Next Article in Journal
Valorization of Reground Pasta By-Product through PHA Production with Phototrophic Purple Bacteria
Previous Article in Journal
A Numerical Case Study of Particle Flow and Solar Radiation Transfer in a Compound Parabolic Concentrator (CPC) Photocatalytic Reactor for Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Improved Structural Description of Different γ-Al2O3 Materials Using Disordered δ5-Al2O3 Phase via X-ray Pair Distribution Function Analysis

1
Institute of Crystallography, RWTH Aachen University, 52066 Aachen, Germany
2
Jülich Centre for Neutron Science (JCNS-3), Forschungszentrum Jülich GmbH, 52425 Jülich, Germany
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(4), 238; https://doi.org/10.3390/catal14040238
Submission received: 24 February 2024 / Revised: 29 March 2024 / Accepted: 31 March 2024 / Published: 3 April 2024
(This article belongs to the Section Nanostructured Catalysts)

Abstract

:
Extensive research has been conducted in the past on the crystallographic characteristics of γ-Al2O3 support materials due to their advantageous properties in heterogeneous catalysis. While their structure is most commonly described as spinel, their intrinsic disorder and nanostructure have prompted alternative models involving tetragonal space groups, supercells, or occupancy of non-spinel positions. X-ray pair distribution function (PDF) analysis has further postulated the existence of short-range order domains with structural remnants from boehmite precursors from which γ-Al2O3 is commonly prepared via calcination. In this PDF study, we now show that a recently theoretically found monoclinic δ5-Al2O3 phase is, in fact, best suited for describing the structure of different commercial Al2O3 supports, as well as a self-prepared and an industrial Ni/Al2O3 methanation catalyst. Furthermore, in situ experiments under catalytic cycling in the methanation reaction demonstrate that the nanoscale structure of this δ5 phase is preserved during cycling, pointing towards the high stability of the therein-represented disorder. A complete description of the disordered Al2O3 support structure is crucial in the field of heterogeneous catalysis in order to distinguish disorder within the bulk support from additional interfacial restructuring processes such as surface oxidation or spinel formation due to nanoparticle–support interactions during catalytic cycling in in situ scattering experiments.

Graphical Abstract

1. Introduction

In the quest for efficient heterogeneous catalysts, nanostructured γ-Al2O3 has garnered considerable attention as a highly stable support material. Its high industrial relevance is reflected in its usage across various catalytic processes [1,2,3]. While α-Al2O3 is the thermodynamically most stable modification, its comparably small specific surface of ca. 5 to 20 m2/g prevents its widespread application [4]. In contrast, γ-Al2O3 features a significantly higher surface area of about 150 to 200 m2/g, while being stable in temperature ranges of room temperature up to 700 °C, in which many heterogeneously catalyzed reactions can be run [5,6,7]. Due to this high surface area stemming from the micro- and mesoporosity of γ-Al2O3, catalytically active metal and metal oxide nanoparticles can be stabilized in a well-dispersed state without significant sintering for long periods of time in industrial applications such as the methanation reaction running on Ni/γ-Al2O3 [8,9].
Despite its wide usage, the crystal structure of γ-Al2O3 is still posing a challenge to structural refinements [10,11,12,13]. Originally, a cubic spinel structure with partial occupancies of the Al3+ cations (see Figure 1b) has been proposed based on the Rietveld analysis of powder X-ray diffraction (PXRD) and neutron diffraction supported by density functional theory (DFT) calculations [14,15,16,17]. Yet, the partial occupancies implemented in an average crystallographic model using a typically sized spinel unit cell (7.9 Å) [18] could not satisfactorily describe the disordered crystal structure of γ-Al2O3 [19]. To account for remaining misfits to the experimental data, tetragonal distortions [20,21] or non-spinel occupancies have been considered [18]. Furthermore, models employing supercells were created, which contained different stacking faults or antiphase boundaries [17]. An alternate description of the γ-Al2O3 structure was then suggested using a c-symmetry-based tetragonal model in space group I41/amd; see Figure 1a [22]. Since γ-Al2O3 is prepared from boehmite precursors via calcination, various studies have additionally pointed out that the actual γ-Al2O3 structure is highly dependent on the precursor structure and the history of thermal treatment [19,23].
With the rise of the pair distribution function (PDF) analysis of high-energy X-ray scattering data in the field of nanostructured materials [24], the structure of γ-Al2O3 has been revisited. The PDF, labeled G(r), can be understood as a histogram of the interatomic distances rij between the atoms i and j in the sample and is thus highly suited for structurally characterizing disordered materials:
G r = 1 N r i , j f i ( Q ) f j ( Q ) f ( Q ) 2 δ r r i j 4 π r ρ 0
The interatomic distances are weighted by the scattering powers of the two contributing atoms, i.e., the X-ray atomic form factor fi(Q) for atom i, and the average scattering length of all atoms f ( Q ) 2 , taking into account the atomic number density ρ0 and the total number of atoms N in the sample [25]. Experimentally, the PDF is gained by Fourier transformation of PXRD data over very large scattering angles. Hereby, this technique not only exploits the scattering from within the Bragg peaks but also exploits any diffuse scattering in between and below the Bragg peaks, which contains further information about the disorder. Using PDF analysis, Paglia et al. extended their previous tetragonal model of the average structure of γ-Al2O3 [22] by showing that the short-range structure of γ-Al2O3 deviates from the average one [19]. The local structure extends about 1 nm and was explained by stacking faults in the oxygen sublattice, being a remnant of the precursor structure boehmite [19,22].
Operando PXRD and PDF measurements of catalytic cycling are becoming increasingly popular to derive structure–activity relationships since the PDF can be used to track the structure and particle size of the catalyst particles, as well as simultaneously access the structure of the support material—even with subsecond time resolution at modern synchrotron radiation facilities [26,27,28]. Simultaneously, the detection and interpretation of ever smaller signals and so-far non-described structural residuals in refinements become ever more important for catalyst design in light of metal–support interactions, for instance.
Various PDF refinements on different γ-Al2O3 structures using the improved tetragonal model achieved by Paglia et al. [19] feature different degrees of structural residuals in the different curves of the fits. These residuals have been reasoned, for instance, with slightly different γ-Al2O3 structures due to the different calcination procedures of boehmite precursors or with interfacial restructuring of the γ-Al2O3 [8,29]. Yet, so far, an improved structural description has remained out of reach.
Very recently, for γ-Al2O3, Kovarik et al. found co-existing spinel and δ-Al2O3 domains in transmission electron microscopy (TEM) images, which supported previous ideas that γ-Al2O3 contains several crystallographic domains (twins, stacking faults, supercells) [30]. In earlier works, they had already investigated samples of δ-Al2O3 only and found them to be an intricate intergrowth of closely related polymorphs within the δ-Al2O3 family. These intergrowth structures had been visualized by scanning TEM (STEM) and called δ1, δ2, δ3, and δ4—complemented by a δ5 phase obtained by DFT calculation [31,32,33]. Those phases feature differences in their symmetry (space groups P212121, P21, A2/n, P-1, and A2, respectively), the number of atoms per unit cell, and the lattice parameters (see Figure S1 and Table S4).
In this study, we now employed these δ phases to refine experimental PDF data of different γ-Al2O3 supports and methanation catalysts. Herein, we show that, in many cases, catalysts are assumed to be built upon the γ-Al2O3 but feature a δ5-crystal structure instead.

2. Results and Discussion

2.1. Improved Description of γ-Al2O3 Materials with δ5 Phase

To improve the structural description of Al2O3 materials in PDF refinements, we compared the fit quality of different crystal structures to several Al2O3 materials, which are commercially available as γ-Al2O3: the most widely used spinel structure [18], the tetragonal model [22], and the recently established five intergrowth δ phases found by Kovarik et al. [31]. Figure 2 shows the PDF refinements of the Al2O3 support labeled as SBa200 by Sasol, Hamburg, Germany, being widely employed in catalysis for its high surface area, and Figure S7 compares the experimental PXRD pattern of SBa200 with the simulated PXRD patterns of these three crystallographic phases. The spinel phase describes the Al2O3 structure slightly better than the tetragonal phase in the PDF refinement, which is evidenced by the lower goodness-of-fit parameter Rw of 0.23 for the spinel compared to 0.25 for the tetragonal phase (for details on Rw calculation, see Section 3). For both phases, though, the short-range order for interatomic distances < 15 Å is not well described and shows higher residuals than the average noise level in the data in the difference curves, which is highlighted in Figure 2 by the boxes with dashed lines in the difference curves. The result values of all fit parameters for SBa200 are contained in Table S5, and the corresponding refinements for the supports SBa150 and TH100-150 by Sasol are shown in Figure S2 in the Supplementary Materials.
Fitting the five different δ phases to the experimental PDF of SBa200, phase δ4 did not describe the data at all. The phases δ1 and δ2 fit the data slightly worse or had a similar goodness of fit as the spinel and tetragonal structures, and δ3 improves the fit quality slightly. The monoclinic δ5 phase described the PDF and, in particular, the short-range order < 20 Å the best and better than the spinel model, as shown in Figure 2c, yielding a goodness-of-fit Rw = 0.18. The δ5 phase yielded the overall best-fit results for the supports SBa150 and TH100-150, too. Comparing the resulting fit values for Biso(O) and Biso(Al) for the spinel, tetragonal, and δ5 phase shows that the values are larger for the spinel and tetragonal structure. Both static disorder and thermal disorder can contribute to peak broadening in the PDF; the higher the disorder or thermal motion, the higher the Biso value. Given the higher static disorder with off-centered atomic positions in the δ5 phase, the Biso values for the spinel and tetragonal structure likely do not reflect higher thermal motion but try to make up for structural disorder, which is not included in the crystal structure of the spinel and tetragonal models.
The mismatch of the local structure with the spinel or tetragonal models was addressed before in the PDF study by Paglia et al. [19], where they proposed fitting a P1 structure to the local order from 1.2 to 8 Å in addition to the average tetragonal structure. This P1 model is highly parametrized, though. To prevent fits resulting in too-short bond lengths upon refinement of the atomic xyz positions, we did not refine the xyz positions and employed 17 parameters. Those refinements shown in Figure S3 did not improve the fit quality compared to fits using the δ5 phase.
Hence, this comparison of different crystal structures to the experimental PDF data of commercial γ-Al2O3 suggests that the crystal structure of what is referred to as a commercial γ-Al2O3 is, in fact, a monoclinic δ5 phase.

2.2. Supported Catalysts—Ni/Al2O3

The industrially employed Ni/Al2O3 methanation catalyst labeled SPP2080-IMRC has recently been investigated in a multi-method approach to establish a holistic description, to which we have contributed PDF characterization [8]. The PDF experiments were carried out in situ during reducing conditions (i.e., during catalyst activation). After activation, the nickel species exist as nanoparticles with face-centered cubic lattice (Nifcc). To structurally describe the experimental PDF, the fits consist of two phases: a Nifcc phase with a refinable particle diameter and an Al2O3 phase being either the tetragonal, the spinel, or the δ5 phase, as shown in Figure 3a–c. For all three fits, the nanoparticle diameter and Nifcc lattice parameter were refined to similar values of 2.3 nm and 3.54 Å. On the contrary, the fit qualities differed for the three Al2O3 phases, which was again best described by the δ5 phase. The residual curve of the δ5 phase clearly shows less structural signal in Figure 3c. For the catalyst before activation, the Nifcc phase was replaced by a NiO phase, but the δ5 phase very robustly also gave the best-fit results there, as shown in Figure S4.
In the original publication on the characterization of the SPP2080-IMRC Ni/Al2O3 catalyst [8], the unknown Al2O3 phase was optimized starting from a spinel structure with additional Al3+ ions on non-spinel positions to take into account some disorder. Yet, the short-range order could not be well described, and the goodness of fit remained worse, at Rw = 0.31, than that found here with the δ5 phase.
To investigate the phase stability of the δ5 phase during catalytic operation conditions as impinged by a renewable energy supply, we performed in situ methanation experiments on a model Ni/Al2O3 catalyst at beamline ID15, ESRF. This catalyst with 15 wt% Ni on the SBa200 support was cycled between catalysis (H2:CO2 = 4:1) and hydrogen dropouts at 425 °C four times over a total of 4 h in a flow cell reactor after activation in pure hydrogen (see Methods for details of in situ experiment). Hydrogen dropouts mimic the shortage of hydrogen in an industrial scenario when hydrogen is produced from electrocatalytic water splitting via renewable energies, which undergo fluctuations (sun, wind) [34]. The biphasic PDF refinements of the Nifcc nanoparticle and the δ5 phase revealed that the δ5 phase is structurally stable. The average domain size of the δ5 phase increases only slightly during the catalytic cycling by ca. 60 Å to 69 Å, as shown Figure S5a, which is insignificant growth given the spread of data points and having an interatomic Al-O bond length of 1.82 Å in the tetrahedral sites and 1.90 Å for the octahedral sites. This shows that the nanoscale structure of the δ5 phase is preserved under catalytic cycling, pointing towards the high stability of the therein-represented nanostructured motifs and distortions.

3. Experimental Section

3.1. Synthesis of Ni/Al2O3

The support materials of the nominally γ-Al2O3 structures called SBa200, SBa150, and TH100-150 were provided by Sasol, Hamburg, Germany. These supports differ in surface area and porosity; see Table S6 and Figure S8 in the Supplementary Materials. These support materials were used without further purification or treatment for the preparation of Ni/γ-Al2O3 catalysts following a urea precipitation method according to Grunwaldt et al. [9]: Urea was added under constant stirring into a Ni(NO3)2 water solution to yield a molar ratio of urea: Ni = 2:1 at room temperature. An Al2O3 water suspension (1.6 g Al2O3 in 50 mL water) was added, and the reaction was run for 48 h at 90 °C while constantly stirring. The green precipitate was washed with water and dried overnight in an oven at 90 °C, and the resulting powder was calcined at 400 °C for 3 h.

3.2. Characterization

Surface area and porosity measurements of the three supports SBa200, SBa150, and TH100-150 were performed with the BET-BJH method by adsorption of N2 (assuming 16.2 Å2 as the area of a N2 molecule) at the temperature of liquid nitrogen after outgassing the samples at 400 °C for 2 h. The instrument used for this analysis was the 3P macro 200C, 3P Instruments, Odelzhausen, Germany, and the results are shown in Figure S8.
The industrial methanation catalyst of the SPP2080, Ni/Al2O3, labeled SPP2080-IMRC, was previously thoroughly characterized by Weber et al. [8]. This catalyst features 8.6 wt% Ni.

3.3. X-ray Total Scattering Experiments

Synchrotron: Total scattering data of the powder samples were acquired at the ID15A beamline at the ESRF (Grenoble, France) with an X-ray energy of 65 keV (λ = 0.1907 Å) and a beam size of 150 × 150 μm2 in quartz capillaries with a diameter of 1 mm. Every data set was collected for 1 s with a Pilatus CdTe 2M detector (DECTRIS, Baden-Daettwil, Switzerland). Radial integration was performed with the software xpdtools, version 0.8 [35] and PDF calculation with PDFgetX3, version 2.2.1 [36] using a Qmax of 24 Å−1. The instrumental resolution was determined from a LaB6 standard with qdamp = 0.02 and qbroad = 0.009.
In situ catalysis experiment at synchrotron: These in situ experiments were carried out in a custom-built quartz capillary reactor (outer diameter 1.5 mm, wall thickness 0.05 mm) with heating coils for temperature control and a gas dosing system with mass flow controllers (Bronkhorst, AK Ruurlo, The Netherlands) to adjust the gas flow of He, H2, and CO2, with an overall gas flow of 10 mL min−1, while using a catalyst amount of 7–10 mg. The catalytic and dropout cycles were alternated every 30 min. Data were measured continuously with 1 s for each detector image and 29 s sleep time after each image to prevent ghosting issues on the Pilatus CdTe 2M detector.
Industrial methanation catalyst SPP2080-IMRC: For PDF data collection and processing of the industrial methanation catalyst SPP2080-IMRC, see Weber et al. [8].

3.4. PDF Modeling

PDF refinements were conducted with Diffpy-CMI, version 3.0 [37]. All data sets were fit over a range of 1–80 Å for comparability of fit results between the different supports. For all refinements, first, the lattice parameter(s) and scale were refined, then the domain size and shape parameters of the characteristic functions for the spherical nanoparticles, followed by the atomic isotropic atomic displacement parameter Biso and the parameter δ2 for the correlated atomic motion of nearest neighbors. Finally, occupancies were freed, if included in the model. The fractional atomic coordinates xyz were not fitted to prevent overfitting and unphysical bond distances.
The goodness of fit is defined by Equation (2), where Gobs(r) and Gcalc(r) are the observed and modeled PDFs, w(rr) is the weighting factor of each data point I, and s is the scale factor of the model to the data [38].
R w = i w r i G o b s r i s G c a l c r i 2 i w r i G o b s r i 2 1 2
For the spinel phase, 10 parameters were refined: lattice parameter a, domain size, scale, δ2, and Biso(O). For the four Al sites, one occupancy was refined for each, while their four Biso parameters were combined into a single parameter Biso(Al). The tetragonal phase was refined with 10 parameters: lattice parameters a and c, domain size, scale, δ2, and Biso(O). Again, the Biso of the three Al sites was combined. The oxygen occupancies were not refined, but the three Al ones were. For both spinel and tetragonal models, the oxygen occupancies were not refined, and the resulting elemental ratio Al:O was crosschecked to be close to the nominal Al2O3 composition upon Al occupancy refinement.
The local structure, according to the Paglia model, was fitted with a P1 phase for 1.2 < r < 8 Å, refining lattice parameters a, b, and c; domain size; scale, δ2; and individual Biso(O) and Biso(Al) for each atom, resulting in 17 parameters.
For the δ5 phase, 9 parameters were used: lattice parameters a, b, and c; angle β; domain size; scale; δ2; Biso(Al); and Biso(O). Since the δ5 phase has 42 Wyckoff sites, we used one Biso value for all atom sites of one element. Partial occupancies and xyz positions were not refined for any site.
For refining the in situ data, the fit results of one data set were used as the starting point for the subsequent data set. Additionally, when fit results showed a phase to be <5% of the summed scale, the fit of this data set was repeated without this phase. The value of 5% was chosen because this was the threshold where the fit was able to describe the data, and, for phase fractions < 5%, these phases merely fitted noise.

4. Conclusions

Understanding the structure of a disordered support material, such as γ-Al2O3, is essential for deducing structure–activity correlations and achieving ultimate insight into particle–support interactions, for instance. In this study, the power of PDF analysis for characterizing nanostructured materials was highlighted again in the field of heterogeneous catalysis. For different commercial Al2O3 supports, as well as a self-prepared and an industrial Ni/Al2O3 methanation catalyst, we obtained a superior structural description using a recently found monoclinic δ5 phase compared to the established spinel and tetragonal models. Furthermore, the δ5 structure remained stable during catalytic cycling, which proves the robustness of this nanostructure against restructuring. Although, pretty commonly, various Al2O3 supports have been associated and labeled as being γ-Al2O3, this PDF study of their local structure revealed that the structure of a significant number of Al2O3 supports is actually better described by the δ5 phase and could, thus, better be called δ5-Al2O3 in the future. More importantly, though, in order to improve the structural characterization of catalytic systems for enhanced structure–activity correlations, we hope that this improved δ5 structure finds its way into databanks and broad application across communities soon. A complete description of the disordered Al2O3 support structure is crucial to the field of heterogeneous catalysis in order to distinguish general disorder within the bulk support from additional interfacial restructuring processes upon catalyst loading or during catalyst cycling, such as surface oxidation or spinel formation due to nanoparticle-support interactions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14040238/s1. Supplementary figures and tables for characterization include tables of PDF fit results, graphs of PDF refinements, XRD, and BET curves. References [18,19,31,32,33,37,39] are cited in the Supplementary Materials.

Author Contributions

F.M. and B.M.D.F. analyzed the PDF data. M.Z. conceptually laid out and led the project. All authors contributed to data analysis and discussion, with a particular effort by F.M. in interpretation. F.M. and B.M.D.F. wrote parts of the original manuscript, and M.Z. contributed and revised it. All authors have read and agreed to the published version of the manuscript.

Funding

We thank the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) via SPP2080 (ZO 369/2-2) and the Exzellenzcluster 2186 ‘The Fuel ScienceCenter’ (ID: 390919832) for financial support.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

We acknowledge beamtime at beamline ID15A at ESRF (Grenoble, France), proposal CH-6069, and local support by Stefano Checchia, as well as support by our former colleague Nils Prinz. We acknowledge Sasol for the provision of samples SBa200, SBa150, and TH100-150, as well as Ronny Zimmermann for the provision of the SPP2080 reference catalyst.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Busca, G.; Spennati, E.; Riani, P.; Garbarino, G. Looking for an Optimal Composition of Nickel-Based Catalysts for CO2 Methanation. Energies 2023, 16, 5304. [Google Scholar] [CrossRef]
  2. Hu, F.; Ye, R.; Lu, Z.-H.; Zhang, R.; Feng, G. Structure–Activity Relationship of Ni-Based Catalysts toward CO2 Methanation: Recent Advances and Future Perspectives. Energy Fuels 2022, 36, 156–169. [Google Scholar] [CrossRef]
  3. Lee, Y.H.; Ahn, J.Y.; Nguyen, D.D.; Chang, S.W.; Kim, S.S.; Lee, S.M. Role of oxide support in Ni based catalysts for CO2 methanation. RSC Adv. 2021, 11, 17648–17657. [Google Scholar] [CrossRef]
  4. Rascón, F.; Wischert, R.; Copéret, C. Molecular nature of support effects in single-site heterogeneous catalysts: Silica vs. alumina. Chem. Sci. 2011, 2, 1449. [Google Scholar] [CrossRef]
  5. Candia, R.; Sørensen, O.; Villadsen, J.; Topsøe, N.Y.; Clausen, B.S.; Topsøe, H. Effect of Sulfiding Temperature on Activity and Structures of CO-MO/Al2O3 Catalysts. ii. Bull. Sociétés Chim. Belg. 1984, 1984, 763–774. [Google Scholar] [CrossRef]
  6. Zhang, J.; Xu, H.; Jin, X.; Ge, Q.; Li, W. Characterizations and activities of the nano-sized Ni/Al2O3 and Ni/La–Al2O3 catalysts for NH3 decomposition. Appl. Catal. A Gen. 2005, 290, 87–96. [Google Scholar] [CrossRef]
  7. Centeno, M.A.; Hadjiivanov, K.; Venkov, T.; Klimev, H.; Odriozola, J.A. Comparative study of Au/Al2O3 and Au/CeO2-Al2O3 catalysts. J. Mol. Catal. A Chem. 2006, 252, 142–149. [Google Scholar] [CrossRef]
  8. Weber, S.; Zimmermann, R.T.; Bremer, J.; Abel, K.L.; Poppitz, D.; Prinz, N.; Ilsemann, J.; Wendholt, S.; Yang, Q.; Pashminehazar, R.; et al. Digitization in Catalysis Research: Towards a Holistic Description of a Ni/Al2O3 Reference Catalyst for CO2 Methanation. ChemCatChem 2022, 14, e202101878. [Google Scholar] [CrossRef]
  9. Mutz, B.; Gänzler, A.; Nachtegaal, M.; Müller, O.; Frahm, R.; Kleist, W.; Grunwaldt, J.-D. Surface Oxidation of Supported Ni Particles and Its Impact on the Catalytic Performance during Dynamically Operated Methanation of CO2. Catalysts 2017, 7, 279. [Google Scholar] [CrossRef]
  10. Prins, R. Location of the Spinel Vacancies in γ-Al2O3. Angew. Chem. Int. Ed Engl. 2019, 58, 15548–15552. [Google Scholar] [CrossRef]
  11. Prins, R. On the structure of γ-Al2O3. J. Catal. 2020, 392, 336–346. [Google Scholar] [CrossRef]
  12. Samain, L.; Jaworski, A.; Edén, M.; Ladd, D.M.; Seo, D.-K.; Javier Garcia-Garcia, F.; Häussermann, U. Structural analysis of highly porous γ-Al2O3. J. Solid State Chem. 2014, 217, 1–8. [Google Scholar] [CrossRef]
  13. Sun, M.; Nelson, A.E.; Adjaye, J. Examination of spinel and nonspinel structural models for gamma-Al2O3 by DFT and rietveld refinement simulations. J. Phys. Chem. B 2006, 110, 2310–2317. [Google Scholar] [CrossRef] [PubMed]
  14. Verwey, E.J.W. The Crystal Structure of γ-Fe2O3 and γ-Al2O3. Z. Für Krist.-Cryst. Mater. 1935, 91, 65–69. [Google Scholar] [CrossRef]
  15. Ushakov, V.A.; Moroz, E.M. Structure of low-temperature γ- and η-Al2O3. React. Kinet. Catal. Lett. 1984, 24, 113–118. [Google Scholar] [CrossRef]
  16. Zhou, R.-S.; Snyder, R.L. Structures and transformation mechanisms of the η, γ and θ transition aluminas. Acta Crystallogr. Sect. B Struct. Sci. 1991, 47, 617–630. [Google Scholar] [CrossRef]
  17. Rudolph, M.; Motylenko, M.; Rafaja, D. Structure model of γ-Al2O3 based on planar defects. IUCrJ 2019, 6, 116–127. [Google Scholar] [CrossRef] [PubMed]
  18. Smrcok, L.; Langer, V.; Krestan, J. Gamma-alumina: A single-crystal X-ray diffraction study. Acta Crystallogr. C 2006, 62, i83–i84. [Google Scholar] [CrossRef] [PubMed]
  19. Paglia, G.; Božin, E.S.; Billinge, S.J.L. Fine-Scale Nanostructure in γ-Al2O3. Chem. Mater. 2006, 18, 3242–3248. [Google Scholar] [CrossRef]
  20. Saalfeed, H. The Dehydration of Gibbsite and the Structure of a Tetragonal γ-Al2O3. Clay Miner. 1958, 3, 249–257. [Google Scholar] [CrossRef]
  21. Wilson, S.J. The dehydration of boehmite, γ-AlOOH, to γ-Al2O3. J. Solid State Chem. 1979, 30, 247–255. [Google Scholar] [CrossRef]
  22. Paglia, G.; Buckley, C.E.; Rohl, A.L.; Hunter, B.A.; Hart, R.D.; Hanna, J.V.; Byrne, L.T. Tetragonal structure model for boehmite-derived γ-alumina. Phys. Rev. B 2003, 68, 144110. [Google Scholar] [CrossRef]
  23. Xie, Y.; Kocaefe, D.; Kocaefe, Y.; Cheng, J.; Liu, W. The Effect of Novel Synthetic Methods and Parameters Control on Morphology of Nano-alumina Particles. Nanoscale Res. Lett. 2016, 11, 259. [Google Scholar] [CrossRef] [PubMed]
  24. Billinge, S.J.L.; Kanatzidis, M.G. Beyond crystallography: The study of disorder, nanocrystallinity and crystallographically challenged materials with pair distribution functions. Chem. Commun. 2004, 749–760. [Google Scholar] [CrossRef]
  25. Billinge, S.J.L. Nanostructure studied using the atomic pair distribution function. Z. Kristallogr. Suppl. 2007, 26, 17–26. [Google Scholar] [CrossRef]
  26. Newton, M.A.; Di Michiel, M.; Ferri, D.; Fernàndez-Garcia, M.; Beale, A.M.; Jacques, S.D.M.; Chupas, P.J.; Chapman, K.W. Catalytic Adventures in Space and Time Using High Energy X-rays. Catal Surv Asia 2014, 18, 134–148. [Google Scholar] [CrossRef]
  27. Zimmerli, N.K.; Müller, C.R.; Abdala, P.M. Deciphering the structure of heterogeneous catalysts across scales using pair distribution function analysis. Trends Chem. 2022, 4, 807–821. [Google Scholar] [CrossRef]
  28. Chupas, P.J.; Chapman, K.W.; Chen, H.; Grey, C.P. Application of high-energy X-rays and Pair-Distribution-Function analysis to nano-scale structural studies in catalysis. Catal. Today 2009, 145, 213–219. [Google Scholar] [CrossRef]
  29. Schlicher, S.; Prinz, N.; Bürger, J.; Omlor, A.; Singer, C.; Zobel, M.; Schoch, R.; Lindner, J.K.N.; Schünemann, V.; Kureti, S.; et al. Quality or Quantity? How Structural Parameters Affect Catalytic Activity of Iron Oxides for CO Oxidation. Catalysts 2022, 12, 675. [Google Scholar] [CrossRef]
  30. Kovarik, L.; Bowden, M.; Khivantsev, K.; Kwak, J.H.; Szanyi, J. Structural complexity of γ-Al2O3: The nature of vacancy ordering and the structure of complex antiphase boundaries. Acta Mater. 2024, 266, 119639. [Google Scholar] [CrossRef]
  31. Kovarik, L.; Bowden, M.; Shi, D.; Szanyi, J.; Peden, C.H.F. Structural Intergrowth in δ-Al2O3. J. Phys. Chem. C 2019, 123, 9454–9460. [Google Scholar] [CrossRef]
  32. Kovarik, L.; Bowden, M.; Shi, D.; Washton, N.M.; Andersen, A.; Hu, J.Z.; Lee, J.; Szanyi, J.; Kwak, J.-H.; Peden, C.H.F. Unraveling the Origin of Structural Disorder in High Temperature Transition Al2O3: Structure of θ-Al2O3. Chem. Mater. 2015, 27, 7042–7049. [Google Scholar] [CrossRef]
  33. Kovarik, L.; Bowden, M.; Genc, A.; Szanyi, J.; Peden, C.H.F.; Kwak, J.H. Structure of δ-Alumina: Toward the Atomic Level Understanding of Transition Alumina Phases. J. Phys. Chem. C 2014, 118, 18051–18058. [Google Scholar] [CrossRef]
  34. Kalz, K.F.; Kraehnert, R.; Dvoyashkin, M.; Dittmeyer, R.; Gläser, R.; Krewer, U.; Reuter, K.; Grunwaldt, J.-D. Future Challenges in Heterogeneous Catalysis: Understanding Catalysts under Dynamic Reaction Conditions. ChemCatChem 2017, 9, 17–29. [Google Scholar] [CrossRef] [PubMed]
  35. Wright, C.J.; Zhou, X.D. Computer-assisted area detector masking. J. Synchrotron Radiat. 2017, 24, 506–508. [Google Scholar] [CrossRef] [PubMed]
  36. Juhás, P.; Davis, T.; Farrow, C.L.; Billinge, S.J.L. PDFgetX3: A rapid and highly automatable program for processing powder diffraction data into total scattering pair distribution functions. J. Appl. Crystallogr. 2013, 46, 560–566. [Google Scholar] [CrossRef]
  37. Juhás, P.; Farrow, C.L.; Yang, X.; Knox, K.R.; Billinge, S.J.L. Complex modeling: A strategy and software program for combining multiple information sources to solve ill posed structure and nanostructure inverse problems. Acta Crystallogr. A Found. Adv. 2015, 71, 562–568. [Google Scholar] [CrossRef] [PubMed]
  38. Peterson, P.F.; Božin, E.S.; Proffen, T.; Billinge, S.J.L. Improved measures of quality for the atomic pair distribution function. J Appl. Crystallogr. 2003, 36, 53–64. [Google Scholar] [CrossRef]
  39. Macrae, C.F.; Sovago, I.; Cottrell, S.J.; Galek, P.T.A.; McCabe, P.; Pidcock, E.; Platings, M.; Shields, G.P.; Stevens, J.S.; Towler, M.; et al. Mercury 4.0: From visualization to analysis, design and prediction. J. Appl. Crystallogr. 2020, 53, 226–235. [Google Scholar] [CrossRef]
Figure 1. Crystal structures of the (a) tetragonal, (b) spinel, and (c) δ5 phases of Al2O3 (from left to right), shown as projections into the ac plane. Oxygen atoms in red, aluminum atoms in blue, with partial occupancies indicated with partially filled blue circles.
Figure 1. Crystal structures of the (a) tetragonal, (b) spinel, and (c) δ5 phases of Al2O3 (from left to right), shown as projections into the ac plane. Oxygen atoms in red, aluminum atoms in blue, with partial occupancies indicated with partially filled blue circles.
Catalysts 14 00238 g001
Figure 2. Refinements of the experimental PDF, G(r), for the Al2O3 support SBa200 by using the (a) tetragonal, (b) spinel, and (c) δ5 structures. Experimental PDF data, G(r) obs, in blue circles, fitted calculated PDF, G(r) calc, as red line, and the difference curve in grey in offset, with the goodness-of-fit Rw in each panel. Boxes in dashed lines around difference curves indicate the standard deviation of data points over the short range up to 20 Å and the medium range > 20 Å.
Figure 2. Refinements of the experimental PDF, G(r), for the Al2O3 support SBa200 by using the (a) tetragonal, (b) spinel, and (c) δ5 structures. Experimental PDF data, G(r) obs, in blue circles, fitted calculated PDF, G(r) calc, as red line, and the difference curve in grey in offset, with the goodness-of-fit Rw in each panel. Boxes in dashed lines around difference curves indicate the standard deviation of data points over the short range up to 20 Å and the medium range > 20 Å.
Catalysts 14 00238 g002
Figure 3. Refinement of experimental PDF data of an industrial methanation catalyst Ni/Al2O3 using two phases, with one phase being Nifcc nanoparticles and the second being the (a) tetragonal, (b) spinel, and (c) δ5 phases for describing Al2O3 support. Experimental PDF data, G(r) obs, in blue circles, fitted calculated PDF, G(r) calc, as red line, and the difference curve in grey in offset, with the goodness-of-fit Rw in each panel. The contribution of the two phases to the total fit curve is shown in the offset: in purple for Al2O3 and in brown for Nifcc.
Figure 3. Refinement of experimental PDF data of an industrial methanation catalyst Ni/Al2O3 using two phases, with one phase being Nifcc nanoparticles and the second being the (a) tetragonal, (b) spinel, and (c) δ5 phases for describing Al2O3 support. Experimental PDF data, G(r) obs, in blue circles, fitted calculated PDF, G(r) calc, as red line, and the difference curve in grey in offset, with the goodness-of-fit Rw in each panel. The contribution of the two phases to the total fit curve is shown in the offset: in purple for Al2O3 and in brown for Nifcc.
Catalysts 14 00238 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Manzoni, F.; Fahl, B.M.D.; Zobel, M. Improved Structural Description of Different γ-Al2O3 Materials Using Disordered δ5-Al2O3 Phase via X-ray Pair Distribution Function Analysis. Catalysts 2024, 14, 238. https://doi.org/10.3390/catal14040238

AMA Style

Manzoni F, Fahl BMD, Zobel M. Improved Structural Description of Different γ-Al2O3 Materials Using Disordered δ5-Al2O3 Phase via X-ray Pair Distribution Function Analysis. Catalysts. 2024; 14(4):238. https://doi.org/10.3390/catal14040238

Chicago/Turabian Style

Manzoni, Fabio, Benjamin M. D. Fahl, and Mirijam Zobel. 2024. "Improved Structural Description of Different γ-Al2O3 Materials Using Disordered δ5-Al2O3 Phase via X-ray Pair Distribution Function Analysis" Catalysts 14, no. 4: 238. https://doi.org/10.3390/catal14040238

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop