Atomically dispersed Cu on Ce1−xRExO2−δ nanocubes (RE = La and Pr) for water gas shift: influence of OSC on catalysis

Muhammad Ridwanac, Rizcky Tamaranya, Jonghee Hanab, Suk Woo Namab, Hyung Chul Hamac, Jin Young Kima, Sun Hee Choia, Seong Cheol Janga and Chang Won Yoon*ac
aFuel Cell Research Center, Korea Institute of Science and Technology, Seoul 02792, Republic of Korea. E-mail: cwyoon@kist.re.kr; cw.yoon@ust.ac.kr
bKorea University, Seoul 02841, Republic of Korea
cClean Energy and Chemical Engineering, Korea University of Science and Technology, Daejeon 34113, Republic of Korea

Received 5th October 2015 , Accepted 15th October 2015

First published on 15th October 2015


Abstract

To elucidate the effect of CeO2 shape and doping on activity, Cu0.02Ce0.98O2−δ and Cu0.02Ce0.86RE0.12O2−δ (RE = La and Pr) were synthesized by a molecular precursor approach. The materials showed distinct activities depending on the shape and composition of CeO2, which was well correlated with their different oxygen storage capacities.


Owing to its intriguing properties, including high oxygen storage capacity (OSC) and oxygen mobility, ceria (CeO2) has been extensively studied as a metal-oxide support for numerous catalysis reactions.1–4 In CeO2-supported catalysts, the OSC of CeO2 plays a crucial role in catalyzing industrially important reactions, particularly those associated with oxygen ion (O2−) transport, such as water gas shift (WGS) and partial oxidation (POX) reactions.1–4 Significant efforts have been made to improve the OSC to facilitate a number of catalytic reactions. One strategy to enhance the OSC of CeO2 is through morphology control. CeO2 nanocubes with (100) facets have increased oxygen vacancies compared with CeO2 irregular nanoparticles (NPs) with (111) facets.5 Alternatively, the incorporation of rare earth elements as dopants into the CeO2 lattice can increase the OSC.1–4 The doping of CeO2 with rare earth metal elements such as Pr, La, and Tb increases the OSC, preventing significant carbon coking and ultimately enhancing the catalyst durability in the steam methane reforming reaction.6 In addition, atomically dispersed metals on oxide supports, Au/M-doped CeO2 (M = La), were recently reported,7,8 demonstrating the importance of active site distribution for the WGS reaction and providing an insight into the design of next generation catalysts with highly dispersed active sites.

Herein, we report atomically distributed Cu catalysts supported on La or Pr-doped CeO2 nanocubes for the WGS reaction. These catalysts were prepared by a molecular precursor method using a Cu-containing organometallic compound. This new synthetic approach for CuCeREO2−δ (RE = La or Pr) in conjunction with shape control and doping strategies provided the uniformly dispersed active sites with increased OSC needed to achieve high activity for WGS.

To improve the OSC of CeO2, we employed the following synthetic approaches: (i) morphology control and (ii) doping with rare earth elements. Method (i) yielded CeO2 irregular nanoparticles (NPs) (1) and CeO2 nanocubes (2), whereas method (ii) gave Ce0.88La0.12O2−δ (3) and Ce0.88Pr0.12O2−δ (4) nanocubes (vide infra). To analyze the shape of the prepared CeO2 supports, morphological studies of 1 and 2 were conducted by HR-TEM. 1, prepared by the co-precipitation method, showed irregular shapes with d-spacing of 0.315 nm (Fig. 1a), indicating the formation of CeO2 (111) planes at the surface.9 In contrast, 2, obtained by the hydrothermal method, exhibited well-defined cubic structures with uniform sizes of 20–40 nm. The HR-TEM image of 2 shows apparent (100) lattice fringes with an interplanar spacing of 0.272 nm (Fig. 1b), suggesting that CeO2 (100) fringes are dominant at the surface of 2.9 Structure determination for 3 and 4 by TEM indicated that doping of La or Pr into 2 did not affect the shape (Fig. 1c and d).


image file: c5ra20557c-f1.tif
Fig. 1 HR-TEM images: (a) CeO2 irregular NPs (1), (b) CeO2 nanocubes (2), (c) Ce0.88La0.12O2−δ nanocubes (3), and (d) Ce0.88Pr0.12O2−δ nanocubes (4).

Materials 1–4 likely have different OSCs depending on the morphology or dopant. The relative quantities for oxygen vacancies of 1–4 were examined using Raman spectroscopy (Fig. 2); the sharp peaks with maxima centered at ca. 465 cm−1 were attributed to the F2g mode, characteristic of the CeO2 fluorite crystal structure.1 Notably, the F2g peaks for 3 and 4 broadened slightly and shifted toward lower frequencies, indicating the formation of solid solutions upon La or Pr incorporation.1 In addition to the F2g mode, 4 had a dominant peak centered at ca. 570 cm−1 owing to oxygen vacancies generated by doping with a trivalent cation; replacement of Ce4+ in the CeO2 lattice with Pr3+ produces an oxygen vacancy to maintain charge neutrality.3 Likewise, the Raman spectrum of 3 had two additional modes at 540 and 600 cm−1 associated with the local vibrations of different oxygen vacancy (VO) complexes.1 Particularly, the vibrational mode at 600 cm−1 originates from the presence of Ce3+–VO complexes in the CeO2 lattice, referred to as the intrinsic vacancy mode, whereas that of ∼540 cm−1 comes from extrinsic vacancy mode related to La doping.10 In the Pr-doped samples the extrinsic vacancy mode appears at ∼570 cm−1. A higher I575/I465 ratio indicates a higher quantity of oxygen vacancies.11 These results suggest that among these materials, the ability for the formation of oxygen vacancies for 4 is superior.


image file: c5ra20557c-f2.tif
Fig. 2 Raman spectra: (a) CeO2 irregular NPs (1), (b) CeO2 nanocubes (2), (c) Ce0.88La0.12O2−δ nanocubes (3), and (d) Ce0.88Pr0.12O2−δ nanocubes (4). Inset presents Raman spectra of 1, 2 and 3 with the range of 500–650 cm−1.

We also employed Ce 3d X-ray photoelectron spectroscopy (XPS) to determine the relative capability for oxygen vacancy formation of 1–4 by assessing their Ce3+/[Ce3+ + Ce4+] ratios. The XPS spectra had seven peaks corresponding to Ce 3d3/2 (897–915 eV) and Ce 3d5/2 (875–897 eV)12 (u and v, respectively, Fig. S1). The relative Ce3+ 3d5/2 (v′)/[Ce3+ 3d5/2 (v′) + Ce4+ 3d5/2 (v, v′′, and v′′′)] ratios for 1–4, determined from the integrated areas of the fitted data (Fig. S1 and Table S1), provide useful information about the oxygen vacancies. 1 and 2, which have different shapes, had Ce3+ fractions of 15% and 27%, respectively, indicating that 2 has increased oxygen ion mobility compared with 1. In contrast, the incorporation of La or Pr into 2 to obtain 3 or 4 did not alter the Ce3+/[Ce3+ + Ce4+] ratio, consistent with a previous result.13

To gain further information about OSC of the materials, we employed thermogravimetric analyses (Fig. S2), according to a previously reported method.5 Since WGS reactions were conducted at <400 °C (vide infra), we determined the quantities of oxygen vacancies using the heating temperature of 400 °C. The catalysts were initially heated from room temperature to 400 °C with air flow. In this process, oxygen vacancies presented in a catalyst were filled by O2. At 400 °C, N2 gas was then supplied for 10 min to abstract the filled oxygen from the catalyst, which resulted in weight loss. Repeated the processes gave the quantities of oxygen vacancies in the catalyst (Table 1). Consistent with the Raman results, the measured OSCs were found to increase in the order of 4 > 3 > 2 > 1.

Table 1 Calculated oxygen vacancies for 1–4 based on thermogravimetric analysis at 400 °C
Samples Weight loss (%) μmol of O2a per gcat
a The quantities of oxygen vacancy = g of decreased weight/molecular weight of O2.5
CeO2 irregular NPs (1) 0.034 10
CeO2 nanocubes (2) 0.052 16
Ce0.88La0.12O2−δ nanocubes (3) 0.072 22
Ce0.88Pr0.12O2−δ nanocubes (4) 0.074 23


The enhanced OSC of the modified CeO2 materials could provide catalysts with improved activity. We incorporated Cu into the crystal structure of CeO2 to generate catalytically active sites for the WGS. Precipitation, wet impregnation, or co-precipitation deposition methods have widely been employed to introduce dopants into metal oxide lattices, but metal aggregates are often formed. Recently, atomically dispersed active sites supported on metal oxides were shown to be highly active for CO oxidation and WGS;7,14 the WGS reaction was accelerated by uniformly dispersed, nonmetallic Au or Pt species strongly associated with CeO2 that generated by removing excess Au or Pt in nanostructured Au- or Pt-CeOx materials.7

We employed a different synthetic strategy using an organometallic precursor, Cu(OCH3)2, to dope Cu atoms into CeO2 with increased distribution (Fig. 3). The surface hydroxyl (–OH) groups functioned as nucleophiles to react with Cu(OCH3)2 to yield Cu-anchored CeO2 and CH3OH as a byproduct (Fig. 3, reactions A and B). A reaction of 2 (0.20 g) with Cu(OCH3)2 (2.9 mg, 0.023 mmol) at 70 °C for 24 h clearly indicated the formation of CH3OH, as evidenced by 1H NMR spectroscopy (Fig. S3). Note that Cu agglomeration was likely minimized by the steric hindrance of the organometallic precursor. The well-dispersed Cu atoms were then incorporated into the CeO2 lattice by calcination at 400 °C (Fig. 3, reaction C). In fact, similar surface grafting methods for CeO2 with organosilanes were employed to improve chemical mechanical polishing.15,16 The as-synthesized catalysts are denoted as Cu0.02Ce0.98O2−δ irregular NPs (Cu-1), Cu0.02Ce0.98O2−δ nanocubes (Cu-2), Cu0.02Ce0.86La0.12O2−δ nanocubes (Cu-3), and Cu0.02Ce0.86Pr0.12O2−δ nanocubes (Cu-4). For comparison, CuCeO2 was prepared by a conventional process using Cu(NO3)2, followed by NaBH4 reduction: Cu0.02Ce0.98O2−δ irregular NPs (Cu-5) and Cu0.02Ce0.98O2−δ nanocubes (Cu-6).


image file: c5ra20557c-f3.tif
Fig. 3 A schematic diagram for the generation of uniformly distributed Cu active sites based on a molecular precursor approach (yellow: Ce atom, red: O atom, and blue: Cu atom).

The CuCeO2-based catalysts were characterized using BET and SEM-EDS (Table 2). SEM-EDS indicated that 0.58–0.74 wt% of Cu was deposited on the catalysts. STEM-EDS mapping indicated the presence of well-dispersed Cu in Cu-2 (as an example, Fig. S4). X-ray diffraction analyses of Cu-1–Cu-4 showed only the characteristic peaks of the CeO2 fluorite cubic structure (Fig. S5), indicating that the Cu atoms were well dispersed and formed a Cu-doped CeO2 solid solution. The low valence state of Cu2+, with its smaller ionic radius, was reported to allow facile substitution of Ce4+ in the CeO2 lattice with simultaneous oxygen vacancy generation to produce a CuxCe1−xO2−δ solid solution.4

Table 2 Characterization of the prepared CuCeO2 catalysts
Catalyst Cu precursor Method Metal loadingb (wt%) BET surface area (m2 g−1)
a Cu reduced using NaBH4.b Measured by SEM-EDS.
Cu-1 Cu(OCH3)2 Molecular precursor approach 0.70 56
Cu-2 Cu(OCH3)2 0.58 28
Cu-3 Cu(OCH3)2 0.62 30
Cu-4 Cu(OCH3)2 0.61 29
Cu-5 Cu(NO3)2·3H2Oa Conventional co-reduction 0.70 56
Cu-6 Cu(NO3)2·3H2Oa 0.74 27


We determined the catalytic activities of Cu-1–Cu-6 for the WGS reactions. First, we verified the influence of CeO2 facets on the WGS (Fig. 4). Cu-2 is expected to show higher activity than Cu-1 since the CeO2 nanocubes have a higher OSC than the CeO2 irregular NPs owing to the high surface energy of the (100) facet, which originates from the instability of the top layer oxygen atoms that bridge two cerium atoms.17 Recently, the OSC of CeO2 was determined using thermogravimetric analysis and, although nanocubes have smaller surface areas than irregular NPs, the OSC of CeO2 nanocubes was nearly 2.6 times higher than that of CeO2 irregular NPs at 400 °C.18 Similarly, the CO conversion using Cu-2 (42%) was greater than that of Cu-1 (30%) at 300 °C (Fig. 4a). Next, we examined the influence of doping by comparing Cu-2, Cu-3, and Cu-4. Compared with Cu-2, Cu-3 and Cu-4 had increased activities, presumably owing to the enhanced OSC caused by doping CeO2 with rare earth elements;1–4 the CO conversion of Cu-4 (74%) was 1.8 times higher than that of Cu-2 at 300 °C. Given the influence of both facet control and doping, Cu-4 has significantly increased activity (2.5 times higher than that of Cu-1 at 300 °C); this result correlates well with the relative OSCs of 1–4. In contrast, Cu-5, with dominant Cu (111) facets, had a higher catalytic activity than the Cu-6 nanocubes, with CO conversions of 32% and 16% at 300 °C, respectively (Fig. 4b). The CO conversion appears to be linear but it is expected to have a sigmoidal shape in an expanded temperature range. An Arrhenius plot obtained using the temperature dependent kinetic data for Cu-4 gave an activation energy of 22 kJ mol−1, which is considerably lower than that of Cu-1 (Fig. S6 and Table S2). We further conducted DFT calculations to determine the influence of the dopants on oxygen vacancy formation at the materials. Our calculations strongly support that the Pr-doped CeO2 likely has the highest oxygen vacancy mobility (Fig. S7 and S8).

 
CO + Cu-Ce-(O)-RE → CO2 + Cu-Ce-( )-RE (1)
 
H2O + Cu-Ce-( )-RE → H2 + Cu-Ce-(O)-RE (2)
 
Overall: CO + H2O → CO2 + H2 (3)


image file: c5ra20557c-f4.tif
Fig. 4 CO conversions during the WGS over the catalysts: (a) Cu-1 (image file: c5ra20557c-u1.tif, pink), Cu-2 (image file: c5ra20557c-u2.tif, black), Cu-3 (image file: c5ra20557c-u3.tif, red), and Cu-4 (image file: c5ra20557c-u4.tif, blue); (b) Cu-5 (image file: c5ra20557c-u5.tif, purple) and Cu-6 (image file: c5ra20557c-u6.tif, green).

Based on the experimental results, a plausible mechanism involves oxygen transfer from CuCeREO2. First, CO reacts with oxygen atoms in the CeO2 surface lattice to generate an oxygen vacancy (eqn (1)). The resulting lattice abstracts an oxygen atom from H2O to produce hydrogen (eqn (2)). The process is highly dependent on the OSC of the catalyst.

In summary, CuCeREO2−δ (RE = La or Pr) catalysts were prepared using the molecular precursor approach to generate Cu active sites well dispersed on CeO2-based supports. These catalysts were superior to those synthesized by the conventional co-reduction method. The catalytic activities for the WGS were enhanced owing to the increased OSC of CeO2; the catalyst with exposed (100) facets had better activity than that with exposed (111) facets. In addition, the doped materials with CeO2 (100) facets had even higher catalytic performance than the undoped CuCeO2 material. Our synthetic approach using an organometallic precursor provides insights into the design of new types of catalysts applicable to numerous chemical transformations.

Acknowledgements

The authors thank the Global Research Laboratory (GRL) Program through the National Research Foundation of Korea funded by the Ministry of Science, ICT and Future Planning of Republic of Korea. The part of the research was also supported by the Fundamental Technology Development Programs for the Future through the Korea Institute of Science and Technology as well as by the National Research Foundation of Korea Grant funded by the Korean Government (MSIP) (University-Institute cooperation program). The computational resources were supported by the supercomputing center of Korea Institute of Science and Technology Information (KISTI) (KSC-2015-C2-002).

References

  1. J. R. McBride, K. C. Hass, B. D. Poindexter and W. H. Weber, J. Appl. Phys., 1994, 76, 2435–2441 CrossRef CAS PubMed.
  2. D. Harshini, Y. Kim, S. Nam, T.-H. Lim, S.-A. Hong and C. Yoon, Catal. Lett., 2013, 143, 49–57 CrossRef CAS.
  3. M. Guo, J. Lu, Y. Wu, Y. Wang and M. Luo, Langmuir, 2011, 27, 3872–3877 CrossRef CAS PubMed.
  4. D. Zhang, Y. Qian, L. Shi, H. Mai, R. Gao, J. Zhang, W. Yu and W. Cao, Catal. Commun., 2012, 26, 164–168 CrossRef CAS PubMed.
  5. J. Zhang, H. Kumagai, K. Yamamura, S. Ohara, S. Takami, A. Morikawa, H. Shinjoh, K. Kaneko, T. Adschiri and A. Suda, Nano Lett., 2011, 11, 361–364 CrossRef CAS PubMed.
  6. D. Harshini, D. H. Lee, J. Jeong, Y. Kim, S. W. Nam, H. C. Ham, J. H. Han, T.-H. Lim and C. W. Yoon, Appl. Catal., B, 2014, 148–149, 415–423 CrossRef CAS PubMed.
  7. Q. Fu, H. Saltsburg and M. Flytzani-Stephanopoulos, Science, 2003, 301, 935–938 CrossRef CAS PubMed.
  8. M. Yang, L. F. Allard and M. Flytzani-Stephanopoulos, J. Am. Chem. Soc., 2013, 135, 3768–3771 CrossRef CAS PubMed.
  9. S. Bernal, G. Blanco, J. J. Calvino, C. López-Cartes, J. A. Pérez-Omil, J. M. Gatica, O. Stephan and C. Colliex, Catal. Lett., 2001, 76, 131–137 CrossRef CAS.
  10. Z. D. Dohčević-Mitrović, M. Grujić-Brojčin, M. Šćepanović, Z. V. Popović, S. Bošković, B. Matović, M. Zinkevich and F. Aldinger, J. Phys.: Condens. Matter, 2006, 18, S2061 CrossRef.
  11. M.-F. Luo, Z.-L. Yan, L.-Y. Jin and M. He, J. Phys. Chem. B, 2006, 110, 13068–13071 CrossRef CAS PubMed.
  12. M. A. Henderson, C. L. Perkins, M. H. Engelhard, S. Thevuthasan and C. H. F. Peden, Surf. Sci., 2003, 526, 1–18 CrossRef CAS.
  13. N. Paunovic, Z. Dohcevic-Mitrovic, R. Scurtu, S. Askrabic, M. Prekajski, B. Matovic and Z. V. Popovic, Nanoscale, 2012, 4, 5469–5476 RSC.
  14. B. Qiao, A. Wang, X. Yang, L. F. Allard, Z. Jiang, Y. Cui, J. Liu, J. Li and T. Zhang, Nat. Chem., 2011, 3, 634–641 CrossRef CAS PubMed.
  15. Z. Zhang, L. Yu, W. Liu and Z. Song, Appl. Surf. Sci., 2010, 256, 3856–3861 CrossRef CAS PubMed.
  16. K. Moller and T. Bein, Chem. Mater., 1998, 10, 2950–2963 CrossRef CAS.
  17. H. Nörenberg and J. H. Harding, Surf. Sci., 2001, 477, 17–24 CrossRef.
  18. T. Adschiri, Y. Hakuta, K. Sue and K. Arai, J. Nanopart. Res., 2001, 3, 227–235 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Materials and detailed preparation procedures, catalyst characterization, catalytic reaction, analysis of products, and DFT calculations. See DOI: 10.1039/c5ra20557c

This journal is © The Royal Society of Chemistry 2015