Synthesis of Mg-Doped Ordered Mesoporous Pd–Al2O3 with Different

Jan 25, 2017 - Employing NH4HCO3 as pore-enlarging agent and P123 as a template agent, ordered mesoporous Mg-doped γ-Al2O3 with different basicity an...
1 downloads 13 Views 3MB Size
Subscriber access provided by University of Newcastle, Australia

Article

Synthesis of Mg-doped ordered mesoporous Pd-Al2O3 with different basicity for CO, NO and HC elimination Yihong Xiao, Xiaohai Zheng, Xiaohua Chen, Lilong Jiang, and Ying Zheng Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/acs.iecr.6b03799 • Publication Date (Web): 25 Jan 2017 Downloaded from http://pubs.acs.org on January 26, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Industrial & Engineering Chemistry Research is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Synthesis of Mg-doped ordered mesoporous Pd-Al2O3 with different basicity for CO, NO and HC elimination Yihong Xiaoa, Xiaohai Zhenga,b, Xiaohua Chenb, Lilong Jianga*, Ying Zhenga,b a

National Engineering Research Center of Chemical Fertilizer Catalyst, Fuzhou University, Fuzhou, Fujian 350002, China

b

College of Materials Science and Engineering, Fujian Normal University, Fuzhou, Fujian 350007, China

*Corresponding author: Fax: +86 591 83707796, Tel: +86 591 83731234-8201, E-mail: [email protected]

Abstract: Employing NH4HCO3 as pore-enlarging agent and P123 as a template agent, ordered mesoporous Mg-doped γ-Al2O3 with different basicity and large surface area have been successfully synthesized by a facile sol–gel approach by adjusting the content of magnesium and pH value. Specifically, the result alumina calcined at 1000 °C with large specific surface area (234.4 m2 g–1) and high pore volume (0.72 cm3 g–1) was obtained when the Mg content was 9 wt%. These materials with ordered mesostructure and advantageous structural properties were utilized as carriers of Pd-Al2O3 catalysts for catalytic reaction of simulated exhaust automobile gases including CO, NO and hydrocarbon (HC). The results revealed that the high catalytic performance of Pd/9Mg-OMA originated from its more appropriate basicity, greater PdO dispersion and higher surface area.

Keywords: Sol–gel method; Ordered mesostructure; Mg-doped; Surface area; Acid-base; Palladium catalyst.

1 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1. Introduction Since the environmental laws and regulations of exhaust emissions have become more and more strict, the catalysts with more efficient catalytic performances were required, especially well acid-base property and low temperature activity. Three-way catalysts (TWCs) have been widely used,1 which provides an efficient way to reduce automobile exhaust of carbon monoxide (CO), nitrogen oxides (NOx) and hydrocarbons (HC). The working temperature of gasoline engines generally higher than 1000 ºC, which makes commonly used catalyst support γ-Al2O3 phase transition, sintering2. Therefore, many considerable attentions has been paid to the improvement of thermal stability of γ-Al2O3. Ordered mesoporous alumina (OMA) could effectively restrain the phase transformation from γ-Al2O3 to α-Al2O3 because of its excellent texture properties, such as large BET surface area, tunable pore sizes and long-range ordering of the pore packing, which makes it extensively applied as catalyst carrier.3 Niesz4 firstly managed to synthesize OMA with P123 as the structure-directing agent via a sol-gel based self assembly technique. Yuan et al.5 reported a modified sol-gel method to the synthesis of OMA with highly ordered mesoporous structure, but the specific surface area only remains 116 m2 g–1 after calcined at 1000 °C. Moreover, the application of the catalysts with OMA as carrier has also gained much attention. Li6 applied a facile route to prepared OMA-supported catalysts with Pt or Pd nanoparticles well-dispersed, which exhibits high thermal stability and high catalytic activity when applied in CO oxidation reaction. The evaporation-induced self-assembly (EISA) process was employed by Bordoloi7 to

2 ACS Paragon Plus Environment

Page 2 of 47

Page 3 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

achieve

ordered

mesoporous

mixed

cobalt-aluminum

oxides

with

Co2+

homogeneously distributed in an alumina matrix. The well-ordered 15Co-Al2O3 composite turned out to be highly active in CO oxidation. The surface acid–base properties of the catalyst carrier will also strongly affect the performance of catalyst. Therefore, the alkaline earth metals (including Be, Mg, Ca, Sr or Ba) were added to improve the surface acid–base properties and catalytic performances of the catalysts. The MgO is alkali metal oxides (Hammet constant H = +26.0), which is usually selected as carriers or basic modifiers for Pd supported catalysts.8 Moreover, the thermal stability of alumina can be improved by doping magnesium according to the reference.9 Arnby10 yielded catalysts modified by Mg with high thermal stability through incipient wetness impregnation by using platinum nitrate solution and Mg(NO3)2 as precursor compounds. The low-temperature activity of Pt/γ-Al2O3 catalysts for carbon monoxide oxidation was slightly enhanced by adding magnesium. Yamaguchi11 prepared the Mg-Al composite oxides through the calcination of hydrotalcites. The composite oxide with the mole ratio of Mg/Al=5 possessed the highest catalytic activity when acted as efficient catalysts for CO2 reaction and styrene oxide because of the synergistic reaction of basic and acidic sites. Xu12 yielded ordered mesoporous OM-xMgyAl composite oxides with high thermal stability through one-pot EISA method, the catalysts exhibited high catalytic activities toward the reaction for CO2 reforming of CH4. Herein, a series of thermally stable Mg-doped OMA with strong basicity and large BET surface area were facilely prepared through a simple sol–gel method. The

3 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

increase of BET surface area, pore volume and pore sizes can be attributed to the increase of micellar template sizes as well as the release of CO2 and NH3 in the calcination process by introducing of NH4HCO3. Ordered mesoporous Mg-doped Pd-alumina catalysts were prepared by impregnation method with Mg-doped OMA as carriers. The effect of Mg content on catalytic properties and physicochemical characteristics of the catalysts were investigated. The results reveal that these materials with advantageous structural properties and different basicity show higher catalytic activity for conversion of CO, NO and hydrocarbon (HC) compared with undoped sample.

2. Experimental 2.1. xMg-OMA supports As following, the required amounts of Mg(NO3)2 and approximately 1.00 g EO20PO70EO20 triblock copolymer (Mn=5800, Aldrich) were dissolved in a mixed solution containing anhydrous ethanol (20 mL) and hydrochloric acid (1.0 mL, 37 wt%) at room temperature. And others were then added into the solution in the molar ratio of NH4HCO3: acetic acid: aluminum isopropoxide (AIP) = 0.1: 0.15: 1 under vigorous agitation for at least 5 h, adjusting the final pH for 1.0~1.2. The above solution was dry in air at 60 °C for 6 h and 80 °C for 8 h to evaporate the solvent, respectively. Afterwards, the calcination was taken with the temperature increasing from 25 °C to 550 °C (1 °C min–1 ramping rate) for 4 h in air. Then they were calcined at required temperature (800 °C or 1000 °C) by using a heating rate of 10 °C min–1 during 1 h. The 0, 3, 6, 9, 12 and 15 wt% Mg doped samples were referred as

4 ACS Paragon Plus Environment

Page 4 of 47

Page 5 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

xMg-OMA (x stands for the Mg content in wt%). 2.2. Catalyst preparation Supported Pd catalysts on xMg-OMA were synthesized through incipient wetness impregnation route with 6 h impregnation in Pd(NO3)2 aqueous solution (0.028 g/mL) at room temperature. Then the catalyst precursors were dried at 110 °C for 6 h and this was followed by calcination at 800 °C or 1000 °C for 2 h by using a 10 °C min–1 ramping rate. Then the as-synthesized catalysts were crushed and sieved (20-40 mesh particles). The Pd loading was 0.5 wt% (theoretic value). All the as-prepared catalysts were designated as Pd/xMg-OMA in the following text. 2.3. Characterization techniques X-ray diffraction (XRD) measurements were collected on a Philips MPD diffractometer (X'Pert Pro) using Cu Kα radiation (λ=0.15046 nm), operated at 40 mA and 40 kV. Texture properties of the samples were conducted by N2 adsorption–desorption experiments through an U.S. Micromeritics ASAP 2020M analyzer. Differential thermal analysis (DTA) measurements were accomplished in a perkin-Elmer 1700 device under air atmosphere. Transmission electron microscopy (TEM) studies were performed with an FEI Tecnai G2 F20 S-Twin (Netherlands) transmission microscope operated at 200 kV. X-ray photoelectron spectroscopy (XPS) spectra were acquired in a Thermo ESCALAB 250 spectrometer (USA) using Al Kα X-ray radiation (hυ= 1486.6 eV), the data of which was corrected with the C1s binding energy (BE) at 284.8 eV. CO2 temperature programmed desorption (CO2-TPD) studies were presented at

5 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 47

AutoChem 2920 instrument (Micromeritics) employing a thermocouple detector TCD detector to clarify the adsorption properties of CO2 on catalysts. The sample (0.3 g) was pretreated at 400 °C and held for 1 h in He stream (10 °C min–1, 30 mL min–1) to remove the impurities and moisture, and then cooled down to 50 °C. Then a pure CO2 flow was pumped into the reactor for 1 h. Finally, the CO2-TPD measurements were performed from 50 °C to 800 °C at a heating rate of 10 °C min–1 under He flow. CO pulse chemisorption experiments were accomplished on the same apparatus as described for CO2-TPD. Approximately 0.3g of each sample was pretreated in a pure He flow from 30 °C to 400 °C for 1h at a heating rate of 10 °C min–1, then reduced in the gas flow of 10% H2/Ar (30 mL min–1) at 400 °C during 1 h followed by cooling down to 25 °C to carried out CO chemisorption. The dispersions, the metallic surface area, and particle diameter of PdO were recorded by pulse method. H2

temperature-programmed

reduction

(H2-TPR)

measurements

were

accomplished in a chemisorption analyzer same as CO2-TPD. The catalyst (about 0.3 g) was pre-treated in a flow of helium from 30 °C to 400 °C for 0.5 h to remove water vapor. The TPR profiles were acquired by pumping a 10% H2/Ar (30 mL min–1) through the preheated catalyst with the temperature increasing from 30 °C to 900 °C (10 °C min–1). In situ-diffuse reflectance infrared Fourier transform spectroscopy (in situ-DRIFTS) were performed by using Nicolet-6700 FTIR spectrometer with an MCT detector. The spectra were obtained at 2 cm−1 resolution and an accumulation of 64 scans. The catalysts were pretreated with He at 300 °C for 0.5 h to remove

6 ACS Paragon Plus Environment

Page 7 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

impurities. (1) CO adsorption experiments were carried out under 1%CO/He for 1h and it was purged by helium for 0.5 h to record the spectra again. The sample was first reduced by 10% H2/Ar during 1 h at 300 °C. (2) CO2 chemisorbed experiments were taken after pure CO2 flow was then fed into the reactor for 30 min. The DRIFT spectra were collected at different temperatures. 2.4. Catalytic performance measurement All tests were performed in an evaluation apparatus of simulate automobile exhaust. The reaction mixture containing CO (0.89%), O2 (0.76%), CO2 (11.5%), NO (800 ppm), HC (150 ppm, C3H6) and balance gas (N2) were introduced into the reactor at 45,000 h−1 GHSV, λ=1. A 0.3 g catalyst (20—40 mesh) was held in a quartz tube (Φ 8 mm) by packing quartz wool at the end of the catalyst bed with 210 mL min–1 flow rate. The light-off (T50%) and complete conversion temperature (T100%) was recorded and the CO, NO and HC conversions were calculated by the following equations: CO conversion =

[CO]in - [CO]out ×100% [CO]in

NO conversion =

[NO]in - [NO]out ×100% [NO]in

HC conversion =

[HC]in - [HC]out ×100% [HC]in

3. Results and discussions 3.1. XRD and N2 adsorption measurements The small-angle X-ray diffraction (SXRD) pattern of Mg-doped OMA shown in Fig. 1 was used to assess the formation of ordered mesostructures. The samples

7 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

calcined at 800 °C display two diffraction peak corresponding to the (100) and (110) reflections at around 0.71° and 1.08°, giving a hint that the obtained xMg-OMA possess a 2D hexagonal structure with a P6mm space group.13 With the rise of Mg content, the intensity of the (100) diffraction line of xMg-OMA first increases and then decreases. The phenomenon suggests that a moderate content of Mg is in favor of the higher ordering degree. Simultaneously, the peaks move toward relatively lower values with the increasing of Mg, indicating the segmental existence of Mg2+ in the Al2O3 lattice influence the cell parameter (a0) and d-spacing (d100) values, which were calculated in accordance with Meng14 (Supplementary materials, Table S1). The result can be attributed to the incorporation of Mg species into the matrix of Al2O3 resulting the formation of Mg–O–Al bonds. As for the samples calcined at 1000 °C (Fig. 1B), the diffraction peaks show a obvious shift to a higher angle, which indicates a constriction of the grain size after heated at high temperature.15 The wide-angle XRD patterns of xMg-OMA with different magnesium content were presented in Fig. 2A. It was noteworthy that the un-doped OMA calcined at 800 °C display six diffraction peaks assigned to the (220), (311), (222), (400), (440) and (511) reflections of γ phase alumina16 (JCPDS 10-0425) at around 32.9°, 37.1°, 39.5°, 45.5°, 60.7° and 67.0°, respectively. As for Mg-doped OMA, the intensity of diffraction peaks decrease with the Mg content gradually increasing. When the Mg content is 15 wt%, only one weak diffraction peak is observed in 15Mg-OMA. The results suggest that the crystal temperature of γ-alumina can be improved by Mg doping. When calcined at 1000 °C, all Mg-doped OMA show remarkable MgAl2O4

8 ACS Paragon Plus Environment

Page 8 of 47

Page 9 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

diffraction peaks (JCPDS 87-0345).17 It can be noted that the position of the diffraction peaks shifts to lower 2θ values with increasing Mg content, the lattice parameters of samples also increase, which indicates that Mg species incorporated into the matrix of alumina lattice.18 Besides, all samples heated at 1000 °C (Fig. 2B) display higher intensity of diffraction peaks than these heated at 800 °C (Fig. 2A), which suggests that all of them possess higher crystallinity. Table 1 gives the textural parameters of xMg-OMA calcined at 1000 °C. The BET surface area and pore volume of xMg-OMA gradually increase with increasing the Mg content (3-9 wt%) in comparison with undoped OMA. Meanwhile, 9Mg-OMA displays larger BET surface area (212.3 m2 g–1) and bigger pore volume (0.56 cm3 g–1) than others, the value is also higher than some other Mg-doped OMA under the same calcined temperature reported by Pan19. With the Mg content exceeding 9 wt%, both of the BET surface area and pore volume of xMg-OMA were decreased, indicating that only a small amount of modifier can improve the textural parameters.20 To further improve the pore volume and BET surface area, the NH4HCO3 was introduced into the synthesis system of xMg-OMA. It can be observed from Table 1 that all the samples with NH4HCO3 possess higher BET surface area and larger pore volume than those without NH4HCO3, the improvement of textural properties on the one hand can be attributed to the enlargement micellar template dimensions derived from the existence of additional NH4HCO3 molecules between alumina precursor and P123. On the other hand the enlargement of BET surface area and pore size is due to the change of the intracrystalline porosity during thermal

9 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

decomposition (the release of CO2 and NH3).21 Moreover, the pore size of all samples with adding NH4HCO3 also increase, the results suggest that the pore size can be controlled in a wide range. Therefore, the Pd species may penetrate into the pore due to the larger pore size of support. Taking the well textural properties of samples with adding NH4HCO3 into consideration, the N2 adsorption-desorption isotherms of xMg-OMA are presented in Fig. 3. All the samples performed classic IV-type isotherms with a clear H1-shaped hysteresis loops, a typical feature for mesoporous materials with regularly cylindrical pores according to the sharpness of the capillary condensation step.22 The pore size distributions for supports with adding NH4HCO3 calcined at 1000 °C, derived from the desorption branch by using BarrettJoyner-Halenda (BJH) calculations, which shows that Mg-OMA-9 represented a narrower pore size distribution around 9.8 nm (Fig. 3 (B)). DTA analysis shown in Fig. 4 was performed to study the thermal stability of as-synthesized xMg-OMA. There are two clear exothermic peaks in all curves. As for Mg-OMA-0, the DTA curve of which shows a clear peak located at 849 °C, which can be ascribed to the formation of the γ phase alumina. Another exothermic peak centered at 1154 °C is assigned as the phase transformation of γ-alumina to stable α-Al2O3.23 The position of peaks for the Mg-doped OMA shifts to higher temperature with increasing content of Mg. The results can be ascribed to the formation of MgAl2O4 delay the migration of Al3+, which restrains the phase transformation of alumina.24 As is known to all that γ-alumina possesses a spine structure, in which two

10 ACS Paragon Plus Environment

Page 10 of 47

Page 11 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

kinds of vacancies (tetrahedral and octahedral) randomly assigned on its surface. The ionic radius of Al3+ is smaller than Mg2+ and it can insert into the vacancies of γ-Al2O3, resulting in lattice expansion18. As a result, the cationic deficiency of γ-Al2O3 decreased, which improves the thermal stability of obtained alumina. In view of the fine structure properties of xMg-OMA, the TEM analysis presented in Fig. 5 was carried out to confirm the presence of well-defined hexagonal mesostructures with p6mm symmetry. It can be clearly observed that 9Mg-OMA displays the alignment of highly ordered hexagonal pores along [001] (insets) and the arrangement of parallel cylindrical channels along [110] after being calcined at 800 °C (Fig. 5A),25 which is well consistent with the SXRD results (Fig. 1A). After the calcined temperature increases to 1000 °C, its ordered mesoporous structure can still be well retained (Fig. 5B). The results reveal that the obtained xMg-OMA possesses highly ordered mesostructure and high thermal stability.

3.2. Catalytic activity of the catalysts The complete-conversion temperature (T100%) of HC, CO and NO for the catalysts Pd/xMg-OMA are presented in Table 2. It can be observed from the table that whether the catalysts heated at 800 °C or 1000 °C, the complete-conversion temperature of Pd/xMg-OMA lower than that of un-doped samples, indicating that the doping of Mg enhance the catalytic activities of Pd supported alumina. It is also noticeable that the complete-conversion temperatures of Pd/9Mg-OMA for all the target pollutants lower than others. Combined with the textural parameters (Table 1) and thermal stability (Fig. 4) of supports as well as the Pd dispersion (Table 4) of all

11 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 47

catalysts, the conclusion is draw that the large surface area and high thermal stability of supports are in favor of the Pd dispersion, which makes the target pollutants full contact with active component and brings about the improvement of the catalytic performance.26, 27 While excessive Mg content (12 and 15 wt%) may decrease the catalytic performance due to it can weaken the interaction of Pd–Al2O3. Therefore, one can draw a conclusion that the excessive content of magnesium may produce a negative impact on the catalytic activity. After being calcined at 1000 °C, the catalytic activities of the catalysts suffered a certain degree of decrease. This phenomenon was suggested to be fairly related to the high temperature treatment, which makes the active sites of Pd evaporation and sintering.28 It is worth noting that the complete-conversion temperature (T100%) of HC and NO for the catalysts Pd/9Mg-OMA are 326.3 and 342.8 °C, both of them are lower than the results over Pd-Ce0.7Zr0.3O2-Al2O329 and Pd/(Ce, Zr)Ox–Al2O330 three-way catalysts. On the other hand, the well catalytic activities can be attributed to the number of exposed active centers to the reactants on the mesopores surface and the stabilized active sites from the confinement effect of the mesostructure of alumina.31 The existence of confinement in the mesostructure of the xMg-OMA is conducive to prevent particles of PdO from sintering under high temperature treatment and catalytic reaction. From this point of view, the Pd supported catalysts prepared with xMg-OMA as carrier can effectively improve the catalytic activity. Meanwhile, MgO is usually selected as basic supports or modifiers, which improves the surface acidity and basicity of catalysts

and

generally

results

in

enhancing

the

12 ACS Paragon Plus Environment

catalytic

activity.

The

Page 13 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

characterizations of CO2-TPD and CO2 in situ-DRIFTS were taken below to further investigate theirs surface acidity and basicity.

3.3. Catalytic stability of the catalysts The evaluation of the long-term stabilities of the catalysts was carried out under given conditions: 280 °C, 210 mL min–1, catalyst 0.3g, GHSV=45,000 h−1, λ=1. Fig. 6 illustrates the profiles of CO, HC and NO conversions with time on stream over the Pd/9Mg-OMA catalyst. Pd/9Mg-OMA calcined at 800 °C exhibits high catalytic activities (e.g., around 99 %, 96 % and 91% for the conversions of HC, CO and NO, respectively) and stable catalytic behavior in the whole 30 h time on stream. It is worth noting that the Pd/9Mg-OMA calcined at 1000 °C showed stable activity during 30 h of reaction, the catalytic conversions of HC, CO and NO were 90%, 85% and 80% after 5 h reaction. Besides, the conversion of NO on this catalyst reach 80 % after 3 h reaction and then decrease a little (5%) after 30 h time on stream. The results suggest that the Pd/9Mg-OMA catalyst exhibits favorable long-term stability during the whole 30 h studied, implying a promising catalyst for CO, NO and HC elimination.

3.4. Acid-base properties CO2-TPD analysis collected in Fig. 7 was conducted to determine the basicity and base strength of the as prepared catalysts. Commonly, it was known to us that the adsorption of CO2 on the weak basic sites could be desorbed under low temperature and that desorbed at higher temperature would be called strong basic sites according to Pino.32 It can be seen that the TPD profiles for all the Pd/xMg-OMA catalysts were

13 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

similar in the shapes. It was observed that the OMA without doping magnesium promotion behaved a strong desorption peak at ca. 154 °C, which would be assigned to the weakly chemical adsorption of CO2 in the framework and physisorption of CO2 on account of high BET surface areas and pore volumes.33 It was also observed that the desorption peaks showed some shift to higher temperatures with magnesium content increasing from 3% to 9% while that of Pd/12Mg-OMA and Pd/15Mg-OMA suffered some decline. The results reveal that the suitable amount of the Mg modifier contributed to intensifying the basic intensity. As for the desorption peaks around 298 °C and 543 °C, they ought to be related to the chemisorbed CO2 over moderate and strong intensities basic sites.34 On the other hand, the amount of Mg obviously affected the intensities of these desorption peaks. The quantitatively distribution of surface basic sites of different strength of the catalysts derived from CO2-TPD profiles are summarized in Table 3. Among the Pd supported catalysts, the basicity of Pd/Mg-OMA-9 is larger than others. It is interesting found that the amount of CO2 desorption for all basic sites of Pd/Mg-OMA-x (x = 0, 3, 6, and 9) retained gradually increased with an increase of magnesium content. The phenomenon can be ascribed to the fact that Mg2+ has incorporated into the Al3+ lattice and creates a surface defect to make up for the formation of positive electric charge, which makes the adjacent surface oxygen ion becomes unsaturated, bring about a generation of surface magnesium aluminate with highly basicity.35 With the magnesium content further increasing, the amount of CO2 desorption suffered some decline because of the charge compensation effect on the surface above-mentioned became less important. The

14 ACS Paragon Plus Environment

Page 14 of 47

Page 15 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

result can be attributed to the formation of bulk magnesium aluminate spinel of the Pd/xMg-OMA (x = 12 and 15) as shown in WXRD patterns (Fig. 2B). It is noteworthy that the decrease of BET surface area of 12Mg-OMA and 15Mg-OMA could be another important cause for the decline of basicity.36 Therefore, it can be deduced that Pd/9Mg-OMA possess higher basicity compared to the catalysts with other magnesium content due to the formation of basic surface magnesium aluminate and high BET surface area.

3.4. In situ DRIFTS spectra of chemisorbed CO2 Fig. 8 presents the In situ DRIFTS spectra obtained on Pd/xMg-OMA after CO2 adsorbed at 30 °C and evacuated at 180 °C. There are three surface species of adsorbed state CO2 reflecting different types of surface alkaline sites were detected on the Pd supported catalysts. Bicarbonate formation (weak-strength basic sites) involves surface hydroxyl groups, which shows an asymmetric O-C-O stretches mode at 1410 cm–1. Unidentate and bidentate carbonate (strong-strength alkaline sites) formation requires surface alkaline oxygen atoms.37 It shows an asymmetric O-C-O stretches centered in 1510 cm–1. An asymmetric O-C-O stretches around 1635 cm−1 can be ascribed to bidentate carbonates (medium-strength alkaline sites).38 After evacuation at 180 °C, bicarbonate on the surface of the catalysts disappears while the unidentate and bidentate carbonates still maintained. It was also observed that Pd/9Mg-OMA possess higher band intensity after CO2 adsorption at 30 °C or evacuation at 180 °C. The results suggest that the basic sites strength increase with the appropriate Mg content were added in, which is consistent with the results of CO2-TPD analysis.

15 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

3.5. CO pulse chemisorption The CO pulse chemisorption was performed to test Pd dispersion of the catalysts, the results of which are summarized in Table 4. The dispersion, metallic surface area and particle sizes of active component were calculated using average CO:Pd adsorption stoichiometries of 1:1.39 As for the catalysts treated at 800 ºC, the metallic surface area and the dispersion of Pd gradually increase with the Mg content increasing, while the active particle sizes gradually decrease. Pd/9Mg-OMA shows a higher dispersion value (34.3%) and may be attributed to the strong interaction between Mg-doped OMA support and Pd. On one hand, the incorporation of Mg species into Al2O3 lattice with the formation of Mg–O–Al makes the defects of Al2O3 increases and which could be the strong site for Pd supported catalysts. From another point of view, the mesopore barriers are beneficial to protect the Pd particles located at the pore surface of Pd/xMg-OMA from growing.40 Besides, the high BET surface area and highly ordered mesostructure of Pd/Mg-OMA-9 may be another reason for the high dispersion of active component, leading to the higher catalytic activity than other catalysts.

3.6. In situ DRIFTS spectra of chemisorbed CO Representative in situ-DRIFTS spectra of CO adsorption are presented in Fig. 9 to reveal the chemical state of PdOx species. In the DRIFT spectra of CO adsorption (Fig. 9A), the main bands in the range of C–O stretches frequency could be assigned to linear carbonyl groups with Pd2+ complexes (Pd2+–CO) at 2040-2230 cm–1; another band for Pd/0Mg-OMA centered at 1915 cm–1 are associated with bridging carbonyl

16 ACS Paragon Plus Environment

Page 16 of 47

Page 17 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

ligands with Pd+ complexes (Pd+2–CO).41 It can be observed that the wave numbers of catalysts gradually shift to higher positions with the increase of Mg content from 3-9 wt% and the Pd/9Mg-OMA shows a clear band at 1948 cm–1. This phenomenon may be due to the higher BET surface area of Mg-doped OMA and smaller PdO particle sizes make it more likely to interact with Mg modified support.42 It is also found that the band at about 1988 cm–1 of Pd/9Mg-OMA displays a clearly stronger intensity than other catalysts. One possible reason is that there is relatively much more number of active sites on the mesopore surface of Pd/9Mg-OMA highly dispersed,43 which is well consistent with dispersion of the Pd species shown in Table 4. After getting rid of CO from the catalysts, the bands related to bridge-bound carbonyl CO still maintained while the peaks intensity of linearly bound CO weakened. The reducibility of Pd supported catalysts is one of the influences on catalytic activities.44 The H2-TPR profiles of Pd/xMg-OMA catalysts are established in Fig. 10. One peak centering at 82 °C was observed on Pd/0Mg-OMA, which can be assigned to the PdO species on the mesopores surface of the catalyst carrier easily reduced under lower temperature.45 It is also found that the reduction peak gradually shifted to lower temperature with increasing the Mg content, which indicates that the presence of magnesium improves the reducibility of PdO species.46 This behavior is attributable to the highly dispersion of Pd and the strong interaction between Mg-doped OMA carrier and PdO. These results indicate that the co-existence of Mg and Pd result in a synergetic effect, which improves the reducibility of Pd/xMg-OMA catalysts.

17 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

3.7. Pd dispersion and chemical state Fig. 11 gives the TEM images of the Pd/9Mg-OMA catalyst calcined at 800 °C and 1000 °C. After heating at 800 °C, the catalyst possessed highly ordered alignment of mesopores along [001] direction, which may be a reason for the high catalytic activity of Pd/9Mg-OMA due to the ordered structure was beneficial to the full access between reaction gas and active component. For the catalyst calcined at 1000 °C, the ordered mesostructure still could be maintained though accompanied by some damage, which fully demonstrates that the obtained catalyst possesses highly thermal stability. It is interesting to find that there is a particle on the catalyst surface (marked by a circle), the interplanar spacing (ca. 0.26 nm) is determined to be correspoding to the PdO (101) lattice fringes47. Besides, the particle diameter estimated from the image were found to be 6.4 nm, which is close to that of value obtained from the results of CO pulse chemisorption (Table 4).

3.8. XPS spectra characterization The XPS study on the Pd/0Mg-OMA, 9Mg-OMA and Pd/9Mg-OMA shown in Fig. 12 was carried out to investigate the chemical state of Mg, O, Al and Pd. The binding energy (BE) of Mg 2p spectra of 9Mg-OMA and Pd/9Mg-OMA are 50.8 and 51.1 eV, both of them are assigned to Mg2+ in Mg–O–Al bonds.48 The BE value of Al 2p for 9Mg-OMA and Pd/9Mg-OMA are 73.6 and 73.9 eV, both of them are lower than pure alumina (74.2 eV). The BE value of O 1s for 9Mg-OMA (531.9 eV) and its catalyst (532.1 eV) are slightly different from that of pure MgO (529.8 eV), γ-alumina (531.1 eV) and PdO (530.7 eV). The Mg–O–Al bonds are usually suggested to be connected with the formation of the spinel crystalline phase (MgAl2O4), and 18 ACS Paragon Plus Environment

Page 18 of 47

Page 19 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

strengthen the interaction between Mg and Al in the catalysts. This interaction is present in the crystal lattice of samples, in which Mg occupies tetrahedral sites and Al octahedral.49 XPS spectra of Pd/0Mg-OMA and Pd/9Mg-OMA presented a Pd 3d5/2 peak located at around 337.1 and 337.3 eV, which corresponds to the BE values of PdO species.50 The migration to the higher binding energy can be attributed to the interaction between support and active component resulting from the introduction of Mg. The turnover frequencies (TOF) values of Pd-supported catalysts were shown in Table 5. The dispersion of active metal was determined from the results of CO chemisorption (Table 4). It can be clearly observed from the catalysts calcined at 800 °C that the TOF values firstly increased as Mg loading on the support increased and then decreased when the Mg content more than 9 wt%. As for catalysts with different Mg content calcined at 1000 °C, the TOF values of HC, NO, and CO was observed to decrease when the content of Mg is larger than 9 wt%. The results may be attributed to a great deal of active component nanoparticles on the surface of Pd/9Mg-OMA,51 which consistent with the results of catalytic performance (Table 2) and Pd dispersion (Table 4) for the catalysts with different content of Mg.

4. Conclusions In conclusions, the preparation of ordered mesoporous Mg-doped γ-aluminas with different basicity via a facile sol-gel approach through using P123 as template and NH4HCO3 as pore-enlarging agent was reported. The as-prepared thermally stable OMA supports with favorable textural and structural characteristics were utilized as

19 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the Pd based catalysts toward the catalytic converters of HC, CO and NO elimination. It was observed that the introduction of magnesium effectively enhanced the BET surface area and thermal stability. It is of interesting to confirm that Pd/9Mg-OMA catalyst with moderate basic modifier and advantageous texture parameters displays better catalytic performances than other catalysts. From the above discussions, these favorable advantages make ordered mesoporous Mg-doped Pd-alumina a potential catalyst for purification of automobile exhaust and the treatment of dye-containing wastewater.

Supporting Information Interplanar spacing and lattice parameter of samples prepared with different Mg content calcined at different temperatures (Table S1); Small-angle (A) and wide-angle (B) XRD patterns for the catalysts prepared with different content of Mg calcined at 1000 °C (Figure S1); Interplanar spacing and lattice parameter of catalysts prepared with with NH4HCO3 and different Mg content calcined at different temperatures (Table S2).

Author Information Corresponding Author *E-mail: [email protected]

Notes The authors declare no competing financial interest

Acknowledgements The authors are grateful to financial support from the National key research and

20 ACS Paragon Plus Environment

Page 20 of 47

Page 21 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

development program (2016YFC0203903), the Project (class A) in Fujian province department of education (No. JA14060) and Industry-university-institute cooperation projects of Fujian province (No. 2014H6015).

21 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References (1) Prasad, V.S.; Aghalayam, P. Microkinetic modeling of the effects of oxygen on the catalytic reduction of NO on Pt and Rh in automotive after treatment. Ind. Eng. Chem.

Res. 2016, 55, 9362−9371. (2) Cho, K.M.; Park, S.; Seo, J.G.; Youn, M.H.; Nam, I.; Baeck, S.H.; Chung, J.S.; Jun, K.W.; Song, I.K. Effect of calcination temperature of alumina supports on the wax hydrocracking performance of Pd-loaded mesoporous alumina xerogel catalysts for the production of middle distillate. Chem. Eng. J. 2009, 146, 307–314. (3) Abidli, A.; Hamoudi, S.; Belkacemi, K. Synthesis, characterization and insights into stable and well organized hexagonal mesoporous zinc-doped alumina as promising metathesis catalysts carrier. Dalton Trans. 2015, 44, 9823–9838. (4) Niesz, K.; Yang, P.; Somorjai, G.A. Sol-gel synthesis of ordered mesoporous alumina. Chem. Commun. 2005, 15, 1986–1987. (5) Yuan, Q.; Yin, A.X.; Luo, C.; Sun, L.D.; Zhang, Y.W.; Duan, W.T.; Liu, H.C.; Yan, C.H. Facile synthesis for ordered mesoporous γ-aluminas with high thermal stability.

J. Am. Chem. Soc. 2008, 130, 3465–3472. (6) Li, Z.X.; Shi, F.B.; Li, L.L.; Zhang, T.; Yan, C.H. A facile route to ordered mesoporous-alumina-supported catalysts, and their catalytic activities for CO oxidation. Phys. Chem. Chem. Phys. 2011, 13, 2488–2491. (7) Bordoloi, A.; Sanchez, M.; Noei, H.; Kaluza, S.; Großmann, D.; Wang, Y.; Grünert, W.; Muhler, M. Catalytic behaviour of mesoporous cobalt-aluminum oxides for CO oxidation. J. Catal. 2014, 2014, 807545.

22 ACS Paragon Plus Environment

Page 22 of 47

Page 23 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

(8) Han, S.J.; Bang, Y.; Lee, H.; Lee, K.; Song, I.K.; Seo, J.G. Synthesis of a dual-templated MgO-Al2O3 adsorbent using block copolymer and ionic liquid for CO2 capture. Chem. Eng. J. 2015, 270, 411–417. (9) Kumar, P.A.; Reddy, M.P.; Hyun-Sook, B.; Phil, H.H. Influence of Mg addition on the catalytic activity of alumina supported Ag for C3H6-SCR of NO. Catal. Lett. 2009,

131, 85–97. (10) Arnby, K.; Törncrona, A.; Skoglundh, M. Influence of ammonia on CO and methanol oxidation over Pt/γ-Al2O3 catalysts modified by Mg. Appl. Catal. B:

Environ. 2004, 49, 51–59. (11) Yamaguchi, K; Ebitani, K; Yoshida, T; Yoshida, H; Kaneda, K. Mg-Al mixed oxides as highly active acid-base catalysts for cycloaddition of carbon dioxide to epoxides. J. Am. Chem. Soc. 1999, 121, 4526–4527. (12) Xu, L; Song, H; Chou, L. Ordered mesoporous MgO-Al2O3 composite oxides supported Ni based catalysts for CO2 reforming of CH4: Effects of basic modifier and mesopore structure. Int. J. Hydrogen. Energ. 2013, 38, 7307–7325. (13) Li, L.L.; Duan, W.T.; Yuan, Q.; Li, Z.X.; Duan, H.H.; Yan, C.H. Hierarchical γ-Al2O3 monoliths with highly ordered 2D hexagonal mesopores in macroporous walls. Chem. Commun. 2009, 41, 6174–6176. (14) Meng, Y.; Gu, D.; Zhang, F.; Shi, Y.; Yang, H.; Li, Z.; Yu, C.; Tu, B.; Zhao, D. Ordered mesoporous polymers and homologous carbon frameworks: amphiphilic surfactant templating and direct transformation. Angew. Chem. 2005, 117, 7215–7221. (15) Wu, Q.; Zhang, F.; Yang, J.; Li, Q.; Tu, B.; Zhao, D. Synthesis of ordered

23 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 47

mesoporous alumina with large pore sizes and hierarchical structure. Micropor.

Mesopor. Mater. 2011, 143, 406–412. (16) Semeykina, V.S.; Parkhomchuk, E.V.; Polukhin, A.V.; Parunin, P.D.; Lysikov, A.I.; Ayupov, A.B.; Cherepanova, S.V.; Kanazhevskiy, V.V.; Kaichev, V.V.; Glazneva, T.S.; Zvereva, V.V. CoMoNi catalyst texture and surface properties in heavy oil processing. Part I: hierarchical macro/mesoporous alumina support. Ind. Eng. Chem.

Res. 2016, 55, 3535–3545. (17) Shuang, L.; Xiaodong, W.; Duan, W.; Rui,R. NOx-assisted soot oxidation on Pt−Mg/Al2O3 catalysts: magnesium precursor, Pt particle size, and Pt−Mg interaction.

Ind. Eng. Chem. Res. 2012, 51, 2271−2279. (18) Sun, L.B.; Tian, W.H.; Liu, X.Q. Magnesia-incorporated mesoporous alumina with crystalline frameworks: a solid strong base derived from direct synthesis. J. Phys.

Chem. C 2009, 113, 19172–19178. (19) Pan, D., Dong, Z., He, M., Chen, W., Chen, S., Yu, F., Fan, B., Cui, X., Li, R. Structural

and

surface

properties

of

highly

ordered

mesoporous

magnesium-aluminium composite oxides derived from facile synthesis. Mater. Chem.

Phys. 2016. 22, 1–10. (20) Sharma, S.K.; Kushwaha, P.K.; Srivastava, V.K.; Bhatt, S.D.; Jasra, R.V. Effect of hydrothermal conditions on structural and textural properties of synthetic hydrotalcites of varying Mg/Al ratio. Ind. Eng. Chem. Res. 2007, 46, 4856–4865. (21) Hu, X.; Liu, Y.; Tang, Z.; Li, G.; Zhao, R.; Liu, C. Fabrication of high-surface-area γ-alumina by thermal decomposition of AACH precursor using

24 ACS Paragon Plus Environment

Page 25 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

low-temperature solid-state reaction. Mater. Res. Bull. 2012, 47, 4271–4277. (22) Liu, P.; He, S.; Wei, H.; Wang, J.; Sun, C. Characterization of α-Fe2O3/γ-Al2O3 catalysts for catalytic wetperoxide oxidation of m-cresol. Ind. Eng. Chem. Res. 2015,

54, 130–136. (23) Zhou, Z.; Zeng, T.; Cheng, Z.; Yuan, W. Preparation and characterization of titania−alumina mixed oxides with hierarchically macro-/mesoporous structures. Ind.

Eng. Chem. Res. 2010, 50, 883–890. (24) Kirszensztejn, P.; Przekop, R.; Szymkowiak, A.; Mackowska, E.; Gaca, J. Preparation of MgO-Al2O3 binary gel system with mesoporous structure. Micropor.

Mesopor. Mater. 2006, 89, 150–157. (25) Xu, J.; Wang, A.; Wang, X.; Su, D.; Zhang, T. Synthesis, characterization, and catalytic application of highly ordered mesoporous alumina-carbon nanocomposites.

Nano. Res. 2011, 4, 50–60. (26) Zhu, Z.; Liu, H.; Sun, H.; Yang, D. Surfactant assisted hydrothermal and thermal decomposition synthesis of alumina microfibers with mesoporous structure. Chem.

Eng. J. 2009, 155, 925–930. (27) Duan, K.; Chen, B.; Zhu, T.; Liu, Z. Mn promoted Pd/TiO2-Al2O3 catalyst for the selective catalytic reduction of NOx by H2. Appl. Catal. B: Environ. 2015, 176, 618–626. (28) Yahyaei, B.; Azizian, S. Rapid adsorption of anionic dyes by ordered nanoporous alumina. Chem. Eng. J. 2012, 209, 589–596.

25 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 47

(29) Shen, M., Yang, M., Wang, J., Wen, J., Zhao, M., Wang, W. Pd/support interface-promoted Pd−Ce0.7Zr0.3O2−Al2O3 automobile three-way catalysts: studying the dynamic oxygen storage capacity and CO, C3H8, and NO conversion. J. Phys.

Chem. C 2009, 113, 3212-3221. (30) Zhao, B., Yang, C., Wang, Q., Li, G., Zhou, R. Influence of thermal treatment on catalytic

performance

of

Pd/(Ce,Zr)Ox–Al2O3

three-way catalysts. J.

Alloy.

Compd. 2010, 494, 340-346. (31) Xu, L.; Song, H.; Chou, L. One-Pot Synthesis of ordered mesoporous NiO–CaO–Al2O3 composite oxides for catalyzing CO2 Reforming of CH4. ACS Catal.

2012, 2, 1331−1342. (32) Pino, L.; Vita, A.; Cipitì, F.; Laganà, M.; Recupero, V. Hydrogen production by methane tri-reforming process over Ni–ceria catalysts: effect of La-doping. Appl.

Catal. B: Environ. 2011, 104, 64−73. (33) Ashok, J.; Kathiraser, Y.; Ang, M.L.; Kawi, S. Bi-functional hydrotalcite-derived NiO–CaO–Al2O3 catalysts for steam reforming of biomass and/or tar model compound at low steam-to-carbon conditions. Appl. Catal. B: Environ. 2015, 172, 116–128. (34) Zhang, G.; An, H.; Zhao, X.; Wang, Y. Preparation of Ca−Zn−Al oxides and their catalytic performance in the one-pot synthesis of dimethyl carbonate from urea, 1,2-propylene glycol, and methanol. Ind. Eng. Chem. Res. 2015, 54, 3515−3523. (35) Han, S.J.; Bang, Y.; Kwon, H.J.; Lee, H.C.; Hiremath, V.; Song, I.K.; Seo, J.G. Elevated temperature CO2 capture on nano-structured MgO–Al2O3 aerogel: effect of

26 ACS Paragon Plus Environment

Page 27 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Mg/Al molar ratio. Chem. Eng. J. 2014, 242, 357–363. (36) Min, J.E.; Lee,Y.J.; Park, H.G.; Zhang, C.; Jun, K.W. Carbon dioxide reforming of methane on Ni–MgO–Al2O3 catalysts prepared by sol–gel method: Effects of Mg/Al ratios. J. Ind. Eng. Chem. 2015, 26, 375–383. (37) Al-Daous, M.A.; Manda, A.A.; Hattori, H. Acid–base properties of γ-Al2O3 and MgO–Al2O3 supported gold nanoparticles. J. Mol. Catal. A-Chem. 2012, 363, 512–520. (38) Cosimo, J.I. Di.; Dıez, V.K.; Xu, M.; Iglesia, E.; Apesteguıa, C.R. Structure and surface and catalytic properties of Mg-Al basic oxides. J. Catal. 1998, 178, 499–510. (39) Kuhn, M.; Lucas, M.; Claus, P. Long-time stability vs deactivation of Pd−Ag/Al2O3 egg-shell catalysts in selective hydrogenation of acetylene. Ind. Eng.

Chem. Res. 2015, 54, 6683−6691. (40) Shi, J.; Zhou, Y.; Zhang, Y.; Zhou, S.; Zhang, Z.; Kong, J.; Guo, M. Synthesis of magnesium-modified mesoporous Al2O3 with enhanced catalytic performance for propane dehydrogenation. J. Mater. Sci. 2014, 49, 5772–5781. (41) Smith, M.L.; Kumar, N.; Spivey, J.J. CO adsorption behavior of Cu/SiO2, Co/SiO2, and CuCo/SiO2 catalysts studied by in situ DRIFTS. J. Phys. Chem. C 2012,

116, 7931–7939. (42) Yang, L.; Yang, X.; Lin, S.; Zhou, R. DRIFTS study of the role of alkaline earths in promoting the catalytic activity of HC and NOx conversion over Pd-only three-way catalysts. Catal. Sci. Technol. 2015, 5, 1495–1503. (43) Zhang, Y.; Cai, Y.; Guo, Y.; Wang, H.; Wang, L.; Lou, Y.; Guo, Y.; Lu, G.; Wang,

27 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Y. The effects of the Pd chemical state on the activity of Pd/Al2O3 catalysts in CO oxidation. Catal. Sci. Technol. 2014, 4, 3973–3980. (44) Li, G.; Zhao, B.; Wang, Q.; Zhou, R. The effect of Ni on the structure and catalytic behavior of model Pd/Ce0.67Zr0.33O2 three-way catalyst before and after aging.

Appl. Catal. B: Environ. 2010, 97, 41–48. (45) Zhang, F.; Chen, J.; Chen, P.; Sun, Z.; Xu, S. Pd nanoparticles supported on hydrotalcite-modified porous alumina spheres as selective hydrogenation catalyst.

AIChE J. 2012, 58, 1853–1861. (46) Yang, L.; Lin, S.; Yang, X.; Fang, W.; Zhou, R. Promoting effect of alkaline earth metal doping on catalytic activity of HC and NOx conversion over Pd-only three-way catalyst. J. Hazard. Mater. 2014, 279, 226–235. (47) Zheng, X.; Chen, X.; Chen, J.; Zheng, Y.; Jiang, L. Synthesis and application of highly dispersed ordered mesoporous silicon-doped Pd-alumina catalyst with high thermal stability. Chem. Eng. J. 2016, 297, 148–157. (48) Valente, J.S.; Lima, E.; Toledo-Antonio, J.A.; Cortes-Jacome, M.A.; Lartundo-Rojas, L.; Montiel, R.; Prince, J. Influence of Mg/Al ratio on the thermokinetic rehydration of calcined Mg−Al layered double hydroxides. J. Phys.

Chem. C 2010, 114, 2089–2099. (49) Gusmano, G.; Montesperelli, G.; Traversa, E. Microstructure and electrical properties of MgAl2O4 thin films for humidity sensing. J. Am. Ceram. Soc. 1993, 76, 743–750. (50) Kim, D.H.; Woo, S.I.; Yang, O.B. Effect of pH in a sol–gel synthesis on the

28 ACS Paragon Plus Environment

Page 28 of 47

Page 29 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

physicochemical properties of Pd–alumina three-way catalyst. Appl. Catal. B:

Environ. 2000, 26, 285–289. (51) Kaneeda, M.; Iizuka, H.; Hiratsuka, T.; Shinotsuka, N.; Arai, M. Improvement of thermal stability of NO oxidation Pt/Al2O3 catalyst by addition of Pd. Appl. Catal.

B-Environ. 2009, 90, 564–569.

29 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table captions Table 1 Textural parameters of the samples prepared with different content of Mg calcined at 1000 °C.

Table 2 The full-conversion temperature (T100%) for CO, HC and NO of catalysts. Table 3 The amount of CO2 desorption from CO2-TPD profiles. Table 4 Metallic dispersion, metallic surface area and active particle diameter of catalysts calcined at different temperatures.

Table 5 The TOF values of catalysts calcined at different temperatures.

Figure captions Fig. 1 Small-angle XRD patterns of the xMg-OMA materials calcined at (A) 800 °C and (B) 1000 °C (a: 0Mg-OMA, b: 3Mg-OMA, c: 6Mg-OMA, d: 9Mg-OMA, e: 12Mg-OMA, f: 15Mg-OMA)

Fig. 2 Wide-angle XRD patterns of the xMg-OMA materials calcined at (A) 800 °C and (B) 1000 °C (a: 0Mg-OMA, b: 3Mg-OMA, c: 6Mg-OMA, d: 9Mg-OMA, e: 12Mg-OMA, f: 15Mg-OMA)

Fig. 3 Nitrogen adsorption-desorption isotherms (A) and pore size distributions (B) of the xMg-OMA materials with NH4NO3 calcined at 1000 °C (a: 0Mg-OMA, b: 3Mg-OMA, c: 6Mg-OMA, d: 9Mg-OMA, e: 12Mg-OMA, f: 15Mg-OMA)

Fig. 4 DTA profiles of the xMg-OMA materials with NH4NO3 calcined at 1000 °C (a: 0Mg-OMA, b: 3Mg-OMA, c: 6Mg-OMA, d: 9Mg-OMA, e: 12Mg-OMA, f: 15Mg-OMA)

Fig. 5 TEM images of 9Mg-OMA calcined at (A) 800 °C and (B) 1000 °C Fig. 6 Catalytic stability of the Pd/9Mg-OMA catalyst with NH4NO3 (under 280 °C) for CO, NO and HC elimination: (A) calcined at 800 ºC, (B) calcined at 1000 ºC.

30 ACS Paragon Plus Environment

Page 30 of 47

Page 31 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Fig. 7 CO2-TPD profiles of Pd/xMg-OMA with NH4NO3 calcined at 1000 °C (a: Pd/0Mg-OMA, b: Pd/3Mg-OMA, c: Pd/6Mg-OMA, d: Pd/9Mg-OMA, e: Pd/12Mg-OMA, f: Pd/15Mg-OMA)

Fig. 8 In situ DRIFTS spectra of chemisorbed CO2 on Pd/xMg-OMA with NH4NO3 calcined at 1000 °C after adsorption and evacuation at (A) 30 oC and (B) 180 oC. (a: Pd/0Mg-OMA, b: Pd/3Mg-OMA, c: Pd/6Mg-OMA, d: Pd/9Mg-OMA, e: Pd/12Mg-OMA, f: Pd/15Mg-OMA)

Fig. 9 In situ-DRIFTS spectras of chemisorbed CO on Pd/xMg-OMA with NH4NO3 calcined at 1000 °C: (A) absorbed CO for 1 h and (B) purged by helium. (a: Pd/0Mg-OMA, b: Pd/3Mg-OMA, c: Pd/6Mg-OMA, d: Pd/9Mg-OMA, e: Pd/12Mg-OMA, f: Pd/15Mg-OMA)

Fig. 10 H2-TPR profiles of Pd/xMg-OMA with NH4NO3 calcined at 1000 °C. (a: Pd/0Mg-OMA, b: Pd/3Mg-OMA, c: Pd/6Mg-OMA, d: Pd/9Mg-OMA, e: Pd/12Mg-OMA, f: Pd/15Mg-OMA)

Fig. 11 TEM images of Pd/9Mg-OMA with NH4NO3 calcined at (A) 800 °C and (B) 1000 °C

Fig. 12 The XPS profiles of samples calcined at 1000 °C. (a: 9Mg-OMA, b: Pd/9Mg-OMA, c: Pd/0Mg-OMA)

31 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 47

Table 1 Textural parameters of the samples prepared with different content of Mg calcined at 1000 °C. Without NH4HCO3 Samples

SBET/m2 g-1

With NH4HCO3

Vp/cm3 g-1 Dp/nm

SBET/m2 g-1

Vp/cm3 g-1

Dp/nm

Mg–OMA-0

162.5

0.48

11.7

183.6

0.52

12.1

Mg–OMA-3

184.2

0.53

10.5

197.3

0.58

11.3

Mg–OMA-6

192.1

0.51

9.4

211.7

0.66

10.9

Mg–OMA-9

212.3

0.56

9.8

234.4

0.72

10.5

Mg–OMA-12

189.6

0.49

9.1

207.8

0.63

10.8

Mg–OMA-15

178.9

0.42

10.6

191.5

0.57

11.3

Table 2 The full-conversion temperature (T100%) for CO, HC and NO of catalysts. 800 °C

1000 °C

Catalysts CO

HC

NO

CO

HC

NO

Pd/Mg-OMA-0

327.7

323.3

333.5

357.5

358.5

364.5

Pd/Mg-OMA-3

312.3

317.5

324.8

348.2

342.6

359.3

Pd/Mg-OMA-6

301.5

295.2

310.4

342.5

337.9

351.8

Pd/Mg-OMA-9

287.9

281.7

292.6

336.8

326.3

342.8

Pd/Mg-OMA-12

296.3

302.5

317.2

343.7

344.2

356.5

Pd/Mg-OMA-15

313.4

321.6

328.3

352.6

351.6

363.4

32 ACS Paragon Plus Environment

Page 33 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Table 3 The amount of CO2 desorption from CO2-TPD profiles. Amount of CO2 desorption ( unit area/g) Catalysts

1st peak

2nd peak

3rd peak

Total

Pd/Mg-OMA-0

1.02

0.34

4.52

5.88

Pd/Mg-OMA-3

1.14

0.39

5.78

7.31

Pd/Mg-OMA-6

1.25

0.45

6.56

8.26

Pd/Mg-OMA-9

1.37

0.63

7.87

9.87

Pd/Mg-OMA-12

1.19

0.46

7.21

8.86

Pd/Mg-OMA-15

1.07

0.36

4.89

6.32

Table 4 Metallic dispersion, metallic surface area and active particle diameter of catalysts calcined at different temperatures. 800 °C

1000 °C

Metal dispersion (%)

Metallic surface area (m2 g–1)

Active particle diameter (nm)

Pd/Mg-OMA-0

23.4

45.6

5.5

17.3

33.5

7.7

Pd/Mg-OMA-3

27.5

49.7

5.2

18.4

35.8

7.2

Pd/Mg-OMA-6

31.4

53.3

4.9

21.4

38.4

6.9

Pd/Mg-OMA-9

34.3

58.9

4.4

24.8

40.3

6.2

Pd/Mg-OMA-12

28.6

54.4

4.7

23.6

37.5

6.5

Pd/Mg-OMA-15

25.5

52.1

5.3

19.3

34.1

6.8

Catalysts

Metal Metallic Active dispersion surface particle (%) area diameter (m2 g–1) (nm)

33 ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 47

Table 5 The TOF values of catalysts calcined at different temperatures. 800 °C

1000 °C

Catalysts CO

HC

NO

CO

HC

NO

Pd/Mg-OMA-0

54.8

35.6

8.7

38.2

24.3

6.7

Pd/Mg-OMA-3

57.6

37.8

9.2

39.3

27.9

7.1

Pd/Mg-OMA-6

63.2

40.4

9.6

43.6

29.4

7.4

Pd/Mg-OMA-9

60.7

43.5

10.1

47.1

31.3

7.8

Pd/Mg-OMA-12

59.5

41.7

9.7

44.3

28.4

7.5

Pd/Mg-OMA-15

55.1

36.9

9.1

41.9

25.5

7.1

34 ACS Paragon Plus Environment

Page 35 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Fig. 1 Small-angle XRD patterns of the xMg-OMA materials calcined at (A) 800 °C and (B) 1000 °C (a: 0Mg-OMA, b: 3Mg-OMA, c: 6Mg-OMA, d: 9Mg-OMA, e: 12Mg-OMA, f: 15Mg-OMA) 66x31mm (300 x 300 DPI)

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 2 Wide-angle XRD patterns of the xMg-OMA materials calcined at (A) 800 °C and (B) 1000 °C (a: 0Mg-OMA, b: 3Mg-OMA, c: 6Mg-OMA, d: 9Mg-OMA, e: 12Mg-OMA, f: 15Mg-OMA) 67x32mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 36 of 47

Page 37 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Fig. 3 Nitrogen adsorption-desorption isotherms (A) and pore size distributions (B) of the xMg-OMA materials with NH4NO3 calcined at 1000 °C (a: 0Mg-OMA, b: 3Mg-OMA, c: 6Mg-OMA, d: 9Mg-OMA, e: 12Mg-OMA, f: 15Mg-OMA) 66x32mm (300 x 300 DPI)

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 4 DTA profiles of the xMg-OMA materials calcined at 1000 °C (a: 0Mg-OMA, b: 3Mg-OMA, c: 6Mg-OMA, d: 9Mg-OMA, e: 12Mg-OMA, f: 15Mg-OMA) 70x54mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 38 of 47

Page 39 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Fig. 5 TEM images of 9Mg-OMA calcined at (A) 800 °C and (B) 1000 °C 58x24mm (300 x 300 DPI)

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 6 Catalytic stability of the Pd/9Mg-OMA catalyst with NH4NO3 (under 280 °C) for CO, NO and HC elimination: (A) calcined at 800 ºC, (B) calcined at 1000 ºC. 67x32mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 40 of 47

Page 41 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Fig. 7 CO2-TPD profiles of Pd/xMg-OMA with NH4NO3 calcined at 1000 °C (a: Pd/0Mg-OMA, b: Pd/3Mg-OMA, c: Pd/6Mg-OMA, d: Pd/9Mg-OMA, e: Pd/12Mg-OMA, f: Pd/15Mg-OMA) 80x72mm (300 x 300 DPI)

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 8 In situ DRIFTS spectra of chemisorbed CO2 on Pd/xMg-OMA with NH4NO3 calcined at 1000 °C after adsorption and evacuation at (A) 30 oC and (B) 180 oC. (a: Pd/0Mg-OMA, b: Pd/3Mg-OMA, c: Pd/6Mg-OMA, d: Pd/9Mg-OMA, e: Pd/12Mg-OMA, f: Pd/15Mg-OMA) 65x30mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 42 of 47

Page 43 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Fig. 9 In situ-DRIFTS spectras of chemisorbed CO on Pd/xMg-OMA with NH4NO3 calcined at 1000 °C: (A) absorbed CO for 1 h and (B) purged by helium. (a: Pd/0Mg-OMA, b: Pd/3Mg-OMA, c: Pd/6Mg-OMA, d: Pd/9Mg-OMA, e: Pd/12Mg-OMA, f: Pd/15Mg-OMA)

67x32mm (300 x 300 DPI)

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 10 H2-TPR profiles of Pd/xMg-OMA with NH4NO3 calcined at 1000 °C. (a: Pd/0Mg-OMA, b: Pd/3Mg-OMA, c: Pd/6Mg-OMA, d: Pd/9Mg-OMA, e: Pd/12Mg-OMA, f: Pd/15Mg-OMA)

85x81mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 44 of 47

Page 45 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Fig. 11 TEM images of Pd/9Mg-OMA with NH4NO3 calcined at (A) 800 °C and (B) 1000 °C 56x22mm (300 x 300 DPI)

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 12 The XPS profiles of samples calcined at 1000 °C. (a: 9Mg-OMA, b: Pd/9Mg-OMA, c: Pd/0Mg-OMA) 143x146mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 46 of 47

Page 47 of 47

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

For Table of Contents Only

ACS Paragon Plus Environment