(Cr) because of the higher electron density of Pd resulting from the electron donor ability of the NH2 functional group. However, the same activities were observed for Pd70Ni30/MIL-101(Cr) and Pd70Ni30/MIL-101-NH2(Cr), which is due to a less uniform distribution of Pd nanoparticles in Pd70Ni30/MIL-101-NH2(Cr) originated from amorphization of MIL-101-NH2(Cr) structure during the reduction process. In contrast, Pd70Ni30/MIL-101(Cr) revealed the stable structure and activity during reduction and CO oxidation for a long time.  相似文献   

14.
Calcined and reduced catalysts Pd/LaBO3 (B = Co, Fe, Mn, Ni) were used for the total oxidation of toluene. Easiness of toluene destruction was found to follow the sequence based on the T50 values (temperature at which 50% of toluene is converted): Pd/LaFeO3 > Pd/LaMnO3+δ > Pd/LaCoO3 > Pd/LaNiO3. In order to investigate the activation process (calcination and reduction) in detail, the reducibility of the samples was evaluated by H2-TPR on the calcined catalysts. Additionally, characterization of the Pd/LaBO3 (B = Co, Fe) surface was carried out by X-ray photoelectron spectroscopy (XPS) at each stage of the global process, namely after calcination, reduction and under catalytic reaction at either 150 or 200 °C for Pd/LaFeO3 and either 200 or 250 °C for LaCoO3. The different results showed that palladium oxidized entities were totally reduced after pre-reduction at 200 °C for 2 h (2 L/h, 1 °C/min). As LaFeO3 was unaffected by such a treatment, for the other perovskites, the cations B are partially reduced as B3+ (B = Mn) or B2+ even to B0 (B = Co, Ni). In the reactive stream (0.1% toluene in air), Pd0 reoxidized partially, more rapidly over Co than Fe based catalysts, to give a Pd2+/Pd4+ and Pd0/Pd2+/Pd4+ surface redox states, respectively. Noticeably, reduced cobalt species are progressively oxidized on stream into Co3+ in a distorted environment. By contrast, only the lines characteristic of the initial perovskite lattice were detected by XRD studies on the used catalysts. The higher activity performance of Pd/LaFeO3 for the total oxidation of toluene was attributed here to a low temperature of calcination and to a remarkable high stability of the perovskite lattice whatever the nature of the stream which allowed to keep a same palladium dispersion at the different stages of the process and to resist to the oxidizing experimental conditions. On the contrary, phase transformations for the other perovskite lattices along the process were believed to increase the palladium particle size responsible of a lower activity.  相似文献   

15.
A series of sulfated zirconia supported Pd/Co catalysts was synthesized by the sol–gel method and examined for NOx reduction by methane. The NO conversion increased up to a Co/S ratio of 0.43, and then decreased at a higher Co loading (Co/S = 0.95). Sulfate content was also essential for obtaining high selectivity to molecular nitrogen. A catalyst loaded with 0.06 wt.% Pd, 2.1 wt.% Co and 2.1 wt.% S (Pd/Co-SZ-2) exhibited remarkable performance under lean conditions and displayed stability in a long-term durability test using a synthetic reaction mixture containing 10% water vapor. This catalyst exhibited the highest sulfur retention most probably as cobalt sulfide. Besides, the catalytic oxidation of NO to NOy groups was confirmed by FT-IR, in agreement with the general mechanism for the SCR of NO by hydrocarbons. In the absence of oxygen in the feed stream, the catalyst was highly active for NO reduction with methane. IR stretching bands assigned to N2O and adsorbed nitro groups were identified upon adsorbing NO on Pd/Co-SZ-2. This indicates that under rich conditions disproportionation of NO to N2O and NO2 occurs and confirms that the formation of NO2 species is an essential step for NO reduction by CH4.  相似文献   

16.
Carbon-supported Pd-based binary alloy electrocatalysts (Pd–Co and Pd–Ni) with different particle sizes for polymer electrolyte fuel cells were prepared by a NaBH4 reduction method and investigated to examine effects of the size and lattice constant of the Pd alloy nanoparticles on the oxygen reduction reaction (ORR) activity. The particle size and lattice constant were controlled in the wide ranges 4.2–12.1 and 0.3802–0.3948 nm, respectively by heating the catalysts in specific atmospheres. The alloy structures were characterized by X-ray diffraction, transmission electron microscopy and X-ray absorption fine structure. The electrochemical tests of the Pd–Co/C and Pd–Ni/C catalysts were performed by cyclic voltammetry and rotating disk electrode in 0.1 M HClO4. Nearly linear relationship between the lattice constant and nanoparticle size was observed with the Pd–Co and Pd–Ni nanoparticles. The nanoparticle sizes and lattice constants of the Pd–Co/C and Pd–Ni/C electrocatalysts, which influence the Pd d-band center, showed positive and inverse relations with the ORR specific activities, respectively. The mass activities of the Pd–Co/C and Pd–Ni/C electrocatalysts showed an increasing trend with the lattice expansion.  相似文献   

17.
Ce-Zr solid solution (CexZr1-xO2, CZO) was prepared by the citric acid sol-gel method. The CZO was then used as a support for Pd/CZO catalysts for the oxidative carbonylation of phenol to diphenyl carbonate. The Pd/CZO catalyst showed enhanced activity and diphenyl carbonate selectivity compared with the Pd/CeO2 catalyst. The catalytic performance of Pd/CZO was influenced by the calcination temperature of the CZO support. X-ray diffraction, scanning electron microscopy, N2 adsorption-desorption measurements, X-ray photoelectron spectroscopy and H2 temperature-programmed reduction measurements were used to investigate the effects of Zr doping and calcination temperature. The catalytic performance of Pd/CZO and Pd/CeO2 for the oxidative carbonylation of phenol was affected by several factors, including the specific surface area, Ce3+ and/or oxygen vacancy content, oxygen species type and Pd(II) content of the catalyst. All these properties were influenced by Zr doping and the calcination temperature of the CZO support.  相似文献   

18.
Highly dispersed palladium nanoparticles containing mesoporous silicas MCM-41 and MCM-48 were prepared by one-pot synthesis. The method consists of the simultaneous formation of CTA+ surfactant templating MCM-41 mesophase and CTA+ micelle-capped PdO, which was reduced by hydrogen to Pd metal with particle size ≈ 2 nm and was observed to stay inside the mesochannels of MCM-41 (pore size ≈ 3.8 nm) by TEM, XAS, and PXRD. During hydrothermal synthesis of Pd/MCM-48, Pd nanoparticles of average size ≈ 6–7 nm were deposited on the MCM-48 of pore size = 4 nm. The deposition is probably derived from ethanol reduction of Pd(II) complex generated from PdCl2 precursor by hydrolysis of TEOS and C12H25(OCH2CH2)4OH surfactant. The formation of Pd(0) from Pd(II) species in solid mesoporous silicas by hydrogen reduction was monitored by in situ XAS, and compared with the formation of Pd(0) from [PdCl4]2−, [PdCl3(H2O)], and Pd(OH)2 by sodium dodecyl sulfate surfactant and alcohol reduction in aqueous solutions.  相似文献   

19.
Selective catalytic reduction of NO with methane (CH4-SCR) in the presence of oxygen excess and water vapour was studied over two bimetallic cobalt/palladium-based FER catalysts, which differ on the order of introduction of metal ions. H2-TPR and UV–vis analysis showed that the simple change in the order of addition of metals to catalyst, gives rise to totally diverse species (Co2+ ions, Co oxides, Co-oxo cations and Pd species) both in type and quantity but also in location within zeolite framework. Experiments of TPD and TPSR of NO and NO2 provided important information to establish a relation between the various active sites formed on both catalysts and their function in the reaction mechanism. The importance of NO2 in the mechanism of NO reaction with CH4 and O2 was explored and the catalyst with a higher capacity to retain adsorbed NO2 is the less active for deNOx. The preparation of a bimetallic catalyst active for NO reduction must provide the proximity between Co and Pd species, and the presence of Co-oxo cations together with palladium species seem to be essential. Furthermore, a suitable amount of cobalt oxides must exist in order to originate NO2 that is the main reaction intermediate. Nevertheless, an excessive amount of these Co species can lead to an increase of adsorbed NO2, which reduces the rate of the reaction of some of the mechanism steps.  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