Supporting Information for: High Temperature Structural Stability of Ceria Based Inverse Opals Danielle C. Casillas,‡ Dan C. Wilkinson,‡ Chun-Han Lai,‡ Michael Ignatowich,‡‡ Stephen K. Wilke,ζ Sossina M. Haile,§,¶,ζ,‡‡ and Bruce S. Dunn‡,† ‡ Department of Materials Science and Engineering, University of California, Los Angeles, Los Angeles, California 90095 § Materials Science and Engineering, Northwestern University, Evanston, Illinois 60208 ¶ Applied Physics, Northwestern University, Evanston, Illionis 60208 ζ ‡‡ Materials Science, California Institute of Technology, Pasadena, California 91125 Chemical Engineering, California Institute of Technology, Pasadena, California 91125 FFT Image Analysis: The fast Fourier transform (FFT) is described by equation 1: 𝑁1 −1 𝑁2 −1 𝑋(𝑘𝑥 , 𝑘𝑦 ) = ∑ ∑ 𝑥(𝑛1 , 𝑛2 𝑘 𝑛 𝑘𝑦 𝑛2 ) −2𝜋𝑖( 𝑥 1 + 𝑁1 𝑁2 )𝑒 (1) 𝑛1 =0 𝑛2 =0 where 𝑥 (𝑛1 , 𝑛2 ) is the array of brightness values and 𝑘/𝑁 is the spatial frequency. In order to compute the FFT, the fft2(X) function in Matlab was implemented. The image is converted to grayscale, a threshold applied, and cropped to remove the information bar. The one-dimensional discrete Fourier transform (DFT) is computed first for each column, and subsequently computed for the row of each result. This results in a circular convolution, and the FFT data has the same dimensions as the original image. Next, Y=fftshift(x) is used to move the zero frequency component to the center, and the absolute value is taken in order to remove the imaginary values from the analysis. The absolute value is the intensity value seen in the final FFT image: 𝐼(𝑘𝑥 , 𝑘𝑦 ) = |𝑋(𝑘𝑥 𝑘𝑦 )| (2) If the image contains spatial order, this will result in an array of spots similar to a diffraction pattern. The pixels comprising the spot at the center of the FFT image represent the average brightness of the image. Additionally, the nearest neighbor spots to the zeroth frequency value represent the lowest frequency occurring order. For this study, these are the data of interest. The sum of the FFT along an axis of symmetry is taken by rotating the image and summing the columns of the resulting FFT (Fig. S1). (a) (b) (c) (d) Fig. S1 (a) Typical ZSC20 1µm inverse opal SEM image with threshold applied, (b) FFT image before rotation, (c) FFT image after rotation, (d) sum of FFT image brightness along x-axis (matrix columns). The spatial order parameter extracted from these FFT analyses is defined as the ratio of the normalized first order FFT sum peak and the full width at half maximum (FWHM) of the peak normalized by the total width of the image in pixels.24 In order to calculate the raw spatial order parameter, γ, the peak intensity and full width at half maximum (FWHM) must be extracted. The findpeaks function in Matlab was used, and the standard deviation of each peak was taken to estimate the FWHM. The peak minimum was subtracted from the peak maximum to find the height. These results correlate well with those of the FFT analysis. Here, γr > 0.4 indicates complete retention of ordered porosity, while 0.2 < γr < 0.4 denotes a gradual loss of ordered porosity, and materials with γr < 0.2 have random porosity. ZSC20 inverse opals retain ordered porosity at 1000 °C for 12 hours, while ZSC30 and ZSC40 inverse opals maintain their structures at 1100 °C for 12 hours. X-Ray Diffraction Powder XRD patterns for ZSC20, 30, and 40 after annealing at 1100 °C for 12 hours (Fig. S2 (a)) show peak splitting, indicating formation of a second phase with smaller lattice parameter for zirconia content greater than 20 at%.46 The shift of the (100) reflection at 29˚to lower 2θ and the appearance of a shoulder at 30˚ point to a segregation of material into zirconium-rich tetragonal, and zirconium-deficit cubic oxides. For Zr substitution greater than 20%, the tetragonal phase becomes apparent upon annealing. Tetragonal phase formation is suppressed (or perhaps undetectable due to peak broadening) in nano-sized grains. However, as shown here, upon annealing at high temperatures significant grain growth occurs. Powder X-ray diffraction (XRD) also shows the emergence of a tetragonal, zirconium-rich phase. In Fig. S2, the XRD patterns ZSC30 and ZSC40 annealed at 1100 ˚C contain extra reflections indicative of the tetragonal phase. Consequently, for extended use of these materials at high temperatures, Zr additions should remain around 20 atomic percent in order to achieve a highly defective lattice without tetragonal phase formation. In Fig. S2(b), the dependence of γ on crystallite size is shown for as-prepared samples with nominal pore size of 1 m, and for identical materials annealed at 1100 ˚C for 12 hours. All compositions in the as-synthesized state display an initial γ of 10-11 and an initial crystallite size of 5-15 nm, and the two parameters are relatively uncorrelated. The post-annealed materials show a stronger correlation between these two parameters, indicating that the processes leading to grain growth also result in loss of pore ordering. The retention of order and small grain size with increasing Zr content suggests that the mixed cation compositions have lower mean cation mobility. (b) (a) 2θ (degree) Fig. S2 (a) Powder XRD patterns for ZSC20, ZSC30, and ZSC40 inverse opals as-synthesized and annealed at 1100°C for 12 hours. Lines indicate peak locations for tetragonal Ce0.86Zr0.14O2 reference JCPDS 00-038-1437). (b) γ as a function of zirconium dependent grain size for ZSC as-synthesized (closed symbols) and annealed at 1100˚C (open symbols). Crystallite size for ZSC10 from reference 35. Raman Spectroscopy Raman spectroscopy (Renishaw, 514nm) was used to verify subtle changes in the oxygen sublattice such as vacancy formation and pseudo-cubic (t”) phase emergence. Fig. S3 shows the Raman spectra of ZSC20 inverse opals annealed at 1100°C for 0.5 and 12 hours. The main cubic phase vibrational mode peak occurs at 464 cm-1 for pure ceria, whereas the defect band (oxygen vacancy) is at 600 cm-1, and a band at 307 cm-1 is characteristic of a tetragonal distortion. Peak positions are designated by 2, 3 and 1 in Fig. S3, respectively.39 The cubic peak position is slightly shifted due to the addition of zirconia, and shift of the cubic peak to higher wavenumbers with annealing occurs due to crystallite growth. This phenomenon is likely due to phonon confinement in small nanoparticles, in addition to lattice expansion, and these effects disappear as the nanocrystals grow.47,48 The defect band is indicative of oxygen vacancy concentration, and is associated with the pseudocubic phase. In this phase, cation positions remain unchanged and oxygen positions are slightly shifted rom their locations in the cubic lattice. Fig. S3 Raman spectra of ZSC20 1µm inverse opal (a) assynthesized, and annealed at 1100 °C for (b) 0.5 hours and (c) 12 hours. Pseudocubic phase, cubic phase, and oxygen vacancy vibrational modes locations are indicated by 1, 2, and 3, respectively. None of the ZSC samples annealed at 1000 °C showed evidence of tetragonal t phase formation, which would be indicated by peaks in addition to 1, 2 and 3, defined in Fig. S3. The Raman spectra of ZSC30 and ZSC40 did, however, contain small broad peaks at 307 cm-1 and 600 cm-1, indicating the emergence of the defective pseudocubic t” phase (Fig. S4). ZSCX Inverse opals annealed at 1100 °C for 1 hour are shown to contain defective pseudocubic phase when X>10. However, undesired tetragonal phases (t and t’) begin to emerge where X>20, as indicated by the presence of additional peaks ~260 cm-1. The amount of Zr is the key parameter which determines the relation between coarsening and the ability of inverse opals to maintain their structures. Large grains allow facile crack propagation along grain boundaries, and thus make inverse opals more susceptible to fracture and loss of order, in addition to the coarsening effects of long term high temperature exposure.49,50 Fig. S4 Raman spectra of ZSC inverse opals annealed at 1000 °C and 1100 °C for 1 hour with (a) 0, (b) 10, (c) 20, (d) 30, (e) 40 atomic percent zirconia. Pore Size Estimation for Facile Gas Transport An effort towards the fabrication of large (> 1µm) pore ceria based inverse opals was motivated by potential mass transport improvements. The small pore radii typical of catalyst materials dictates the fluid flow regime. Within the molecular flow regime, the mean free path of the fluid is much less than the diameter of the pore, and particle-particle interactions dominate flow. In contrast, within the Knudsen flow regime the mean free path of the gas is greater than the pore diameter, and particle-pore wall interactions dominate. For gases, typical diffusion coefficients within the molecular regime and Knudsen regime are 10-1 cm2/s, and 10-5 to 10-2 cm2/s, respectively.51,52 To determine the diffusion regime within the pores it is useful to compare the mean free path of the gas molecule to the pore diameter, which results in Knudsen number:51 𝐾𝑛# = 𝜆 𝑘𝐵 𝑇 = 𝑑 √2𝜋𝜎 2 𝑝𝑑 (3) where d is the pore diameter, λ is the mean free path, kB is the Boltzmann constant, T is the temperature, σ is the hard sphere diameter of a gas molecule, and p is the pressure. When Knudsen number is much smaller than unity, molecular flow is dominant.52 The hard sphere radius can be estimated using the second coefficient of the virial expansion, which simplifies to the second parameter of the Van der Waals equation of state (Eq. 4). This parameter, the Van der Waals volume, takes into account the volume exclusion of gas molecules.53 For a rough approximation, the hard sphere radius is determined from the Van der Waals volume of the gas using the relation: 4 𝜎 3 𝑏 = 𝑁𝑎 ∗ 𝜋 ( ) 3 2 (4) Where b is the Van der Waals volume [m3/mol] and Na is Avogadro’s number. To estimate the minimum pore radius which ensures flow within the molecular regime, the Knudsen number is set to 0.5 at atmospheric pressure, and the pore window diameter can be determined. From visual inspection of pores, the window diameter is ~1/3 of the pore diameter. Table S1 outlines the minimum pore diameters for the gases in this system using tabulated b values.53 A minimum pore diameter of ~4.5µm should guarantee molecular flow. Table S1 Estimated minimum pore diameters for flow within the molecular regime. Gas Min. Pore Diameter (μm) b/10-5 3 (m /mol) 800°C Water 3.1 4.2 Oxygen 3.2 4.2 Hydrogen 2.7 4.8 BET Analysis Brunauer-Emmett-Teller (BET) surface area and pore size distribution analysis were carried out at 77 K using a gas adsorption analyzer (Micromeritics ASAP 2010) with N2 adsorption isotherms. Before analysis, 200 mg test samples were outgassed at 110 °C under vacuum overnight to ensure complete drying. The nitrogen adsorption isotherms were taken from a pressure range between 5 μm Hg and 960 μm Hg. Table S2 shows that values of BET surface area typically fall between ~ 25 m2 g-1 and 50 m2 g-1 for different ZSC20 inverse opals and powders. Table S2 BET surface area for ZSC20 powders and inverse opals of different pore sizes. Inverse opal samples Surface area (m2 g-1) Control powders 300 nm 650 nm 1 μm 51.8 ± 3.7 36.2 ± 8.8 31.3 ± 1.2 25.1 ± 1.4 Hydrogen Production Rate (mL/min/g) Hydrogen production by ZDC20: 4.0 3.5 3.0 300 nm 650 nm 1000 nm Control Powder 2.5 2.0 1.5 1.0 0.5 0.0 0 1 2 3 4 5 6 7 8 9 10 Time (min) Fig. S5: Hydrogen production profiles for different ZDC20 inverse opal microstructures and control powder. Chemically reduced ZSC20 inverse opals having 300 nm, 650 nm and 1 µm nominal pore sizes were evaluated for hydrogen production efficacy as follows. Samples 400 mg in mass were placed in a horizontal tube furnace and heated at rate of 10 ˚C/min in air to 800 ˚C. After purging with Ar gas, the samples were reduced by exposure to a mixture of 3% H2 and 20% H2O in Ar for 20 minutes at a flow rate of 200 sccm (standard cubic centimeters per minute, implying a gas velocity at the sample of 18 cm/s and reaction zone flush time of less than 1 s). The reactor was then purged with Ar for 5 minutes (1000 sccm and gas velocity of 92 cm/s) to remove residual hydrogen, following which, the inverse opals were oxidized using a wet Ar stream for 20 minutes (20% H2O, 200 sccm). The reactor was then purged again with Ar for 5 minutes before beginning another cycle. The process was repeated 12 times. Representative hydrogen production profiles obtained during the inverse opal oxidation step are presented in Figure S5. The integrated hydrogen production (area under the curves) is similar between structures, an expected result because the total production reflects the thermodynamic properties of the material. The minimal influence of pore size on the features of the hydrogen evolution profiles is tentatively assigned to competing effects of (slightly) decreasing surface area and presumably increasing gas diffusion with increasing pore size. References 46. H. Jen, G. Graham, W. Chun, R. McCabe, J. Cuif, S. Deutsch, and O. Touret, “Characterization of Model Automotive Exhaust Catalysts: Pd on Ceria and Ceria–Zirconia Supports,” Catal. Today, 50, 309-328 (1999). 47. J. E. Spanier, R. D. Robinson, F. Zhang, S. –W. Chan, and I. P. Herman, “Sizedependent Properties of CeO2-y Nanoparticles as Studied by Raman Scattering,” Phys. Rev. B, 64, 245407 (2001). 48. R. Si, Y. Zhang, S. Li, B. Lin, and C. Yan, “Urea-Based Hydrothermally Derived Homogeneous Nanostructured (Ce1-xZrx)O2 (x=0-0.8) Solid Solutions: A Strong Correlation between Oxygen Storage Capacity and Lattice Strain,“ J. Phys. Chem. B., 108, 12481-12488 (2004). 49. X. You, T. Si, N. Liu, P. Ren, P. Xu, and J. Feng, ”Effect of Grain Size on Thermal Shock Resistance of Al2O3-TiC Ceramics,” Ceram. Int. 31, 33-38 (2005). 50. F. Knudsen, “Dependence of Mechanical Strength of Brittle Polycrystalline Specimens on Porosity and Grain Size,” J. Am. Ceram. Soc. 42, 376-387 (1959). 51. M. Davis and R. Davis, Fundamentals of Chemical Reaction Engineering, 4th ed., McGraw Hill, New York, 2003. 52. H. Fogler, Elements of Chemical Reaction Engineering, 5th Ed.; pp. 757-801, Prentice Hall, Inc., New Jersey, 2006. 53. R. Weast, Handbook of Chemistry and Physics. 53rd Ed., Chemical Rubber Co., Cleveland, 1972.
© Copyright 2025 Paperzz