Philipp
Simons‡
a,
Kierstin P.
Torres‡
a and
Jennifer L. M.
Rupp
*ab
aDepartment of Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. E-mail: jrupp@mit.edu
bDepartment of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
First published on 22nd March 2022
Spray pyrolysis offers excellent control over microstructure and cost-effective processing of ceria ceramic thin films for many applications such as catalysis, solid oxide fuel cells, solar-to-fuel conversion, and resistive switching. To control the properties of ceria thin films that are relevant for such applications, such as crystal orientation, grain size, crystallinity, and to achieve crack-free deposition, systematic processing guidelines for sprayed ceria are needed. In particular, the crystal orientation of ceria plays a significant role in the catalytic activity of ceria, and therefore controlling and measuring crystallite texture for catalytic applications is highly relevant for catalysis. In this work, we systematically scan the parameter space of deposition and post-annealing parameters to control various mechanical and structural properties that impact film functionality. We identify a critical film thickness of ceria during wet-chemical deposition of 300 nm in the initial amorphous phase, due to the interplay of ceria crystallization, solvent evaporation, and thermal expansion. Through XRD, the transition from amorphous to the cubic crystalline phases and grain growth with annealing is observed. Furthermore, a thermally induced, prominent shift in texture from the [111] to [200] direction is detected, which corresponds to the {100} ceria facet of high catalytic activity. Interestingly, we observe that the peak position of the Raman F2g stretching mode correlates with the measured texture coefficient and exceeds the bulk value of 465 cm−1 for highly textured films. While it has been previously observed that thermal annealing can induce stark texture in ceria thin films, we contribute with the new insight that Raman spectroscopy can be used to detect such texture. Overall, this offers a combination of cheap processing and simple characterization of the ceria crystal orientation.
Thin films have been demonstrated to differ in their electro-chemo-mechanics14 from their bulk counterparts, exhibiting higher amounts of strain ascribed to interactions with the substrate,15,16 and chemical strain from changes in the lattice and their defect chemistry.17 Also, it was reported that strain can impact their ionic and electronic conductivities, defect formation and association.16,18 For example, free-standing ceria membranes have been investigated as a model system for micro solid oxide fuel cell membranes, and the impact of local stress–strain fields on the oxygen ionic conductivity of ceria solid solutions was confirmed for device integration.5,19 Other examples for electro-chemo-mechanic interactions have been demonstrated for gadolinia-doped ceria (GDC) with giant electrostriction,20,21 and their suitability as electrochemical actuators.22
Besides strain affecting the defect chemistry, also the microstructure of ceria such as the grain size and grain boundary density, as well as crystal orientation play a critical role in controlling properties such as catalytic activity or conductivity. For example, nanocrystalline ceria exhibits dramatically higher electronic conductivity than bulk or microcrystalline ceria, due to size dependent grain boundary segregation of impurities.23 Ceria surfaces with different orientations show different levels of catalytic activity, with (100) planes showing higher catalytic activity than (111) planes,24 due to the more favorable energetics for oxygen vacancy formation on the open (100) surface when compared to the closed-packed (111) surface.25 However, computational and experimental results so far largely indicate that the (111) plane is the lowest energy state, and therefore the preferred termination in randomly oriented powders as well as thin films.26,27 Given the different catalytic activity of the different crystal terminations, it is important to be able to control the orientation of ceria during synthesis, and particularly to be able to do so at scale for commercial applications. There have been demonstrations of assembling ceria nanocrystals with controlled surface exposure, using hydrothermal synthesis.28 Interestingly, multiple studies have furthermore shown that upon thermal annealing or deposition under relatively high temperature, texture can be induced into previously randomly oriented ceria, with a preference for the development of the higher energy (100) plane.29–33 These findings indicate that the thermodynamic description of the relative order of stability of the different ceria planes is incomplete. Ultimately, the ability to control the preferred ceria crystal orientation in a simple and scalable way is important in engineering ceria thin films for catalytic applications.
Given their relatively low cost and scalability, wet-chemical thin film deposition techniques for ceria and its solid solutions such as spray pyrolysis are advantageous in comparison with vacuum based techniques which often require more expensive and complex equipment.34 Additionally, wet-chemical synthesis techniques typically require significantly lower processing temperatures often at one fifth to one sixth of the melting temperature (between 300 °C and 500 °C for ceria), and do not require sintering to densify as is common for pellet processing from powders. The ceramic wet-chemical synthesis method of spray pyrolysis offers good control over the microstructure35 and a low-cost option to coat large areas with thin films at low temperatures. Unlike ceramic manufacturing routes that involve sintering of powder green bodies, it is a characteristic feature of spray pyrolysis that the as-deposited films are typically amorphous or biphasic amorphous-crystalline, which can be transformed to their crystalline phase by thermal annealing.35–37 In spray pyrolysis, metal salt is dissolved in an organic solvent, and this precursor is sprayed on a heated substrate. Upon impact on the heated substrate, the organic solvent evaporates and the metal salt simultaneously crystallizes, forming a ceramic oxide thin film. An interesting feature of wet-chemical synthesis techniques is their use of organic precursors, which have been shown to control the amorphous-to-crystalline structure through their chain length, and to incorporate into the film structure and act as space fillers during the metal-salt-to-oxide pyrolysis reaction.38 In addition, the choice of organic precursor is critical in controlling the thin film deposition because the temperatures of solvent evaporation and metal salt crystallization need to be tuned such that both reactions occur simultaneously. Consequently, the choice of solvent greatly affected the film's short-range order and crystallization kinetics. In summary, amorphous films deposited via wet-chemical techniques have fundamental structural deviations from classic crystalline ceria synthesized at high temperatures of up to two thirds of the melting temperature, via particle densification.
Despite previous work on the crystallinity and grain growth of ceria thin films deposited by spray pyrolysis,30,35,38 systematic studies on the critical film thickness, and crystal texture evolution as a function of deposition and annealing temperature are still missing to date. In order to fully harness the potential of spray pyrolysis, it is necessary to understand how the different processing parameters affect the material structure, and subsequently, relevant properties such as ionic conductivity. Clear guidelines are needed on how processing parameters affect the mechanical stability, as well as critical thresholds that lead to processing failures such as cracking. In particular, the rational engineering of ceria catalyst materials requires good control over the crystal orientation, and it is desired to achieve cheap and scalable pathways to assemble high catalytic activity ceria with 〈100〉 orientation, and to characterize the structure of such materials quickly and easily. With singular reports that thermal annealing can induce a preferred 〈100〉 orientation – which was assumed thermodynamically unfavorable in ceria – we see promise in systematically studying thermal processing parameters of sprayed ceria to induce texture evolution in a controlled fashion. This would provide a new, simple, and cheap tool for the rational design of catalytic materials.
In this work, we systematically study the impact of deposition parameters and thermal treatment on the mechanical and lattice vibrational properties of cerium oxide. For this, we select ceria thin films deposited via spray pyrolysis, a wet-chemical precipitation-based deposition technique, and subject thin films to various thermal annealing conditions to vary crystallinity and study thin film orientation. We determine critical film thicknesses to provide guidelines for practical thin film deposition. Moreover, we complement crystal structure analyses through X-ray diffraction with Raman spectroscopy to study how vibrational properties correlate with the crystal orientation and texture of thin films. Through this, we collectively combine the low-cost fabrication method of spray pyrolysis with accessible Raman spectroscopy, and demonstrate how these methods can be used to fabricate and characterize ceria thin films with controlled crystal orientation for catalytic and electrochemical applications.
To determine the critical thickness of sprayed ceria thin films, we deposited films on Si3N4 coated (100) silicon wafers, arranged such that the center of a wafer was aligned with the spray gun nozzle. Fig. 1a schematically shows the spraying plume in which the atomized precursor is deposited from the spray gun onto the substrate in a conical distribution. In this flow pattern, precursor flow was highest at the center of the spray cone and radially reduced outward. This resulted in a radial thickness gradient along the wafer schematically depicted in Fig. 1a where higher volumes of the precursor are deposited closer to the center of the wafer. This inherent thickness gradient created a controlled means of determining the critical thickness during a single deposition. The film thickness distribution was then measured via surface profilometry at evenly spaced points from the center of each wafer outwards.
Fig. 1b displays the film thickness measured via profilometry as a function of distance from wafer center for films deposited at 350 °C, 400 °C, 450 °C, and 500 °C. In spray pyrolysis, the temperature needs to be high enough such that the organic precursor undergoes pyrolytic decomposition during deposition, i.e., the simultaneous evaporation of the organic solvent and the crystallization of the metal salt upon impact on the heated substrate. As such, we selected temperatures near and above the boiling point of the highest boiling point precursor used, bis-2-ethyl hexyl phthalate. Controlling the deposition temperature during the spray pyrolysis process allows for control over the rates of the two concurrent reactions, evaporation and crystallization, which in turn controls the deposition rate, microstructure, density, and potential formation of organic residues in the film. At the center, films are the thickest, with a maximum thickness of approximately 2.5 μm, and radially decreasing thickness from the center. Optical micrographs were taken at each profilometry measurement point to determine the relationship between film thickness and film morphology. From optical micrographs of the film deposited at 500 °C in Fig. 1a (refer to ESI 3† for the remaining three deposition temperatures), one can see at the center of the spray cone and at a distance of 1 cm from it, that the films exhibit a substantial amount of cracking on the range of tens of microns in crack length and also some loose powder at the surface. On the other hand, at a position of the substrate about 2–3 cm away from the center, films are dense, smooth, and crack-free. Moreover, the films show typical signs of dense pyrolysis films with rims of dried micron-sized droplets originating from the droplet-by-droplet deposition nature of spray pyrolysis. By correlating optical micrographs and measured thicknesses of films at each point and deposition temperature, we confirm successful film deposition at all deposition temperatures and observe cracking when the thickness exceeded approximately 300 nm with a maximum thickness measured of 2.5 μm, Fig. 1b. An interesting observation is that this critical thickness, defining dense and crack-free film deposition, is independent of deposition temperature. In the following, we discuss in more detail the origin of this temperature-independent critical thickness.
The upper limit of thin film thickness during spray pyrolysis deposition, and crack formation above this critical thickness, is driven by two main factors: first, the formation of cracks is initiated when it is energetically favorable to release strain energy, which here is due to amorphous or mixed phase films with out-of-equilibrium bond lengths, as well as incorporated solvents. Second, during wet-chemical deposition, the evaporation of precursor on the substrate surface and the droplet deposition of additional precursor occurs simultaneously. Due to the steady competition between solidification and related loss of volume, and the arrival of new liquid solvent, crack formation is dependent on the flow rate, which needs to be below the evaporation rate of precursor solvent. On the first factor, one can turn to Griffith fracture mechanics.39 This widely used model is applicable to brittle glasses and ceramics; for example, it has been used to understand the critical thickness of amorphous silicon thin films for Li-ion battery electrodes.40 When cracks form, there is an increase in surface energy that stems from the formation of two new surfaces at the crack and a decrease in strain energy due to the release of internal stress. According to the Griffith criterion, cracking occurs when the release of strain energy is greater than or equal to the increase in surface energy due to cracking. Thin films exhibit a high amount of strain, ascribed to the mixed-phase, largely amorphous nature of sprayed ceria thin films, as well as precursor residues present in sprayed thin films. In the present study of sprayed ceria thin films, we can apply the reasoning of the Griffith criterion: as film thickness increases, the strain energy in the film increases, reaching a point at the critical thickness where the cracks can form to alleviate the buildup in strain energy (see ESI†). Because the total film strain energy is dependent on thickness, there exists a critical thickness in which there is enough strain energy for cracking to occur. When films are thicker than the critical thickness, cracks form. On the second factor, one expects cracks to form, and powdery deposition to occur, when the incoming precursor flow rate exceeds the evaporation rate of solvents upon impact with the substrate. The spray plume in the present experiments exhibits a radial flow rate gradient with decreasing precursor flow rate away from the spray plume center. Therefore, as we move outward from the center of the plume, the flow rate relative to the evaporation rate is reduced, leading to a geometric point where dense films start to form. We explain the temperature-independent critical film thickness of 300 nm observed here by the complex interplay of these two factors, which both feature a characteristic critical value as it is experimentally observed in sprayed ceria thin films.
We first investigate the as-deposited films. Fig. 2 depicts the XRD data of the as-deposited films deposited at 350 °C, 400 °C, 450 °C, and 500 °C. XRD peaks of as-deposited films are broad with low intensity, showing a slight increase in intensity and decrease in full-width half-maximum (FWHM) at higher deposition temperatures. Broad peaks with low intensity are characteristic for ceria deposited via spray pyrolysis and indicate that a significant fraction of the film is amorphous.35 The increasing intensity and sharpness of peaks for increasing deposition temperature is caused by thermally promoted nucleation and grain growth. In both Fig. 2 and 3, it is observed that the XRD peaks of the films deposited at 400 °C are generally broader than at 350 °C. We attribute this to fluctuations in film thickness, specifically that the films deposited at 400 °C are thinner than the other samples, leading to an overall weaker XRD signal.
We detail the effect of post-annealing temperature between 500 °C and 800 °C on the synthesized film microstructure with the corresponding XRD patterns in Fig. 3a–d. No cracking was observed for any of those thin films. In general, we see the following qualitative trends: firstly, increasing the deposition temperature results in slight increases in peak intensity and decrease in FWHM, and secondly, increasing the post-anneal temperature causes peak FWHM to drastically decrease by approximately 60–67% and intensity to further increase. To analyze this in more detail, we plot in Fig. 4 average grain sizes calculated using the Scherrer equation, based on the FWHM value of the (111) peak, as a function of post-annealing and deposition temperature. Strictly speaking, the Scherrer equation provides a lower bound on crystallite size. Here, we corrected for instrument peak broadening, but not for other second order effects such as micro-strain. Average grain sizes of different samples here range from 2 nm to 14 nm and increase with annealing temperature from 500 °C to 800 °C. Closer inspection reveals that we have a higher impact of post-annealing temperature than effective deposition temperature for the films. The high errors in grain size for as-deposited films are due to the nature of the low intensity, broad XRD peaks characteristic of amorphous films. At the highest annealing temperature of 800 °C, we also observe large error bars in the grain size, corresponding to fluctuations in the FWHM and intensity of different XRD peaks. We attribute this to the presence of texture which affects relative intensity, a feature that will be discussed in more detail below.
Furthermore, the sharpening and intensification of XRD peaks with annealing temperature can also be ascribed to an increase in crystallinity. It has been demonstrated that as-deposited spray pyrolysis films tend to be primarily amorphous with small, dispersed seed crystals and fully crystallize upon annealing.30,35,36,41 Extrapolating data from a Johnson–Mehl–Avrami–Kolmogorov (JMAK) analysis of spray-pyrolysis-deposited ceria films from a previous investigation35 reveals that annealing at 500 °C for 5 hours results in a fully crystalline film if seed crystals are present as-deposited. The presence of broad, weak diffraction peaks in the as-deposited films confirms that some crystallites are present, yet the overall low signal of these peaks implies a largely amorphous structure.
In addition, the XRD data reveals interesting trends in the texture, or preferred orientation, induced by the thermal treatment of sprayed ceria thin films. For the majority of films, the (111) diffraction peak shows the highest intensity, in agreement with the randomly oriented ceria powder reference. The preference for the 〈111〉 orientation is expected due to the high stability of the (111) plane.26,42,43 However, for samples deposited at substrate temperatures of 450 °C and 500 °C, a gradual shift of the preferred orientation from the 〈111〉 to the 〈200〉 orientation is observed with annealing. Remarkably, after annealing at 800 °C, the (200) peak has the highest intensity, meaning that annealing induced texture with a preference for (200) orientation in the plane parallel to the substrate. To better quantify this phenomenon, the texture coefficients of the (200) peak, presented in Table 1 and Fig. 5, were calculated using eqn (1):
T deposition/°C | T annealing/°C | Texture coefficient (200) | Lattice constant/nm |
---|---|---|---|
350 | As deposited | — | — |
500 | 1.247 | 0.541 ± 0.009 | |
650 | 1.295 | 0.541 ± 0.010 | |
800 | 1.481 | 0.539 ± 0.009 | |
400 | As deposited | — | — |
500 | 1.003 | 0.541 ± 0.032 | |
650 | 1.338 | 0.542 ± 0.008 | |
800 | 1.451 | 0.538 ± 0.030 | |
450 | As deposited | — | — |
500 | 1.489 | 0.543 ± 0.006 | |
650 | 1.771 | 0.541 ± 0.007 | |
800 | 2.198 | 0.540 ± 0.006 | |
500 | As deposited | — | — |
500 | 1.487 | 0.543 ± 0.010 | |
650 | 1.559 | 0.542 ± 0.005 | |
800 | 2.016 | 0.540 ± 0.007 |
Texture coefficient:
![]() | (1) |
Because as-deposited films are mostly amorphous with only a minimal number of small crystallites, the concept of texture coefficient is not applicable, and we exclude these films from the texture coefficient analysis. Turning to the texture coefficients for films post annealed at 500 °C, we observe that films deposited at 350 °C and 400 °C have values close to 1, revealing a preference for the 〈111〉 orientation as expected from the polycrystalline ceria reference data. However, films deposited at 450 °C and 500 °C also post annealed at 500 °C have texture coefficients of approximately 1.5, indicating a significant preference to the [200] direction relative to the reference even at this low annealing temperature. As the post-annealing temperature increases for all deposition temperatures, the texture coefficient increases, reaching a value of approximately 1.5 for films deposited at 350 °C and 400 °C and 2 for films deposited at 450 °C and 500 °C. Texture coefficients of 2 indicate that we obtain films with a remarkably high texture in the [200] direction at high deposition and annealing temperatures. The majority of films here have texture coefficients greater than 1, which suggests that there is an overall stronger presence of 〈200〉 oriented grains, compared to the polycrystalline ceria reference, that increases with temperature.
Such a thermally prompted shift in preferred orientation has previously been observed on ceria thin films deposited via spray pyrolysis29,30 pulsed laser deposition,31 and electron-beam-evaporation.33 Furthermore, this shift in preferred orientation appears to be independent of the substrate, as it was also observed for thin films deposited on MgO (see ESI†). This is in agreement with some reports in the literature, where such texture has been observed in ceria films deposited on sapphire substrates45 and on ceria nanoparticles deposited on Ru(0001).25 However, the reason for this prevalence of the (200) peak with increasing processing temperatures has been a point of speculation, as the (200) surface has been shown to be unstable due to its relatively high surface energy and net dipole moment arising from its arrangement of alternately charged planes.42 Several density functional studies have confirmed that the ordering of surface energies of the low index ceria planes from highest energy to lowest energy is: {100} > {110} > {111}.26,27 Consequently, the transition from 〈111〉 to 〈200〉 oriented films is expected to be energetically unfavorable and unstable; one would expect that upon annealing, films would relax to the most energetically favorable configuration. Therefore, our experimental results reported here indicate that the energetics calculated in these studies must be incomplete. Recent studies have identified the emergence of {100} terminated facets in ceria nanoparticles25,32 and 〈100〉 preferred orientations in thin films31,45 upon thermal treatment. It has been suggested that surface defects,32 such as oxygen ion vacancies,25 impurities, or surface reconstructions46 may facilitate the stabilization of {100} surfaces. In a computational study on {100} terminated ceria surfaces, it was determined using density functional theory that various reconstructions of the {100} surface could become energetically favorable at lower oxygen partial pressures by the formation of oxygen defects.25 Annealing films that were deposited at higher temperatures may provide the required energy needed to reconstruct or form oxygen vacancies at the surface. However, in the present study, annealing was conducted under stagnant air and not under reducing conditions as described in ref. 25. Therefore, the analysis only provides a partial explanation for how 〈200〉 oriented films could be stabilized and energetically favorable, as this study shows that a preferred (200) texture is induced at ambient oxygen partial pressures. Therefore, further theoretical studies are required to fully determine the energetics of surface orientation and texture in ceria and explain the preferred (200) texture observed here.
Understanding how to tune ceria processing such that the {100} facets are prevalent is appealing for catalyst applications and a point of active research. {100} surfaces have been demonstrated experimentally and computationally to have higher catalytic performance and activity28,47 than {111} surfaces as a result of the higher surface energy of {100} surfaces, which leads to more favorable oxygen vacancy formation energetics, and in turn a higher affinity to participate in oxygen-exchanging reactions. Therefore, the simple method of controlling film texture with a preferred 〈100〉 orientation presented here shows promise in controlling and engineering catalytic surface activity of ceria.
To complete the XRD analysis, we turn to the lattice parameter, which can be useful in detecting lattice strain. Table 1 shows the lattice parameter at each deposition and annealing temperature. Lattice parameters range from 0.539 nm to 0.542 nm which is close to the fully relaxed value 0.541 nm for ceria.48 Slight fluctuations in lattice parameters are seen, yet there is no clear trend with annealing or deposition temperatures. While this implies the presence of some degree of lattice strain, the lack of trend with temperature shows that it is likely not thermally induced.
Fig. 7 displays Raman spectra around the F2g mode for each deposition and annealing temperature, for further analysis of this characteristic Raman peak. For all spectra, except for the as-deposited film at 400 °C, the F2g peak is present. The overall peak intensity was weak due to the low thickness of the films, and for the film as-deposited at 400 °C, no F2g peak was observed, likely due to the weak signal of this thin film. Due to film thickness variations and instrument limitations on focusing quality during measurements, relative peak intensities fluctuated greatly, so we cannot draw any conclusions from peak intensity. As-deposited peaks are relatively broad, due to their amorphous microstructure. The lack of periodicity in an amorphous solid results in lattice vibrations that are more disordered and differ from the phonons of a crystalline solid. This impacts Raman scattering, causing broader Raman peaks with an asymmetric line shape.51 As annealing temperature increases, peaks narrow, indicating a transition from amorphous to crystalline microstructures52 and grain growth,53 agreeing with XRD results. Furthermore, changes in the F2g Raman mode position with annealing are observed. This is shown in Fig. 8 where the F2g mode position is plotted against annealing temperature and listed in Table 2. We classify two regions of F2g peak positions: Region I (shaded in gray) and Region II, being below 465 cm−1 and above 465 cm−1, respectively.
![]() | ||
Fig. 7 Raman spectra of the F2g peak of as-deposited and annealed films deposited at (a) 350 °C, (b) 400 °C, (c) 450 °C, (d) and 500 °C. |
T deposition/°C | T annealing/°C | F2g peak position/cm−1 |
---|---|---|
350 | As deposited | 450 |
500 | 462 | |
650 | 464 | |
800 | 466 | |
400 | As deposited | Peak not visible |
500 | 465 | |
650 | 469 | |
800 | 470 | |
450 | As deposited | 459 |
500 | 467 | |
650 | 467 | |
800 | 469 | |
500 | As deposited | 459 |
500 | 463 | |
650 | 467 | |
800 | 469 |
In Region I, F2g mode positions of as-deposited films are significantly below the characteristic bulk value of 465 cm−1, with a general trend of increasing F2g wavenumber with increasing post-anneal temperature. As established by XRD results, as-deposited films at all deposition temperatures exhibit a significant fraction of amorphous microstructure. Because annealing increases both grain size and crystallinity, a combination of these features prompts the general increase Raman mode frequency with annealing up to the bulk standard of 465 cm−1.
The observation that for all as-deposited samples where an F2g peak was present, the F2g peak position was significantly lower than the expected bulk value of 465 cm−1 is expected: The amorphous structure is generally more disordered and has longer average cerium–oxygen bonds than the crystalline phase, which results in a lowered F2g mode frequency for the largely amorphous as-deposited films. Additionally, the film as-deposited at 350 °C stands out with a notably low F2g wavenumber of 450 cm−1. While this low wavenumber can in part be attributed to its largely amorphous microstructure with a small number of dispersed nano-crystallites, it is also important to note that the film as-deposited at 350 °C was the only one to exhibit Raman-active organic precursor residues. The incorporation of organic precursor residues as space-filler in the structure of ceria films increases the cerium–oxygen bond distance, resulting in a lower F2g mode frequency. As the residues evaporate during annealing, space in the film where residues resided is freed, allowing for closer packing of the cerium and oxygen ions and consequently an increase in F2g mode frequency with annealing. Similar arguments have been made elsewhere that organic residues from the spray pyrolysis precursor can become incorporated into the film structure and result in a less compact structural packing.38
Four additional annealed films in Region I exhibit Raman peaks at or below the bulk peak position of 465 cm−1. As shown from the XRD analysis in the previous section, particle size within thin films here is in the range of a few nanometers and increases with annealing temperature. It is well-known that for nanocrystalline ceria the F2g peak is shifted to lower wave numbers compared to bulk ceria and that this shift increases with decreasing grain size.38,53–55 This phenomenon is due to phonon confinement: when grain-size is on the nanoscale, phonons become constrained within the small grains, resulting in Raman peaks at lower frequencies with more asymmetric broadening compared to in the bulk state.56–58 Therefore, given the particle size at the order of 5–10 nm for these films, phonon confinement can well explain the F2g peak shift below the bulk ceria reference and the increase in wavenumber of this shift with increasing annealing temperature.
While there is general agreement in literature regarding the trends observed in Region I, the results from Region II are more surprising. In Region II, the F2g mode similarly increases with annealing temperature. However, the F2g peak exceeds the 465 cm−1 bulk value, reaching values around 470 cm−1 for films deposited at 400 °C, 450 °C, and 500 °C and post-annealed at 800 °C. Such drastic changes in Raman shift and high F2g mode frequencies are often indicative of lattice strain, which is particularly prominent in thin films. In literature, changes in F2g mode frequencies on the scale shown here have been a result of chemical strain from doping,59 physically applied strain,60 and strain from the substrate15 in thin films. However, films here are undoped, no external stress is applied, and the deposition of amorphous ceria with subsequent crystallization indicates that the substrate should not influence the strain either. Turning to the lattice parameters obtained from XRD measurements, Table 1, the macroscopic strain in all thin films is small and without a clear trend, which is inconsistent with the clear increase in F2g frequency with annealing temperature. One would expect the lattice parameter to decrease with increasing Raman shift, as outlined by the Grüneisen parameter, which characterizes the correlation between lattice volume and vibrational frequency.13 Raman shift is directly related to bond strength, which increases with decreasing bond length. Therefore, a noticeable decrease in lattice volume would be expected for such a drastic increase in F2g frequency, which is however not observed in the long-range order as measured by XRD. In addition, we confirmed via XRD that there is no difference in the observed long-range order in thin films deposited on both the Si/Si3N4 substrates and the MgO substrates, as shown in the ESI (ESI 4†). In addition, it is highly unlikely that the present observations of induced texture and an apparent correlation with the Raman peak position are caused by trace impurity diffusion. First of all, the preferred orientation was observed for both the MgO and Si/Si3N4 substrates, i.e., the effect is independent of the underlying substrate. In particular, amorphous Si3N4 is an excellent diffusion barrier and highly thermally stable, and to the best of the authors' knowledge, there are no reports of significant Si diffusion from Si3N4 into ceria at the relatively low annealing temperatures of this study, between 350 °C and 800 °C. In addition, no extrinsic impurity peaks were observed under Raman, which would be expected for the solid solution formation or grain boundary diffusion of trace elements.17
A recent study on epitaxially grown ceria has observed that apparent peak shifts toward higher Raman wavenumbers in ceria may be caused by breaks of crystal symmetry, and that the apparent Raman shift is actually a peak splitting of the triply degenerate F2g mode.61 However, this peak splitting should be detectable under polarized Raman spectroscopy, and we did not find any effect of polarization on the measured Raman peak position in this present study.
Interestingly, we observe a correlation between the F2g peak position and the orientation of the grains, viz. the texture coefficient, as shown in Fig. 9. Both parameters, the F2g peak position and the texture coefficient, increase with increasing annealing temperature. We observe that for samples annealed at 500 °C, the F2g peak position is close to the bulk value of 465 cm−1, and the (200) texture coefficient is close to 1. Therefore, samples annealed at 500 °C exhibit structural properties in agreement with the literature of equilibrated, stoichiometric cerium oxide.17 In contrast, for samples annealed at 650 °C and 800 °C, we observe significant texture in favor of the 〈100〉 orientation, as well as high Raman F2g wavenumbers between 465 cm−1 and 470 cm−1. This suggests that the increase in texture for the 〈100〉 orientation with annealing may cause a decrease in the cerium–oxygen bond length at the near order within the thin film surface or volume. Similar to the work by Sediva et al.,61 it can be hypothesized that a symmetry break, in this case in the form of induced texture, induces a Raman peak shift toward higher wave numbers which cannot be explained by other effects such as phonon confinement or local defects. This is a first indication that Raman spectroscopy could be used to measure thin film texture in ceria.
We newly report crystal texture data across a wide range of deposition and post-annealing temperatures, and find that a preferred orientation detected in XRD correlates to unusual peak shifts of the Raman F2g oxygen breathing mode to higher wavenumbers. We find that the energetically unfavorable, but highly catalytically active 〈100〉 orientation becomes preferred after thermal annealing, and that Raman spectroscopy can be a useful, accessible tool to detect such orientation evolution for catalysis research.
Collectively, this contributes to the fundamentals understanding of ceramic synthesis by spray pyrolysis, and specifically on the material class of ceria. In addition, it provides a cost-effective method to vary future catalyst coatings of ceria with orientations to control their functionality. For this we identify a new correlation between the F2g Raman stretching mode and crystal orientation that can be used as a marker for future designs of thin film deposition pathways and catalytic coatings.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1ta09115h |
‡ The authors contributed equally to this manuscript. |
This journal is © The Royal Society of Chemistry 2022 |