文章快速检索    
  分子催化  2017, Vol. 31 Issue (1): 1-10
0

引用本文 

ZHU Jing, WANG Yan-ping, LI Zhe. The Complete Oxidation of Ethanol Over Porous Pd/TiO2-Al2O3 Catalysts[J]. Journal of Molecular Catalysis (China), 2017, 31(1): 1-10.
朱靖, 王彦平, 李哲. 多孔Pd/TiO2-Al2O3催化剂上乙醇完全氧化的研究[J]. 分子催化, 2017, 31(1): 1-10.

Foundation

Project supported by the State Key Program of National Natural Science of China (Grant No. 21336006); Shanxi Province Scientific and Technological Project (20140313002-2)

First author

Jing Zhu (1991-), Woman, master, Study on application and modification of the catalysts for the total oxidation of ethanol from emission of ethanol-gasoline vehicle

Corresponding author:

LI Zhe, E-mail:lizhe@tyut.edu.cn

文章历史

Received date: 2016-12-03
Revised date: 2017-01-19
The Complete Oxidation of Ethanol Over Porous Pd/TiO2-Al2O3 Catalysts
ZHU Jing, WANG Yan-ping, LI Zhe     
Abstract: Porous composite materials Al2O3-TiO2 calcined at different temperatures supported palladium catalysts were synthesized by template-impregnation technique and used for catalytic oxidation of ethanol. The as-prepared samples were cha-racterized by means of low angle and wide angle X-ray diffraction (XRD), N2 adsorption-desorption isotherms, Transmission electron microscopy (TEM), Fourier transformed infrared (FT-IR) and XPS. Pd/TiO2-Al2O3 calcined at 250 ℃ was much more active on the conversion of ethanol and CO2 yield than any other ones. High surface area, well-dispersed metal palladium nanoparticles and rich chemisorbed oxygen are probably responsible for the high catalytic activities.
Key words: porous Al2O3-TiO2 composites     palladium catalyst     ethanol oxidation    
多孔Pd/TiO2-Al2O3催化剂上乙醇完全氧化的研究
朱靖, 王彦平, 李哲     
太原理工大学 化学化工学院, 山西 太原 030024
摘要:利用模板剂-浸渍法制备出了不同焙烧温度下的多孔复合材料TiO2-Al2O3和Pd/Al2O3-TiO2催化剂来催化氧化乙醇.样品经过XRD, FT-IR, 孔结构分析、TEM、XPS、脉冲吸附、NH3-TPD等进行表征分析. 250 ℃焙烧的Pd/TiO2-Al2O3催化剂具有最高的乙醇转化率和CO2生成率.高比表面积、均匀分散的金属Pd颗粒和丰富的表面吸附氧是其具有高的催化活性的主要原因.
关键词: 多孔Al2O3-TiO2复合材料     Pd催化剂     乙醇氧化    

Volatile organic compounds (VOCs) have been considered the main contributor to air pollution[1].The conventional approach for VOCs destruction is catalytic combustion. In that field of catalysis, noble metal/ TiO2 catalysts have been extensively reported and stated to be highly efficient[2-3]. The reason of the highly appreciated noble metal/ TiO2 catalyst lies in two aspects. One is the unique semiconductive properties of TiO2 material, for the ability of being particle reduced can help activate oxygen. The other is the strong metal-support interaction between palladium and TiO2 which can inhibit agglomeration of metal particles. However, on the field of classical impregnation method to prepare Pd/TiO2 (P25), Vera P. Santos have found that although the total oxidation of ethanol could be obtained at the temperature of 180 ℃, the total conversion into CO2 can only be achieved at 260 ℃[4]. Hence, even though significant improvement has been made towards developing noble metal catalysts, challenge still remains. It is well known that total oxidation into CO2 of VOCs over noble metal catalysts should be occur at lower temperature.

Porous TiO2 materials especially mesoporous TiO2 now have been occupied an important position in the field of environmental technology, such as photocatalytic degradation of toxic organic contaminants, photocatalytic water pollutant and titania film might be used in Dye Sensitized Solar Cells[5]. Ordered porous TiO2 has attracted increasing interest for its high porosity and specific surface area, uniform aperture size and narrow pore size distribution[6-9]. While little literature has been reported about the mesoporous powder TiO2 used as support of noble metal catalyst. Actually the pore wall of orderly porous TiO2 is usually composed of nanometer TiO2 particles, which can combine the excellent characteristics of both ordered porous which is considered to be beneficial to the diffusion of the reactants and products and nanometer TiO2 which is well known for its effectiveness nontoxicity and chemical stability. However, there are some obvious drawbacks in applying porous TiO2, such as low thermal stability, low density of surface active site and weak surface acidity, which significantly limit the efficiency of catalysis[10-11]. Efforts have been paid to improve properties of TiO2 porous material. There are a number of strategies reported among which incorporation TiO2 with another metal oxide has been highlighted[12]. Cause composite materials can achieve the possession of high thermostability of anatase, orderly porous channel and defect TiO2 nanoparticles with abundant chemical pro-perties. Alumina, due to its affordability, easy accessibility and high chemical stability, has been extensively studied. The literature has investigated the effect of preparation method on the properties of Al2O3-TiO2 and found that the addition of Al2O3 to TiO2 can produce novel feature because of their unique chemical and physical properties[13-14].

Supported palladium catalyst is well-known for its high catalytic performance in the alcohol oxidation; however, development in catalytic selectivity has been in the quest and more efforts must be paid to enhance palladium based catalysts related to specific organic component reaction. An important factor concerning the VOCs oxidation over noble metal based catalysts is that oxidation of VOCs is considered to be structure sensitive reaction. At the base of fixed active components, turnover rates could be changed by the structure of catalysts. The influence of particle size on the catalytic combustion of different VOCs has been reported[4]. While as to ethanol oxidation over palladium catalyst, more efforts have to be taken to understand mechanism of ethanol oxidation for the connection between catalytic performance and catalyst properties depends mainly on the type of VOC.

In our previous work, nano Al2O3-TiO2 prepared by sol-gel method has been utilized as the support of Pd catalysts. And it found that Pd/Al2O3(5%)-TiO2 catalyst showed a higher catalytic activity than those of either Pd/TiO2 or Pd/Al2O3 for the total oxidation of ethanol. Nonetheless, only limited research attention has been devoted to coupling mesoporous TiO2 with Al2O3, and conducted the resulting titania-composite materials as the carrier of supported catalysts[15]. In this work, the template and sol-gel methods were employed for the preparation of mesoporous TiO2-Al2O3. Palladium as active component was impregnated on the composite support, and the activity and yield of CO2 of this catalyst produced by ethanol combustion was de-monstrated.

1 Experimental 1.1 Preparation of Al2O3 modified porous TiO2

Titania and alumina sols were prepared separately, and mixed together to obtain porous Al2O3-TiO2 (95%). The preparation details of is alumina sol was described in our previous work[16]. Titania sol was synthesized by a traditional template method[17]. Ti (OC4H9)4 was added to a HNO3 aqueous solution slowly under vigorous stirring at 313 K to get a transparent solution. Alumina sol was added into the system, and kept stirring for another 1 h. The transparent sol was separated from hydrolysis products and aged at ambient temperature for 3 h. Ethanol solution of dodecylamine was added into the transparent sol under vi-gorous stirring and then obtained a transparent sol which was treated with ultrasonic for 1 h, aged at 353 K for 24 h in a closed flask. The resulting white gel was transferred to a Petri dish and dried at 353 K for another 24 h to give a yellow powder. The yellow powder was rinsed with absolute ethanol and subsequently dried at 353 K overnight. The products were calcined from 523 to 673 K. Pd was loaded on the carriers by a wet-impregnation method, and all samples contained 1% Pd theoretically. The samples prepared in different conditions were designed as Pd/AT-n, Pd/T-n, in which n denotes different calcination temperatures, AT denotes the resulting titania-composite Al2O3-TiO2, and T denotes single TiO2.

1.2 Catalyst evaluation

The catalytic properties of the catalysts were eva-luated using a fixed-bed reactor connected with an online gas chromatograph. The compositions of ethanol and products were analyzed using a GC-950 apparatus with two FID detectors, one of which incorporated a methanator. Approximately 0.30 g of catalyst was used (0.450~0.280 mm), and the feed gas consisted of 0.5% C2H5OH, 5% air, and a balance of N2. The flow rate was 100 mL/min of total gas and the gas hourly space velocity (GHSV) was 24 000 h-1. Ethanol conversion (Xi) and the yields (Yj) of by-pro-ducts and carbon dioxide from ethanol combustion were calculated according to the formulas:

${X_i} = \frac{{{C_{{\rm{i, in}}}}-{C_{{\rm{i, out}}}}}}{{{C_{{\rm{i, in}}}}}}\;\;\;\;{Y_{\rm{j}}} = \frac{{{C_{j, {\rm{out}}}}}}{{{C_{{\rm{i, in}}}}}} \times \frac{1}{n} \times 100\% $

Where Ci and Cj are the concentrations of ethanol and products from ethanol combustion, respectively, n is the products and reactants carbon atom ratios. And

${C_{\rm{i}}} = {A_{\rm{i}}} \times {f_{\rm{i}}}\;\;\;{C_{\rm{j}}} = {A_{\rm{j}}} \times {f_{\rm{j}}}$

where A and f are the peak areas and the relative correction factors of reactants (i) or products (j), respectively.

1.3 Characterizations

Low-angle X-ray diffraction (XRD) patterns of all samples were collected in θ-2θ mode using Rigaku D/MAX-2550 apparatus (Cu Kα1 radiation, λ=1.540 6 Å), operated at 40 kV and 200 mA. Wide-angle XRD patterns were collected in the same mode, but operated at 100 mA. Intensities of the diffraction peaks were recorded in the 2θ range 5°~80° with a step size of 0.01°, and the scanning speed was 8 °/min.

Transmission electron micrographs (TEM) were carried out on a Tecnai G2F20, operating at 200 kV. The sample was diluted in ethanol and dropped onto a copper grid coated with a thin film of carbon.

The porous texture of the samples was analyzed by N2-adsorption at 77 K. By using a Micromeritics ASAP 2000 system, the BET and BJH methods were applied for the determination of the specific surface area, and the mean mesoporous equivalent diameter, respectively. Before measurement, samples were evacuated at 473 K.

FT-IR analysis was carried out using NICOLET360 type spectrometer. The spectra of solids were obtained using KBr pellets. Before measurement, the catalyst and potassium bromide were mixed together in accordance with the quality ratio 1:100. The vibrational transition frequencies are reported in wave numbers (cm-1).

XPS analysis was carried out on a Kratos Axis Ultra DLD multifunctional surface analysis photoelectron spectrometer with X-ray source at 450 W and Al-Kα radiation (1 486.6 eV). The element binding energies were referenced to the C 1s line at 284.6 eV and measurements were restricted an estimated error of ± 0.1 eV.

The Pd dispersion was determined by pluse H2 chemisorption on a Micrometrics AutoChem Ⅱ 2920 instrument with a TCD. Before measurement, 100 mg sample was pretreated by helium flow for 1h at 200 ℃ to remove gaseous species adsorbed, and then cooling down to room temperature. The H2/Pd average stoichiometry of 1 has been assumed for the calculation of dispersion[18].

NH3-TPD-MS experiment was performed on a TP-5080. Prior to experiment, 100 mg catalyst was pretreated at 300 ℃ for 30 minutes in He atmosphere to remove the adsorbed matter and then decreased to 50 ℃ to adsorb ammonia. Physically adsorbed ammonia was removed by purging with He. And then the temperature was increased linearly at 10 ℃/min from 50 to 810 ℃ to desorbed chemisorbed NH3 and the gas coming from the reactor be analyzed by QIC mass spectrometry detector.

2 Results and discussion 2.1 Phase and morphology analysis

A series of composite Al2O3-TiO2 porous materials and palladium based catalysts were synthesized by a template associated wet-impregnation method. Evidence for template removal of the as-prepared samples has been confirmed by FT-IR (Fig. 1).

Figure 1 FTIR spectra of different supports and catalysts calcined at different temperatures

Their low-angle XRD patterns are shown in Fig. 2a. The low-angle XRD patterns of Pd/T-150, Pd/AT-150, Pd/T-250, Pd/AT-250 appear only one broad peak respectively. The peaks at the lower angle suggest the existence of a large lattice plane distance (D value), which correspond to the mesoporous structure with not well-ordered pores in these materials. As only one peak is observed in the low-angle range, the mesoporous structure of the samples is typically of long-range order in nature[19]. In low-angle XRD patterns of mesoporous silica HMS with wormhole framework, only one broad peak could be observed, which was similar to that of our sample[20]. So the samples were not layered structures but wormhole framework structures and that has been proven by further characteristic method TEM (Fig. 2a insert). The XRD pattern of catalysts calcined at 400 ℃ does not show any peak, indicating no mesoporous structure in the samples[21]. This means that higher calcination temperature will lead to the collapse of the pore structure.

Figure 2 Low angle powder XRD (a), TEM of Pd/AT-250(a insert), and high angle powder XRD (b) patterns of supports and catalyst calcined at 250 ℃ patterns of catalysts calcined at 250 ℃

In order to confirm crystal structure of the synthesized mesoporous carriers and catalyst calcined at 250 ℃, XRD analysis was performed in a wide angle range, as shown in Fig. 2b. As seen in the patterns, the characteristic peaks of anatase phase appear in all samples, while the brookite phase appears in the single component carrier T-250 and the composite oxide carrier AT-250. It also can be seen that the intensity of the diffraction peak of brookite becomes weaker when Al2O3 was introduced into mesoporous TiO2, which means that the addition of Al2O3 would restrain the generation of brookite TiO2. No peak correlated to γ-Al2O3 or PdO was observed, probably due to the relatively low contents and/or high dispersion of γ-Al2O3 and PdO over the catalysts[22].

2.2 N2 adsorption analysis

In order to further examine the porous structure of supports and catalysts calcined at 250 ℃, N2 adsorption-desorption isotherms were employed. According to the IUPAC classification[23], the adsorption isotherms of all samples as shown in Fig. 3 exhibit the typical Ⅳ isotherm, which indicates the existence of mesoporous structure. Moreover, a clear absorption hysteresis loop located at 0.40 < P/P0 < 1.0 and a flat "S"-shape is observed in the isotherms of T-250 and AT-250, which belongs to the H1 type hysteresis loop, suggesting the presence of uniform pores in the samples. The curves of Pd/T-250 and Pd/AT-250 belongs to type H2, suggesting that the presence of slit-like pores in the samples. Fig. 3 insert shows the pore size distribution curve calculated by the BJH method from the desorption branch of a nitrogen isotherm. It can be seen that the diameter distribution of samples at 250 ℃ mainly concentrated in the ranges of mesopores from 3 to 20 nm, which can be concluded that the samples have good mesoporous structure.

Figure 3 Nitrogen adsorption-desorption isotherms and Pore size distribution (insert) of supports and catalysts calcined at 250 ℃

The BET surface areas, pore sizes and pore volumes of samples are summarized in Table 1. Without Al2O3, the BET surface areas of T-250 and Pd/T-250 were 187 and 127 m2·g-1, respectively. The surface areas of samples AT-250 and Pd/AT-250 were increased to 256 and 206 m2·g-1 respectively when Al2O3 was added. This may due to Al2O3 itself has a larger surface area. Meanwhile, the surface area of Pd/AT calcined at 400 ℃ sharply decreased to 69 m2·g-1, that could be ascribed to the higher calcination temperature, resulting in the collapse of the pore structure.

Table 1 Specific surface areas and pore structure parameters of supports and catalysts
2.3 Pd dispersion

Pd dispersion of different catalysts was shown in Fig. 4. It can be seen that the dispersion of palladium on the Pd/AT-250 was higher than that on Pd/AT-150 and Pd/AT-400, and the change feature is consistent with their surface area. It is known that higher dispersion of active components resulted in higher thermal stability and catalytic performance.

Figure 4 Pd dispersion of palladium catalysts calcined at different temperature
2.4 XPS

We carried out XPS to determine chemical states of elements Ti, O and Pd in pure TiO2, Al2O3-TiO2, catalysts Pd/TiO2 and Pd/Al2O3-TiO2 calcined at 250 ℃ and shown in Fig. 5. When compared Pd/Al2O3-TiO2 with Pd/TiO2 calcined at 250 ℃, there is no significant difference of chemical state of palladium species. And the increased catalytic performance of ethanol oxidation over palladium catalysts could be ascribed to Pd dispersion.

Figure 5 XPS spectrogram of Pd 3d (a); Ti 2p (b); O1s (c) in Pd/TiO2 and Pd/Al2O3-TiO2 calcined at 250 ℃

As shown in Ti 2p, the main peak of Ti 2p3/2 on the T-250 is located at 458.4 eV, which can be ascribed to Ti4+ of TiO2. In comparison, the corresponding peak showed a negative shift of about 0.4 eV over the AT-250. The shift of Ti 2p3/2 peak to a lower position suggests that particle Ti4+ might be existed in lower chemical state like Ti3+ for the introduction of Al2O3. It has been proven that at the presence of oxygen vacancies assistant with TiOx (x < 2), content of chemisorption oxygen can be enhanced[2].

According to asymmetry of O 1s spectra, we use peak fitting to further figure out chemical state of oxygen species at the surface of palladium catalysts. The main peak at around 530.0 eV could be attributed to the lattice oxygen of TiO2. And the O could be attributed to chemisorption oxygen. It can be seen in the O1s of catalyst Pd/T-250 and Pd/AT-250, the peak position of O shifts to lower energy by about 0.4 eV compared with Pd/T-250, and this is consistent with Ti 2p curves. And the content of chemisorption oxygen over Pd/AT-250 was significantly larger than that of Pd/T-250. The well-dispersed palladium species and rich chemisorbed oxygen is closely related to the activity of Pd catalysts on ethanol oxidation.

2.5 NH3-TPD measurement

To analyze acidity in sample, NH3-TPD profiles corresponding to the sample calcined at 150 and 250 ℃ is shown in Fig. 6. It is obviously to see a desorbed peak at low temperature which was related to weak acidity[24]. And the peak area is almost unchanged between different catalysts indicating the weak acidity is consistent with each other. While samples calcined at 150 ℃ shows a broad desorption peak at temperature range 400~600 ℃ which is attributed to strong acidity resulted from the residual template under low calcination temperature. And the peak area was a little different between Pd/TiO2 and Pd/AT which was for the different acidity of Al2O3 and TiO2. Reference has shown that catalyst acidity was introduced by support than metal. Additionally, for supported Pd catalyst, the larger amount of strong acidity would lead to difficulty in Pd oxidation and restrain ethanol oxidation[25]. And this could be an aspect for lower catalytic performance of catalysts calcined at 150 ℃.

Figure 6 NH3-TPD of catalysts calcined at different temperature
2.6 Catalytic activity evaluate

The ethanol catalytic oxidation activity over different Pd catalysts was evaluated. Evidence of by-product yields has shown the route of ethanol oxidation among various catalysts (Fig. 7c-f). The first step is ethanol dehydrogenized into acetaldehyde, which was then at the presence of oxygen further transformed into CO and CO2. It is found that the difference between various catalysts lies in the yield of CO. CO yield on Pd/AT-250 is apparent lower than that on any other catalysts and that could be due to the high specific area. The catalytic performance of different catalysts was investigated by detecting the oxidation proportions of ethanol and yield to CO2 and the results are shown in Fig. 7. It is found that ethanol conversion and yield to CO2 on Pd/TiO2-Al2O3 were higher than that on the others, and that might be due to the higher surface area and the relatively higher Pd dispersion. From the profiles in Fig. 7(a), the significant decrease of ethanol conversion and CO2 yield on Pd/TiO2 when reaction temperature was higher than 150 ℃(calcined temperature) could also be of another proof for the destroyed structure and decreased surface area.

Figure 7 Catalytic activity, yield of CO2 and other products from ethanol oxidation over catalysts calcined at different temperatures

The catalyst Pd/AT-250 exhibited the highest activity and achieved complete oxidation of ethanol at 275 ℃. While at reaction temperature below 150 ℃, it can be seen that the conversion of ethanol over catalysts calcined at 150 ℃ (Pd/T-150) is higher than that at higher calcination temperature. Calcination can help remove organic components and also do harm to porous structure. The more ordered porous might be beneficial for reactant and surface active sites interaction. Additionally, the introduction of aluminium could restrain crystalline growth. And efforts have been paid to prepare anatase at low temperature for the stable phase[26]. In general, anatase is more active than rutile and brookite for catalytic activity. So it can be proposed that the Pd/T-150 with relatively higher catalytic performance than that on the others is due to the stable anatase and the relation between anatase and surface active sites. While it can be seen the extremely lower performance when reaction temperature is higher than 200 ℃ over Pd/T-150. High temperature treatment would cause residual organic species diffusion to form compound resulting in sintering of active sites and deteriorating catalytic activity[26]. Fig. 7(b) presents the yield of CO2 over Pd-based catalysts calcined at diffe-rent temperatures, which was constant with catalytic activity of ethanol conversion. The catalyst Pd/AT-250 had the best yield to CO2, and its yield to CO2 reached 80% at 275 ℃. The porous TiO2 supported Pd catalyst with high BET surface area can improve the dispersion of surface Pd active sites. The higher dispersed active sites can improve catalytic efficiency towards ethanol convert to CO2. Moreover, the abundant chemisorbed oxygen atoms can enhance the relation between oxethyl produced at the first step of ethanol oxidation reaction and oxygen to form CO2.

2.7 Stability test

The stability test under different temperature and different time of ethanol oxidation over Pd/AT-250 and Pd/T-150 was shown in Fig. 8. The flow rate was 100 mL/min of total gas and the gas hourly space velocity (GHSV) was 24 000 h-1.

Figure 8 stability test under different temperature and different times of ethanol oxidation over Pd/AT-250 and Pd/T-150

The test temperature range is 75~500 ℃. For catalyst used to remove emission of vehicle, it should maintain high catalytic performance at high temperature and long work time. From Fig. 8a, we can see the temperature of total oxidation of ethanol is 275 ℃ over Pd/AT-250 while that over Pd/T-150 is 325 ℃ and it decreased obviously as the temperature increased continuously. It is noting that at temperature range of 275~400 ℃, the higher catalytic performance towards ethanol conversion and CO2 yield could be maintained over Pd/AT-250 which means that the relatively higher thermostability of Pd/AT-250, while as the temperature is higher than 400 ℃, the ethanol conversion and CO2 yield is decreased. The decreased thermostability of Pd based catalyst could be attributed to the relatively lower support structure, and higher reaction temperature leads to carbon deposition and active sites sintering. From Fig. 8b, we carried out stability test at different time. It could be found that at reaction time ethanol conversion decreased from 98.7% to 96.2% and the average decreased rate is 0.25%/h-1 which revealed the higher stability.

While the ethanol conversion and CO2 yield over Pd/T-150 is decreased from 84.5% to 8.29% which means the lower stability, revealing that the increased calcination temperature and the introduction of alumi-nium could improve the stability of Pd/TiO2 catalyst.

3 Conclusions

A series of wormhole-like porous material Al2O3-TiO2 and Pd/Al2O3-TiO2 catalysts have been synthesized by template-impregnation technique. The catalyst Pd/Al2O3-TiO2 calcined at 250 ℃ exhibited the highest activity and yield of CO2 and that can be attributed to the high surface area, larger pore size, higher-dispersed palladium species and the abundant chemisorbed oxygen.

Acknowledgements: Authors acknowledge the financial support from Project supported by the State Key Program of National Natural Science of China (Grant No. 21336006), and Shanxi Province Scientific and Technological Project (20140313002-2).
参考文献
[1] a. Ghadikolaei M A. Effect of alcohol blend and fumigation on regulated and unregulated emissions of IC engines-A review[J]. Renew Sust Ener Rev, 2016, 57: 1440-1495.
b. Zhang Jian-qi (张剑琦), Li Li (李莉), Liu Di (柳迪), et al. Preparation of three-dimensionally ordered macroporous composite ZrO2-TiO2 and its photocatalytic degradation of organic pollutantsunder multiple modes 3DOM (TiO2-ZrO2复合材料制备与多模式光催化降解有机污染物) [J]. J Mol Catal (China)(分子催化), 2015, 29 (4): 348-358.
c. Du Jun-chen (杜君臣), Chang Shi-ying (常仕英), Huang Wei-qiang (黄卫强), et al. Progress of the Pd catalysts for methane oxidation under low temperature (甲烷低温氧化Pd催化剂的研究进展) [J]. J Mol Catal (China)(分子催化), 2015, 29 (5): 482-493.
[2] a. Huang H, Leung D Y C. Complete oxidation of formaldehyde at room temperature using TiO2 supported metallic Pd nanoparticles[J]. ACS Catal, 2011, 1 (4): 348-354.
b. Zhang Li-na (张丽娜), Deng You-quan (邓友全), Shi Feng (石峰). Preparation of Fe-doped TiO2 for the selective oxidation of aromatic alcohols with oxygen under visible light irradiation (Fe掺杂改性TiO2的制备及其可见光催化醇氧化性能研究) [J]. J Mol Catal (China)(分子催化), 2015, 29 (2): 179-187.
c. Sun Sheng-nan (孙圣楠), Li Chun-hu (李春虎), Yang Wei-wei (杨微微), et al. Photocatalytic removal of NO from flue gas by TiO2 loaded on semi-coke prepared by sol-gel method (溶胶-凝胶法制备TiO2负载活化半焦光催化烟气脱硝) [J]. J Mol Catal (China)(分子催化), 2015, 29 (2): 188-196.
d. Wang Ying-jun (汪颖军), Xi Hui-yao (席慧瑶), Suo Yan-hua (所艳华), et al. Preparation methods of Ni-WO3/TiO2-ZrO2 and study on their properties (不同方法制备Ni-WO3/TiO2-ZrO2催化剂及其性能研究) [J]. J Mol Catal (China)(分子催化), 2015, 29 (5): 414-421.
[3] Huang H, Ye X, Huang H, et al. Mechanistic study on formaldehyde removal over Pd/TiO2 catalysts: Oxygen transfer and role of water vapor[J]. Chem Engineer J, 2013, 230: 73–79. DOI:10.1016/j.cej.2013.06.035
[4] a. Bezemer G L, Bitter J H, Kuipers H P C E, et al. Cobalt particle size effects in the fischer-tropsch reaction studied with carbon nanofiber supported catalysts[J]. J Amer Chem Soc, 2006, 128 (12): 3956-3964.
b. Guo Zhi-min (郭志敏), Yuan Jian (袁坚), Peng Xiao-jin (彭小晋), et al. The effect of sodium salts on catalytic activity of V2O5-WO3/TiO2 over NH3-SCR reaction (钠盐对V2O5-WO3/TiO2催化剂NH3-SCR反应活性的影响) [J]. J Mol Catal (China)(分子催化), 2016, 30 (6): 547-556.
[5] Mahoney L, Koodali R. Versatility of evaporation-induced self-assembly (EISA) method for preparation of mesoporous TiO2 for energy and environmental applications[J]. Materials, 2014, 7(4): 2697–2746. DOI:10.3390/ma7042697
[6] Zhu C, Guo C, Dai Z. Fabrication and study of porous metals with a highly regular structure[J]. Mater Lett, 2013, 94: 189–192. DOI:10.1016/j.matlet.2012.12.034
[7] Pabst W, Gregorová E. Conductivity of porous materials with spheroidal pores[J]. J Euro Cera Soc, 2014, 34(11): 2757–2766. DOI:10.1016/j.jeurceramsoc.2013.12.040
[8] Holland B T, Blanford C F, Stein A. Synthesis of macroporous minerals with highly ordered three-dimensional arrays of spheroidal voids[J]. Science, 1998, 281(5376): 538–540. DOI:10.1126/science.281.5376.538
[9] Lukens W W, Yang P, Stucky G D. Synthesis of mesocellular silica foams with tunable window and cell dimensions[J]. Chem Mater, 2001, 13(1): 28–34. DOI:10.1021/cm991109c
[10] Liu Y, Li Y, Wang Y, et al. Sonochemical synthesis and photocatalytic activity of meso-and macro-porous TiO2 for oxidation of toluene[J]. J Hazar Mater, 2008, 150(1): 153–157. DOI:10.1016/j.jhazmat.2007.04.088
[11] Elder S H, Gao Y, Li X, et al. Zirconia-stabilized 25-Å TiO2 anatase crystallites in a mesoporous structure[J]. Chem Mater, 1998, 10(10): 3140–3145. DOI:10.1021/cm980237i
[12] Liu S, Liu G, Feng Q, et al. Al-doped TiO2 mesoporous materials: synthesis and photodegradation properties[J]. J Poro Mater, 2009, 17(2): 197–206.
[13] Toba M, Mizukami F, Niwa S I, et al. Effect of preparation methods on properties of alumina/titanias[J]. J Mater Chem, 1994, 4(4): 585–589. DOI:10.1039/jm9940400585
[14] Zhang W, Li R, He H, et al. Synthesis of mesoporous TiO2-Al2O3 binary oxides photocatalyst by sol-gel method using PEG1000 as template[J]. Inter J Photo, 2012, 2012: 1–7.
[15] Zhu G, Yang B, Wang S, et al. Nanocrystallites-forming hierarchical porous Ni/Al2O3-TiO2 catalyst for dehydrogenation of organic chemical hydrides[J]. Intern J Hydro Ener, 2011, 36(21): 13603–13613. DOI:10.1016/j.ijhydene.2011.07.112
[16] Li Z, Quan Y, Chang Y, et al. Peparation and characterization of nanoscale γ-Al2O3/TiO2 composite[J]. J Mol Catal (China), 2007, 21(5): 417–422.
[17] Sheng Q, Yuan S, Zhang J, et al. Synthesis of mesoporous titania with high photocatalytic activity by nanocrystalline particle assembly[J]. Micro Mes Mater, 2006, 87(3): 177–184. DOI:10.1016/j.micromeso.2005.06.036
[18] Peng J, Wang S. Performance and characterization of supported metal catalysts for complete oxidation of formaldehyde at low temperatures[J]. Appl Catal B: Enviro, 2007, 73(3/4): 282–291.
[19] Wang X, Yu J C, Ho C, et al. Photocatalytic activity of a hierarchically macro/mesoporous titania[J]. Langmuir, 2005, 21(6): 2552–2559. DOI:10.1021/la047979c
[20] Pauly T R, Pinnavaia T J. Pore size modification of mesoporous HMS molecular sieve silicas with wormhole framework structures[J]. Chem Mater, 2001, 13(3): 987–993. DOI:10.1021/cm000762t
[21] Smarsly B, Grosso D, Brezesinski T, et al. Highly crystalline cubic mesoporous TiO2 with 10-nm pore diameter made with a new block copolymer template[J]. Chem Mater, 2004, 16(15): 2948–2952. DOI:10.1021/cm0495966
[22] Zhao J, Wang Y, Lou X, et al. Effects of adding Al2O3 on the crystal structure of TiO2 and the performance of Pd-based catalysts supported on the composite for the total oxidation of ethanol[J]. Inorg Chim Acta, 2013, 405: 395–399. DOI:10.1016/j.ica.2013.06.028
[23] Yu J C, Zhang L, Yu J, et al. Direct sonochemical preparation and characterization of highly active mesoporous TiO2 with a bicrystalline framework[J]. Chem Mater, 2002, 14(11): 4647–4653. DOI:10.1021/cm0203924
[24] González C A, Bartoszek M, Martin A, et al. Hydrodechlorination of light organochlorinated compounds and their mixtures over Pd/TiO2-washcoated minimonoliths[J]. Indus & Engineer Chem Res, 2009, 48(6): 2826–2835.
[25] Liu Z F, Tabora J, Davis R J, et al. Relationships between microstructure and surface acidity of Ti-Si mixed oxide catalysts[J]. J Catal, 1994, 149(1): 117–126. DOI:10.1006/jcat.1994.1277
[26] Shen Q, Li S, Zong J, et al. Low-temperature preparation of macroporous anatase TiO2 thin film and its photocatalytic activity[J]. Rare Met Mater Engineer, 2010, 39(A2): 36–39.