Next Article in Journal
Enhanced Performance of GaAs Metal-Oxide-Semiconductor Capacitors Using a TaON/GeON Dual Interlayer
Previous Article in Journal
The Role of Nanofluids in Renewable Energy Engineering
Previous Article in Special Issue
Nanoparticulate Perovskites for Photocatalytic Water Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Direct Evidence of Dynamic Metal Support Interactions in Co/TiO2 Catalysts by Near-Ambient Pressure X-ray Photoelectron Spectroscopy

1
European Synchrotron Radiation Facility, CS 40220, CEDEX 9, 38043 Grenoble, France
2
Centre RAPSODEE UMR CNRS 5302, IMT Mines Albi, Université de Toulouse, Campus Jarlard, CEDEX 09, 81013 Albi, France
3
Laboratoire de Physique et Chimie des Nano-Objets (LPCNO), Université de Toulouse, INSA, UPS, CNRS, LPCNO, 135 Avenue de Rangueil, 31077 Toulouse, France
4
LCC, CNRS-UPR 8241, ENSIACET, Université de Toulouse, 31030 Toulouse, France
5
Interface Design, Helmholtz-Zentrum Berlin für Materialien und Energie GmbH (HZB), Albert-Einstein-Str. 15, 12489 Berlin, Germany
6
Energy Materials In-Situ Laboratory Berlin (EMIL), Helmholtz-Zentrum Berlin für Materialien und Energie GmbH (HZB), Albert-Einstein-Str. 15, 12489 Berlin, Germany
7
Institut de Chimie et Procédés Pour l’Energie, l’Environnement et la Santé (ICPEES), ECPM, UMR 7515 CNRS—Université de Strasbourg, 25 Rue Becquerel, CEDEX 02, 67087 Strasbourg, France
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(19), 2672; https://doi.org/10.3390/nano13192672
Submission received: 7 September 2023 / Revised: 24 September 2023 / Accepted: 25 September 2023 / Published: 29 September 2023

Abstract

:
The interaction between metal particles and the oxide support, the so-called metal–support interaction, plays a critical role in the performance of heterogenous catalysts. Probing the dynamic evolution of these interactions under reactive gas atmospheres is crucial to comprehending the structure–performance relationship and eventually designing new catalysts with enhanced properties. Cobalt supported on TiO2 (Co/TiO2) is an industrially relevant catalyst applied in Fischer−Tropsch synthesis. Although it is widely acknowledged that Co/TiO2 is restructured during the reaction process, little is known about the impact of the specific gas phase environment at the material’s surface. The combination of soft and hard X-ray photoemission spectroscopies are used to investigate in situ Co particles supported on pure and NaBH4-modified TiO2 under H2, O2, and CO2:H2 gas atmospheres. The combination of soft and hard X-ray photoemission methods, which allows for simultaneous probing of the chemical composition of surface and subsurface layers, is one of the study’s unique features. It is shown that under H2, cobalt particles are encapsulated below a stoichiometric TiO2 layer. This arrangement is preserved under CO2 hydrogenation conditions (i.e., CO2:H2), but changes rapidly upon exposure to O2. The pretreatment of the TiO2 support with NaBH4 affects the surface mobility and prevents TiO2 spillover onto Co particles.

1. Introduction

Oxide-supported metal catalysts play a significant role in a wide range of industrial chemical processes, while they are crucial to the production of sustainable and clean energy from renewable resources [1,2]. The interaction between the metal and the support is critical for the catalytic performance since it controls important characteristics of the catalyst, such as the dispersion, electronic structure, and stability of the active phase [3]. Although the strength of the metal–support interaction is influenced by many different factors, reducible oxides are generally acknowledged to promote stronger interactions with the metal than the non-reducible ones [3,4,5].
Pt supported on TiO2 is the archetype of the so-called strong metal–support interaction (SMSI) systems, although other Pt-group metals on TiO2 have also been widely investigated [3,4,5]. It is now well established that metal–support interactions are dynamic and their strength may significantly vary depending on the gas environment, ranging from simple charge transfer to more extensive mass transport and surface restructuring [6]. Therefore, a fundamental understanding of SMSI phenomena is of paramount interest for optimizing catalytic performance and designing innovative catalytic systems.
Although the SMSI effect was initially described almost 50 years ago [7], our understanding of it is far from comprehensive, and it remains one of the most researched areas in heterogeneous catalysis [3,4,5,8]. The recent breakthrough developments of advanced material characterization techniques, such as environmental high-resolution transmission electron microscopy (HRTEM), allowed us to visualize the dynamic interplay at the metal–oxide interface during operation, which was not feasible in ex situ and postmortem studies [9,10,11]. HRTEM studies can observe in real time the encapsulation of metal clusters by a thin layer of reduced titanium oxide under reducing conditions, and the formation of a thicker titania overlayer upon subsequent oxidative treatment [9,10,11].
Although less studied than Pt-group metals, transition metals, such as Ni and Co supported on TiO2, are known to undergo restructuring processes under reaction conditions [10]. In particular, encapsulation of Co nanoparticles under a thin reduced titania layer has been observed after H2 treatment of Co/TiO2 catalysts [12,13,14]. Deeper understanding of these interactions is essential in order to improve the performance of such catalysts in industrial applications. For example, Co/TiO2 is one of the most industrially applied Fischer−Tropsch synthesis (FTS) catalysts [2,15]. FTS is used to synthesize diesel from biomass feedstocks, among other products. As for Pt/TiO2 catalysts, microscopy techniques are primarily used in order to visualize changes in the Co-TiO2 interface in response to the gas phase environment [12,13,16]. However, SMSI phenomena over Co/TiO2 catalysts have been mostly noticed by ex situ experiments in which the sample was transferred to the microscope after reduction/oxidation gas treatments. In addition, due to the low contrast difference between Ti and Co atoms, in situ HRTEM studies are more challenging than those over Pt/TiO2.
Instead, X-ray photoelectron spectroscopy (XPS), due to its chemical specificity and surface sensitivity, is a powerful tool to study chemical and structural transitions of Co/TiO2 upon interaction with gas atmospheres. In addition, photoemission experiments are quantitative and give an average information over the entire sampling geometric area (around 100 µm [17]), providing a quite representative image of the overall sample structure. The recent development of XPS spectrometers capable of operating at pressures of a few tens of millibars [18,19,20], referred to as near-ambient pressure XPS (NAP-XPS), provides new opportunities to study SMSI process in reaction environments. However, only a few NAP-XPS studies exist on SMSI effects in TiO2-supported catalysts [21,22,23] and to the best of our knowledge, none concerning the Co/TiO2 system.
Herein, we present a synchrotron-based near-ambient pressure X-ray photoelectron and absorption study of cobalt catalysts supported on pure and NaBH4-modified TiO2 [24], in H2, CO2:H2, and O2 gas atmospheres. This work is unique in the field of SMSI investigations in that soft and hard X-rays are coupled to extend the analysis depth of the photoemission process by an order of magnitude (from about 2.5 to 25 nm). The results show a dynamic response of the catalysts to the gas phase environment and reveal critical differences between pure and NaBH4-modified Co/TiO2 catalysts.

2. Materials and Methods

2.1. Catalysts Preparation

Initially, commercial TiO2-P25 (Evonik) was partially reduced using NaBH4 as the reducing agent, leading to the incorporation of Na and B promoters. More specifically, known amounts of NaBH4 were dissolved in ethanol in a rotavapor flask and then TiO2-P25 powder was added to the solution. The resulting mixture was stirred for 1 h (25 °C, 500 mbar) and subsequently, the ethanol was removed using a rotavapor (80 °C, 150 mbar for 30 min and then 75 mbar for 30 min). Next, the sample was dried overnight at 120 °C in a static oven and then treated in a tubular oven under argon at 320–370 °C for 15 min. Afterwards, the resulting blue product was recovered under Ar and then washed three times with distilled water, followed by washing with absolute ethanol. Finally, the product was dried for 15 h at 150 °C under argon. Textural, chemical, and crystallite properties of the TiO2 and m-TiO2 supports are given in Table S1.
Subsequently, modified and unmodified TiO2 were used to prepare 10 wt% Co-based catalysts by conventional incipient wetness impregnation. To do so, the TiO2 supports were placed in a Schlenk flask and degassed for 2 h under vacuum at 150 °C (oil bath temperature). After cooling down at room temperature, an aqueous metal precursor solution (Co(NO3)2·6H2O) was added under vacuum and continuous stirring. The resulting mixture was sonicated for 30 min, followed by 30 min of stirring. The sonication/stirring sequence was repeated 4 times. The mixture was dried for 24 h at 80 °C followed by 12 h at 120 °C in a static furnace. Finally, the powder was calcined for 4 h at 460 °C under Ar. Hereafter, the unmodified and modified catalysts will be referred to as Co/TiO2 and Co/m-TiO2, respectively. The bulk atomic concentration (at%) of the various catalysts’ components, as calculated by inductively coupled plasma optical emission spectroscopy (ICP-OES) measurements, was 9.6 at% Co for Co/TiO2 and 8.0 at% Co, 1.4 at% Na, and 0.2 at% B for Co/m-TiO2. Because of the low contrast difference between Ti and Co, electron microscopy could not be used to assess the size and distribution of Co particles. Instead, the cobalt crystallite size was determined from XRD measurements using the Scherrer equation. The quantification of the metallic Co crystallite size of the reduced catalysts was not possible due to the overlap of various crystallographic phases. In the case of calcined catalysts, the crystallite size of Co3O4 was determined to be around 37 nm and 44 nm for Co/TiO2 and Co/m-TiO2, respectively.

2.2. Spectroscopic Measurements

Near-ambient pressure X-ray photoelectron spectroscopy (NAP-XPS) and near-ambient pressure hard X-ray photoelectron spectroscopy (NAP-HAXPES) combined with near edge X-ray absorption fine structure spectroscopy (NEXAFS), were applied to probe the surface state of the pre-reduced Co/TiO2 and Co/m-TiO2 catalysts. Experiments were performed at the new CAT branch of the dual-colour Energy Materials In-Situ Laboratory (EMIL) beamline (CAT@EMIL) at the synchrotron radiation facility BESSY II of the Helmholtz Zentrum, Berlin [17,25]. A specificity of the CAT@EMIL beamline is the possibility to use soft and hard X-ray radiation (in our case, 4.9 keV, thus in the energy border between tender and hard X-rays) provided by two undulators, UE48 and CPMU17,monochromatized by plane-grating monochromator (PGM) and a double crystal monochromator (DCM), respectively, and focused to the sample position by means of a set of optical mirrors in a single experiment. The variability of the incident photon energy permits non-destructive depth profiling from the extreme surface to the subsurface. The experimental station hosts SPECS PHOIBOS 150 NAP analyzer with a 2D-CMOS detector.
The two catalysts were initially reduced ex situ at 350 °C in 1 bar 40% H2/Ar flow for 4 h. The reduced powder was pressed into pellets and mounted on a sapphire sample holder between a stainless-steel back-plate and a lid with 5 mm hole. To guarantee the repeatability of the treatment conditions, the two catalysts were put together, in close proximity, on the sample holder (Figure S1). The sample stage was heated from the rear by an IR laser. Gases were introduced into the analysis chamber via calibrated mass flow controllers (Bronkhorst). The gas phase composition was monitored by a differentially pumped quadrupole mass spectrometer (QMS, Pfeiffer PrismaPro) connected to the sample chamber through a leak valve. The NAP-XPS spectra were measured with a PGM (Bestec GmbH, Berlin, Germany) and a 60 um exit slit. The NAP-HAXPES was measured with a DCM (Bestec GmbH, Berlin, Germany) using Si (111) crystal pair.
The two catalysts were first annealed in front of the spectrometer nozzle for 30 min at 350 °C under 5 mbar H2 before being cooled to 200 °C to record the spectra in H2. Then, the gas atmosphere was changed to a CO2:H2 (1:2) mixture with an overall pressure of 2.5 mbar and the temperature was raised to 250 °C where a new set of spectroscopic measurements was recorded. Finally, the CO2:H2 mixture was replaced by 1 mbar O2 and the temperature further rose to 350 °C in order to obtain the characteristic spectra of the oxidized catalysts. The NAP-XPS, NAP-HAXPES, and NEXAFS spectroscopic measurements were carried out consecutively on each sample while maintaining stable gas and temperature conditions.
Quantification of the elements was performed after normalization of the XPS spectra intensities by considering the photon flux and the atomic subshell photoionization cross sections. The photoionization cross sections and the inelastic mean free paths (IMFPs) of photoelectrons were obtained by the SESSA (Simulation of Electron Spectra for Surface Analysis) software (version 2.2) [26]. The incident photon flux was measured based on the sample drain current of a cleaned Au foil. The binding energies (BE) of the photoemission peaks presented here are referred to the Fermi edge cut-off position, measured using the same photon energy with the core-level spectrum measured just before. Unless otherwise stated, the accuracy of BE calibration was estimated to be ±0.1 eV. The XPS spectra were analyzed using CasaXPS software (Casa Software Ltd., version 2.3.25). After linear or Shirley background subtraction, the B 1s, Na 1s, and O 1s spectra were fitted by symmetric Gaussian–Lorentzian line shapes, while constrains on the width and the relative BE position of the fitting components were applied.
The excitation photon energy of the NAP-XPS core-level spectra was selected in such a way that the emitted photoelectrons have comparable kinetic energies (KE). Two sets of NAP-XPS spectra with two different KE (about 180 and 450 eV) were measured, corresponding to two analysis depths (a.d.) of about 2.5 and 5 nm, respectively (the information depth is considered to be 3 times the IMFP) [27]. For all NAP-HAXPES experiments discussed in this work, the photon energy was fixed to 4900 eV (information depth about 25  nm).
The Co L3,2- and Ti L3,2-edge NEXAFS spectra were measured in the total electron yield (TEY) mode, using a Faraday cup installed in the first aperture of the analyzer electrostatic lenses. In general, the analysis depth of NEXAFS measurements in the TEY mode ranged between 5 and 10 nm [28].

3. Results and Discussion

3.1. Chemical State Measured in H2 and CO2:H2 Gas Atmospheres

Figure 1 shows the Co and Ti 2p core level and valence band spectra of the two catalysts measured at 200 °C in 5 mbar H2 and at 250 °C in a 2.5 mbar CO2:H2 mixture. The peak shape and the binding energy of the Co 2p3/2 peak at 778 ± 0.1 eV (Figure 1a) is characteristic of metallic Co [29,30]. The two catalysts have almost identical Co 2p3/2 spectra, suggesting that the modification of TiO2 support by NaBH4 does not notably affect the cobalt chemical state. In addition, Co 2p3/2 seems unaffected after switching the gas atmosphere to the CO2:H2 mixture, as validated by the NEXAFS data (vide infra). The Ti 2p spectra (Figure 1b) are composed of two peaks at 459.1 eV and 464.8 eV, which are ascribed to Ti 2p3/2 and Ti 2p1/2 in TiO2, respectively. The Ti 2p peaks of both samples are identical in H2 and CO2:H2 atmospheres and correspond to the Ti4+ ions of stoichiometric TiO2 [22]. It is interesting that there are no evident peak features at the low BE side of the Ti 2p that could be ascribed to Ti3+ formation [22]. This shows that TiO2 is not reduced under the present hydrogen treatment, a result which is consistent with our NEXAFS Ti L-edge spectra (vide infra).
The valence band (VB) is dominated by the Ti 3d and O 2p states between 10 eV and 3 eV [31] and the characteristic Co 3d sharp cut-off at the Fermi level [32]. The relative height of the two distinct features at 5.2 and 7.2 eV is sensitive to the type of TiO2 polymorph (a-TiO2 (anatase) or r-TiO2 (rutile)) [31]. The VB spectra of the Co/TiO2 and Co/m-TiO2 catalysts look alike, confirming the core-level spectra, which indicated that the two catalysts have identical Co and Ti oxidation states. In addition to that, the stability of the two features corresponding to Ti 3d and O 2p states indicates that there is no phase transition of TiO2 under the reaction conditions employed.
The presence of B 1s and Na 1s photoemission peaks (Figure 2) indicates that Na and B remain on Co/m-TiO2 sample surface after the NaBH4 treatment. Previous reports have shown that the BE of the B 1s is sensitive to the boron local chemical environment [33,34,35,36]. In particular, the B 1s peak of anionic B2− (i.e., TiB2 or CoB) and cationic B3+ (i.e., B2O3) appears around 187.5 eV and 193 eV, respectively, while B occupying substitutional or interstitial TiO2 sites is found between 190 and 192 eV [33,34,35,36]. Under H2, the BE of B 1s peak was measured at 192.3 ± 0.1 eV (Figure 2a), which suggests that B most probably forms an oxide, rather than being integrated in the TiO2 structure by substitution of TiO2 sites. Moreover, the BE at 192.3 eV is significantly lower than that reported for the common B2O3 oxide, indicating that boron is partially reduced forming a substoichiometric oxide (e.g., boron suboxide, B6O). The low width (1.35 eV) and the symmetry of the peak shape under H2 is a sign that this is the unique boron oxidation state. Under CO2:H2, the B 1s peak becomes broader and shows an asymmetry at the high BE side (Figure 2b). Curve fitting of the B 1s suggests the presence of an additional B 1s peak at 193.3 eV, which is compatible with oxidation of B-suboxide to B2O3. This indicates the affinity of B towards the CO2 present in the reaction mixture, which is considered a mild oxidant. One should note here that the QMS results showed negligible H2O production under the low-pressure conditions employed suggesting that CO2 is the only oxidant present in the gas phase.
The Na 1s spectra in H2 (Figure 2c) show a single symmetric peak at 1072.5 eV, which is characteristic of Na2O and/or ionic Na species bound to the surrounding support through –O–Na linkages [37,38]. For the sake of convenience, the two species mentioned above are abbreviated as NaOx. The formation of sodium carbonate or hydroxide species is unlikely, since in that case, the Na 1s would appear at a considerably lower BE [39]. When the catalyst is exposed to CO2:H2 (Figure 2d), an evident shoulder appears at the low BE side of the Na 1s peak. The Na 1s curve fitting shows the presence of an additional Na 1s component shifted by 1 eV towards a lower BE (1071.5 eV). This peak has been previously assigned to sodium titanate species [40,41]. The depth-dependent Na 1s spectra included in Figure 2d clearly show surface enrichment of the 1071.5 eV peak as compared to the one at 1072.5 eV. This signifies that sodium titanates are formed on top of NaOx species.
The O 1s and C 1s spectra are presented in Figure 3. The peak at 530.3 ± 0.1 eV, due to lattice TiO2 species [42], dominates the O 1s spectra of the two catalysts (Figure 3a). No discernible differences can be observed in the O 1s peaks of the two catalysts, thereby hindering the ability to differentiate between the contributions of B and Na oxides in the overall peak. Nevertheless, curve fitting of the O 1s peak allows us to distinguish a small O 1s component at 531.6 ± 0.2 eV, typically connected to surface C=O and/or surface adsorbed −OH groups [42]. Since the contribution of C=O is minor in the C 1s peak (vide infra), the component at 531.6 eV should be mainly attributed to −OH species. This is confirmed by depth-dependent O 1s data (Figure S2), which demonstrate a decrease in the 531.6 eV component at deeper analysis depths. In the case of the Co/TiO2 sample, the fraction of the OH-peak increases and shifts to higher BEs in CO2:H2 suggesting a higher abundance of oxygenated species in this case.
The C 1s peak at 284.6 ± 0.1 eV (Figure 3b) is due to C-C and C-H carbon bonds of adventitious hydrocarbons, as could be anticipated for samples that have previously been exposed to air. Notably, the amount of carbon species is more important for Co/TiO2 than Co/m-TiO2 suggesting that the NaBH4 treatment influences the reactivity of TiO2 towards carbon. The C 1s spectra in H2 and the CO2:H2 mixture appear very similar, apart from a decrease in the peak intensity under reaction conditions, which is translated to less adsorbed carbon. A small feature at the high-BE side of the C 1s peak in CO2:H2 can be attributed to the formation of C-O and/or hydroxy species. However, the absence of a peak at around 290 eV can safely exclude the formation of carbonates [43].
The Co and Ti L-edge X-ray absorption spectra (Figure 4) provide fine details about the electronic and geometric structure of Co and Ti. The Co L-edge (Figure 4a) is composed of two peaks (i.e., Co L3- and L2-edges) due to the spin–orbit coupling of the Co 2p states. The sharp rising edge of Co L3 with the intense maxima at 778.7 eV, as well as the absence of fine structure features at the high photon energy side of the peak, are typical characteristics of metallic Co states [22,30]. Notably, the Co L-edge of the two catalysts is similar and appears to be unaffected by the gas phase conditions, confirming the NAP-XPS results.
The Ti L-edge spectra are included in Figure 4b. The edge splits into two peaks due to the spin–orbit coupling of the Ti 2p states, similar to the Co L-edge. Note that the peak intensity ratio between the two peaks around 460 eV (indicated by the arrows in the figure) is sensitive to the TiO2 crystal symmetry. In particular, the lower photon energy spectral feature is considerably greater in height for a-TiO2 compared to r-TiO2 [22,44]. The Ti L-edge absorption profiles in Figure 4b, including the peak features around 460 eV, correspond to previously reported spectra of a-TiO2 [22,44]. Notably, all the Ti L-edges are comparable in terms of spectral line shape and peak excitation energies, suggesting that the valence state and phase of Ti are not affected by the gas atmosphere, which is consistent with the NAP-XPS results (Figure 1). Please note that, while a-TiO2 is obviously the dominant surface state based on VB photoemission and Ti L-edge absorption spectra, the existence of minor quantities of r-TiO2 cannot be ruled out since their spectroscopic features are obscured by the strong signal of the dominant a-TiO2 phase.
In summary, the analysis of the spectroscopic results reveals that Co stays metallic and TiO2 is completely oxidized under the employed H2 and CO2:H2 conditions. The comparability of spectroscopic data from Co/TiO2 and Co/m-TiO2 samples indicates that modification of TiO2 by Na and B does not influence the surface chemical state or the electronic structure of Co/TiO2 catalyst. Although the CO2:H2 mixture induces further boron oxidization and promotes sodium titanate formation compared to their state in H2, it has no effect on Co or TiO2. These findings are more consistent with a static surface configuration than the well-documented dynamic evolution of SMSI systems. This can be justified by the chemical potential of the gas phase, which is highly reducing in both H2 and CO2:H2 atmospheres. Therefore, below, we investigate the surface dynamics under oxidative conditions by changing the gas environment to O2.

3.2. Surface Transformation upon Exposure to O2

The surface transformation of the reduced samples in an O2 atmosphere is examined next. Figure 5 shows the NAP-XPS and NEXAFS spectra of Co/m-TiO2 and Co/TiO2 recorded in 1 mbar O2 at 350 °C. The Co 2p photoemission peaks (Figure 5) become broader and shift to higher BEs in O2 compared to the previously recorded spectra in CO2:H2. This, together with the presence of the low intensity satellite feature around 790 eV, suggests complete oxidation to Co3O4 [22,29,30]. This is supported by the NEXAFS Co L-edge in Figure 5 (top, right), which is typical of a bulk Co3O4 spinel structure [22,29,30]. The Ti 2p and Ti L-edge spectra look identical to those recorded in CO2:H2, which indicates that the a-TiO2 phase is preserved in O2. The analysis of the B 1s and Na 1s peaks reveals that the proportion of the components at 193.3 eV and 1071.5 eV is enhanced in O2 compared to CO2:H2. (see Figure 2). This supports the hypothesis that the presence of oxidants in the gas environment, including mild oxidants like CO2 (Figure 2), promotes the synthesis of Na-titanate and B2O3.
The spectra of the valence band region confirm the interpretation of the core level peaks. More specifically, the distinct feature appearing at around 2 eV is due to Co 3d states [45], which upon cobalt oxidation are shifted 2 eV below the Fermi edge. On the contrary the features at 10 eV–3 eV region, corresponding primarily to Ti 3d-O 2p states, are identical to those found in H2 and H2:CO2 (Figure 1) confirming the NEXAFS findings. The O 1s spectra of the two samples become significantly broader in O2 compared to the previous reduced state (the FWHM of the O 1s peak increases from around 1.3 to 1.8 eV) and shifts by about 0.2 eV to lower BEs. These changes can be understood by the appearance of a new O 1s component due to Co3O4; however, the absence of clear spectral features in the O 1s peak makes O 1s curve fitting ambiguous, and therefore will not be attempted in this case.

3.3. Effect of the Gas Atmosphere on the Surface Composition

Due to its high surface sensitivity, photoemission spectroscopy is a powerful tool to quantify the atomic concentration at the very surface. Figure 6 compares the surface atomic composition of the Co/m-TiO2 and Co/TiO2 catalysts in reducing, reaction, and oxidizing gas environments calculated from the NAP-XPS data. The bulk concentration based on the ICP-OES analysis is included for comparison. In the NAP-XPS calculations, it is assumed that the sample is homogeneous over the analyzed depth. Noteworthily, for layered samples (as will be discussed later) this approximation overestimates the concentration from the top layer relative to that of the layers below. Therefore, one should be cautious when comparing the absolute values of the NAP-XPS and ICP-OES results. Nevertheless, the quantitative NAP-XPS analysis can be used for comparisons to reveal modifications of the surface composition as a function of the gas environment.
In H2 and CO2:H2 atmospheres, the Co surface concentration of Co/TiO2 (Figure 6a) is about half of that found by ICP-OES, indicating that the surface is enriched with TiO2. However, when the gas environment is changed to O2, the Co concentration increases significantly, greatly beyond that of ICP-OES. This reflects a dynamic response of the surface composition to the gas atmosphere, with TiO2 segregating over Co in reducing conditions, and Co re-emerging back on the surface in an oxidizing atmosphere.
Significant differences between bulk and surface compositions were also found for Co/m-TiO2 (Figure 6b). In particular, in H2 and CO2:H2, the surface contains about 3 times less Co and almost 25 times more Na and B compared to the bulk concentration of these elements. This is a sound evidence that Na and B are the dominant elements on the sample surface. The surface concentration of Co, Ti, and B increase in O2 compared to the previous state in CO2:H2, at the expense of Na. As shown in Figure 5, this is accompanied by enhancement of the component related to Na-titanate species. Thus, when exposed to oxidizing conditions, a portion of the TiO2 support interacts with NaOx at the surface to create Na-titanates. This is not necessarily a solid-state reaction, but might simply be due to sodium migration within the TiO2 lattice. As was showed earlier, Na+ ions may be introduced into the TiO2 host without damaging the anatase structure or altering the TiO2 oxidation state [46].

3.4. Depth Distribution of the Catalyst Components

By taking advantage of the tunability of synchrotron radiation, one can vary the kinetic energy of photoelectrons and in turn the information depth of the measurements. This approach, usually referred to as depth-profiling, allows the quantification of the composition at various depths, which is particularly useful in cases of layered surface morphologies as is obviously the case here. The combination of NAP-XPS and NAP-HAXPES enables the excitation photon energy to be varied between 370 and 4900 eV, thereby extending the sample depth from 2.5 to 25 nm. More details about the relationship between photon energy and analysis depth can be found in the Supplementary Material.
Figure 7a shows the evolution of Co and Ti %at for Co/TiO2, as a function of the analysis depth measured at 200 °C in 5 mbar H2. The Ti atomic fraction is clearly enhanced at the surface (up to 5 nm) compared to the subsurface measurement (25 nm), which can be associated with the encapsulation of Co under a thin TiO2 overlayer (shown in the graphical illustration in Figure 8). This finding is consistent with the low surface fraction of Co compared to the bulk concentration shown in Figure 6a. This is a classical manifestation of SMSIs between Co and TiO2, where the support, in a partially reduced state, migrates onto Co particles during the reduction process [12]. Despite the fact that the SMSIs are frequently accompanied by partially reduced TiO2 species (e.g., Ti3+) [12,22], in our case, the Ti 2p and Ti L-edge spectra did not show any evidence of reduced TiO2 for all analysis depths. This implies that this configuration existed prior to the NAP-XPS studies, most likely happening during the ex situ reduction pretreatment at 1 bar (see Section 2.1), which is much more aggressive than the reducing conditions in the NAP-XPS chamber. Under the CO2:H2 reaction conditions, the depth profile measurements are qualitatively similar to those presented above under H2, suggesting no significant surface reconstruction during reaction, in agreement with the findings of Figure 6a.
In the O2 atmosphere (Figure 7b), the trend between the two elements is reversed compared to H2, with the Co fraction strongly increased near the surface (2.5 nm) but stabilized when deeper layers are probed. This is a clear indication that the thin TiO2 overlayer on Co, which manifested in H2 and CO2:H2 atmospheres, retreats under oxidative conditions and cobalt is exposed on the surface (Figure 8). The reversibility of the TiO2 spillover effect, where Co is oxidized and segregates above TiO2, has been suggested in early SMSI literature [9]. However, to our knowledge this is one of the rare cases where the reversibility of the SMSI effect is evidenced in situ with a method other than microscopy.
Figure 7c shows the depth distribution of Co, Ti, Na, and B for the Co/m-TiO2 sample under H2. The enhanced concentration of Na and B at the more surface-sensitive measurements (i.e., 2.5 and 5 nm) implies that these elements are not equally distributed within the catalyst volume, but preferentially situated on its surface, above Ti and Co (Figure 8). This interpretation agrees with the high surface concentration found for Na and B as compared to the bulk concentration, as shown in Figure 6b.
When the Co/m-TiO2 sample is annealed in O2 (Figure 7d), the Co and Ti evolution with analysis depth remains qualitatively similar to that observed in H2, meaning that B and Na are still dominating the extreme surface. However, a comparison of the Co %at between H2 and O2 atmospheres reveals that there is more Co under O2 for all analysis depths, comparable to the finding for Co/TiO2. The rise in Co percentage under O2 should be linked to the decrease in Na %at. This may be caused by either a drop in the thickness of the Na layer over the Co3O4 particles or its full disappearance. The latter scenario is rather improbable due to the fact that depth profiling measurements (Figure 7d) clearly show that it is Na %at which is enhanced at the outer surface compared to the subsurface, and not Co. One can assume that boron oxides are also preferentially located at the extreme surface, since B %at has a similar evolution with Na as a function of the analysis depth (Figure 7d). Therefore, as shown in Figure 8, we propose that the Na and B surface overlayer formed under reducing conditions is maintained under O2 but decreases in thickness. One should mention here that depth profile measurements cannot distinguish if the overlayer above cobalt is dense or has a certain porosity that would allow access to gases during reaction. The two morphologies (i.e., porous or dense Na and B surface overlayer) are expected to lead to radical differences in the catalytic performance. In particular, a dense surface layer will block the access of the reactants to cobalt, thus decrease the number of active sites and consequently the catalytic activity. On the contrary, in case of a porous overlayer, not only cobalt sites remain accessible to the reactants, but the additives can also act as promoters, enhancing the catalytic turnover.
The aforementioned analysis suggests that the modification of the TiO2 support by Na and B does not affect the chemical state of the cobalt catalyst, but plays a significant role in the arrangement between Co and TiO2 by preventing TiO2 migration over cobalt. At this stage, it is not clear if the lower surface mobility on m-TiO2 is due to the lower reducibility of this support. Under the reduction conditions employed in the NAP-XPS experiment, TiO2 was quite stable for both samples. In addition, H2-TPR measurements were not conclusive about the effect of Na and B in the reducibility of the TiO2. In any case, XPS is the method of choice here since H2-TPR lacks the sensitivity to detect surface reduction of TiO2, which is critical to identifying SMSIs.
Overall, the goal of this study was to examine the evolution of Co/TiO2 catalysts in reactive gas atmospheres while also contributing to our understanding of metal–support interaction mechanisms. The observed surface restructuring accompanied by changes in cobalt phase, seen while switching between reducing and oxidative gas atmospheres, are likely to alter the catalyst’s reactivity. According to our findings, a thin TiO2 layer is anticipated to develop on top of the cobalt during H2 activation of calcined catalysts. This layer will affect the cobalt particle growth and the accessibility of the reactants over cobalt sites. However, whether the geometric and electronic structure of pre-reduced Co/TiO2 catalysts is maintained throughout reaction conditions, or the catalyst undergoes additional alterations, remains open. The measurements in the CO2-FTS-relevant conditions (i.e., CO2:H2 mixture at 250 °C) presented in this paper, suggest that the cobalt chemical state and atomic concentration remain rather stable. In contrast, we found a noticeable change in the chemical state of B and Na in CO2-FTS conditions for Co supported on NaBH4-modified TiO2. In particular, the presence of CO2 in the gas phase oxidizes boron and triggers an interaction at the Na-TiO2 interface. This trend implies that under realistic high-pressure FTS reaction conditions, B2O3 and Na-titanates may be the dominating surface species over the Co/m-TiO2 catalyst. Furthermore, our findings imply that when cobalt is oxidized, for example by H2O production under FTS conditions, unmodified catalysts experience considerable surface restructuring, which is not the case with Co/m-TiO2. Although these findings provide important insights into dynamic surface modifications, it is impossible to predict how the catalytic performance will be modified based solely on these data. Before attempting to connect surface states with catalytic performance, further information about the active phase features, such as nanoparticle size, shape, and exposed facets, is necessary.

4. Conclusions

In summary, in this work, we combined NAP-XPS, NAP-HAXPES, and NEXAFS spectroscopies to explore the in situ interaction of two Co/TiO2 catalysts with H2, CO2:H2, and O2 atmospheres. We found that cobalt is reduced to the metallic state under the employed H2 and CO2:H2 conditions, while it is readily transformed to spinel Co3O4 upon switching to O2. On the contrary, the titania chemical state and structure are not affected by the gas treatment with the a-TiO2 phase dominating the surface. The treatment of TiO2 with NaBH4 left significant amounts of Na and B residuals on the surface even after Co deposition in the aqueous metal precursor solution. The restructuring of the Co-TiO2 interface under operating conditions was directly observed by quantitative analysis of photoemission results combined with depth-profiling measurements. It was shown that, under reduction conditions, cobalt particles are encapsulated below a stoichiometric TiO2 layer. This arrangement seems to be preserved under the employed CO2 hydrogenation conditions, but rapidly changes upon exposure to O2. The pretreatment of the TiO2 support with NaBH4 affects the surface mobility of TiO2 and, to a large extent, prevents spillover onto cobalt. However, in this case, cobalt is covered under a Na and B layer, probably formed upon reduction. To a large extent, this layer is preserved upon O2 exposure, despite a notable decrease in its thickness. Our findings highlight the dynamic behavior of the Co-TiO2 interface in reducing and oxidizing gas atmospheres, which is of particular interest to understand the performance of these materials in catalytic applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13192672/s1, Table S1: Textural, chemical, and crystallite properties of TiO2 supports; Figure S1: Photograph of the sample holder and details about the relationship between photon energy and analysis depth. Figure S2: The O 1s spectra of Co/TiO2 catalysts measured in 2.5 mbar CO2:H2 at 350 °C with 3 different excitation photon energies corresponding to 3 different analysis depths (a.d) (indicated in blue). Spectra are normalized to the same height to facilitate peak shape comparison. Refs. [47,48] are cited in Supplementary Materials.

Author Contributions

Conceptualization, S.Z., D.P.M., P.S. and K.S.; validation, S.Z. and C.S.; formal analysis, S.Z. and D.S.; investigation, S.Z. and A.E.; writing—original draft preparation, S.Z. and D.S.; writing—review and editing, S.Z. and D.S.; supervision, S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The research leading to these results has been supported by the project CALIPSOplus under the Grant Agreement 730872 from the EU Framework Programme for Research and Innovation. Part of this work was supported by the Agence Nationale de la Recherche (project ANR-19-CE07-0030), which is gratefully acknowledged.

Data Availability Statement

Data are available on request.

Acknowledgments

We want to acknowledge HZB for the allocation of synchrotron radiation beamtime (proposal no 222-11356-ST) and BESSY II Synchrotron staff for the collaboration during the experiments. We would like to thank M. Hävecker for his valuable support and M. Barreau for his help during the synchrotron measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, D.W.; Yoo, B.R. Advanced Metal Oxide (Supported) Catalysts: Synthesis and Applications. J. Ind. Eng. Chem. 2014, 20, 3947–3959. [Google Scholar] [CrossRef]
  2. Scarfiello, C.; Pham Minh, D.; Soulantica, K.; Serp, P. Oxide Supported Cobalt Catalysts for CO2 Hydrogenation to Hydrocarbons: Recent Progress. Adv. Mater. Interfaces 2023, 10, 2202516. [Google Scholar] [CrossRef]
  3. Sun, Y.; Yang, Z.; Dai, S. Nonclassical Strong Metal-Support Interactions for Enhanced Catalysis. J. Phys. Chem. Lett. 2023, 14, 2364–2377. [Google Scholar] [CrossRef]
  4. Chen, J.; Zhang, Y.; Zhang, Z.; Hou, D.; Bai, F.; Han, Y.; Zhang, C.; Zhang, Y.; Hu, J. Metal-Support Interactions for Heterogeneous Catalysis: Mechanisms, Characterization Techniques and Applications. J. Mater. Chem. A Mater. 2023, 11, 8540–8572. [Google Scholar] [CrossRef]
  5. Lou, Y.; Xu, J.; Zhang, Y.; Pan, C.; Dong, Y.; Zhu, Y. Metal-Support Interaction for Heterogeneous Catalysis: From Nanoparticles to Single Atoms. Mater. Today Nano 2020, 12, 100093. [Google Scholar] [CrossRef]
  6. Mitchell, S.; Qin, R.; Zheng, N.; Pérez-Ramírez, J. Nanoscale Engineering of Catalytic Materials for Sustainable Technologies. Nat. Nanotechnol. 2021, 16, 129–139. [Google Scholar] [CrossRef] [PubMed]
  7. Tauster, S.J.; Fung, S.C.; Garten, R.L. Strong Metal-Support Interactions. Group 8 Noble Metals Supported on TiO2. J. Am. Chem. Soc. 1978, 100, 170–175. [Google Scholar] [CrossRef]
  8. Xie, Y.; Wen, J.; Li, Z.; Chen, J.; Zhang, Q.; Ning, P.; Hao, J. Double-Edged Sword Effect of Classical Strong Metal–Support Interaction in Catalysts for CO2 Hydrogenation to CO, Methane, and Methanol. ACS Mater. Lett. 2023, 5, 2629–2647. [Google Scholar] [CrossRef]
  9. Frey, H.; Beck, A.; Huang, X.; van Bokhoven, J.A.; Willinger, M.G. Dynamic Interplay between Metal Nanoparticles and Oxide Support under Redox Conditions. Science 2022, 376, 982–987. [Google Scholar] [CrossRef]
  10. Monai, M.; Jenkinson, K.; Melcherts, A.E.M.; Louwen, J.N.; Irmak, E.A.; Van Aert, S.; Altantzis, T.; Vogt, C.; van der Stam, W.; Duchoň, T.; et al. Restructuring of Titanium Oxide Overlayers over Nickel Nanoparticles during Catalysis. Science 2023, 380, 644–651. [Google Scholar] [CrossRef]
  11. Beck, A.; Huang, X.; Artiglia, L.; Zabilskiy, M.; Wang, X.; Rzepka, P.; Palagin, D.; Willinger, M.G.; van Bokhoven, J.A. The Dynamics of Overlayer Formation on Catalyst Nanoparticles and Strong Metal-Support Interaction. Nat. Commun. 2020, 11, 3220. [Google Scholar] [CrossRef] [PubMed]
  12. De La Peña O’Shea, V.A.; Consuelo Álvarez Galván, M.; Platero Prats, A.E.; Campos-Martin, J.M.; Fierro, J.L.G. Direct Evidence of the SMSI Decoration Effect: The Case of Co/TiO2 Catalyst. Chem. Commun. 2011, 47, 7131–7133. [Google Scholar] [CrossRef] [PubMed]
  13. Hernández Mejía, C.; van Deelen, T.W.; de Jong, K.P. Activity Enhancement of Cobalt Catalysts by Tuning Metal-Support Interactions. Nat. Commun. 2018, 9, 4459. [Google Scholar] [CrossRef] [PubMed]
  14. van Deelen, T.W.; Hernández Mejía, C.; de Jong, K.P. Control of Metal-Support Interactions in Heterogeneous Catalysts to Enhance Activity and Selectivity. Nat. Catal. 2019, 2, 955–970. [Google Scholar] [CrossRef]
  15. Van Ravenhorst, I.K.; Hoffman, A.S.; Vogt, C.; Boubnov, A.; Patra, N.; Oord, R.; Akatay, C.; Meirer, F.; Bare, S.R.; Weckhuysen, B.M. On the Cobalt Carbide Formation in a Co/TiO2 Fischer-Tropsch Synthesis Catalyst as Studied by High-Pressure, Long-Term Operando X-ray Absorption and Diffraction. ACS Catal. 2021, 11, 2956–2967. [Google Scholar] [CrossRef]
  16. Qiu, C.; Odarchenko, Y.; Lezcano-Gonzalez, I.; Meng, Q.; Slater, T.; Xu, S.; Beale, A.M. Visualising Co Nanoparticle Aggregation and Encapsulation in Co/TiO2 Catalysts and Its Mitigation through Surfactant Residues. J. Catal. 2023, 419, 58–67. [Google Scholar] [CrossRef]
  17. Follath, R.; Hävecker, M.; Reichardt, G.; Lips, K.; Bahrdt, J.; Schäfers, F.; Schmid, P. The Energy Materials In-Situ Laboratory Berlin (EMIL) at BESSY II. J. Phys. Conf. Ser. 2013, 425, 212003. [Google Scholar] [CrossRef]
  18. Knop-Gericke, A.; Kleimenov, E.; Hävecker, M.; Blume, R.; Teschner, D.; Zafeiratos, S.; Schlögl, R.; Bukhtiyarov, V.I.; Kaichev, V.V.; Prosvirin, I.P.; et al. Chapter 4 X-ray Photoelectron Spectroscopy for Investigation of Heterogeneous Catalytic Processes. Adv. Catal. 2009, 52, 213–272. [Google Scholar] [CrossRef]
  19. Zhong, L.; Chen, D.; Zafeiratos, S. A Mini Review of in Situ Near-Ambient Pressure XPS Studies on Non-Noble, Late Transition Metal Catalysts. Catal. Sci. Technol. 2019, 9, 3851–3867. [Google Scholar] [CrossRef]
  20. Carbonio, E.A.; Teschner, D. Catalysts at Work by Near-Ambient Pressure X-ray Photoelectron Spectroscopy. Catal. Sci. Ser. 2023, 21, 265–382. [Google Scholar] [CrossRef]
  21. Petzoldt, P.; Eder, M.; Mackewicz, S.; Blum, M.; Kratky, T.; Günther, S.; Tschurl, M.; Heiz, U.; Lechner, B.A.J. Tuning Strong Metal-Support Interaction Kinetics on Pt-Loaded TiO2(110) by Choosing the Pressure: A Combined Ultrahigh Vacuum/Near-Ambient Pressure XPS Study. J. Phys. Chem. C 2022, 126, 16127–16139. [Google Scholar] [CrossRef]
  22. Papaefthimiou, V.; Dintzer, T.; Lebedeva, M.; Teschner, D.; Hävecker, M.; Knop-Gericke, A.; Schlögl, R.; Pierron-Bohnes, V.; Savinova, E.; Zafeiratos, S. Probing Metal-Support Interaction in Reactive Environments: An in Situ Study of PtCo Bimetallic Nanoparticles Supported on TiO2. J. Phys. Chem. C 2012, 116, 14342–14349. [Google Scholar] [CrossRef]
  23. Figueiredo, W.T.; Prakash, R.; Vieira, C.G.; Lima, D.S.; Carvalho, V.E.; Soares, E.A.; Buchner, S.; Raschke, H.; Perez-Lopez, O.W.; Baptista, D.L.; et al. New Insights on the Electronic Factor of the SMSI Effect in Pd/TiO2 Nanoparticles. Appl. Surf. Sci. 2022, 574, 151647. [Google Scholar] [CrossRef]
  24. Le Berre, C.; Serp, P.; Messou, D. Method for Preparing a Supported Metal Catalyst, Catalyst Obtained According to This Method and Uses. Google Patents No. WO2021224576A1, 11 November 2021. Available online: https://patents.google.com/patent/WO2021224576A1/en (accessed on 6 September 2023).
  25. Hendel, S.; Schäfers, F.; Hävecker, M.; Reichardt, G.; Scheer, M.; Bahrdt, J.; Lips, K. The EMIL Project at BESSY II: Beamline Design and Performance. AIP Conf. Proc. 2016, 1741, 030038. [Google Scholar]
  26. Smekal, W.; Werner, W.S.M.; Powell, C.J. Simulation of Electron Spectra for Surface Analysis (SESSA): A Novel Software Tool for Quantitative Auger-Electron Spectroscopy and X-ray Photoelectron Spectroscopy. Surf. Interface Anal. 2005, 37, 1059–1067. [Google Scholar] [CrossRef]
  27. Zemlyanov, D.Y. Practical Aspects of XPS: From Sample Preparation to Spectra Interpretation. Catal. Sci. Ser. 2023, 21, 13–50. [Google Scholar] [CrossRef]
  28. Abbate, M.; Goedkoop, J.B.; de Groot, F.M.F.; Grioni, M.; Fuggle, J.C.; Hofmann, S.; Petersen, H.; Sacchi, M. Probing Depth of Soft X-Ray Absorption Spectroscopy Measured in Total-Electron-Yield Mode. Surf. Interface Anal. 1992, 18, 65–69. [Google Scholar] [CrossRef]
  29. Luo, W.; Zafeiratos, S. Tuning Morphology and Redox Properties of Cobalt Particles Supported on Oxides by an in between Graphene Layer. J. Phys. Chem. C 2016, 120, 14130–14139. [Google Scholar] [CrossRef]
  30. Turczyniak, S.; Luo, W.; Papaefthimiou, V.; Ramgir, N.S.S.; Haevecker, M.; MacHocki, A.; Zafeiratos, S. A Comparative Ambient Pressure X-ray Photoelectron and Absorption Spectroscopy Study of Various Cobalt-Based Catalysts in Reactive Atmospheres. Top. Catal. 2016, 59, 532–542. [Google Scholar] [CrossRef]
  31. Breeson, A.C.; Sankar, G.; Goh, G.K.L.; Palgrave, R.G. Phase Quantification by X-ray Photoemission Valence Band Analysis Applied to Mixed Phase TiO2 Powders. Appl. Surf. Sci. 2017, 423, 205–209. [Google Scholar] [CrossRef]
  32. Bettac, A.; Bansmann, J.; Senz, V.; Meiwes-Broer, K.H. Structure and Magnetism of Hcp(0001) and Fcc(001) Thin Cobalt Films on a Clean and Carbon-Reconstructed W(110) Surface. Surf. Sci. 2000, 454–456, 936–941. [Google Scholar] [CrossRef]
  33. Cano-Casanova, L.; Ansón-Casaos, A.; Hernández-Ferrer, J.; Benito, A.M.; Maser, W.K.; Garro, N.; Lillo-Ródenas, M.A.; Román-Martínez, M.C. Surface-Enriched Boron-Doped TiO2 Nanoparticles as Photocatalysts for Propene Oxidation. ACS Appl. Nano Mater. 2022, 5, 12527–12539. [Google Scholar] [CrossRef] [PubMed]
  34. Lu, N.; Quan, X.; Li, J.Y.; Chen, S.; Yu, H.T.; Chen, G. Fabrication of Boron-Doped TiO2 Nanotube Array Electrode and Investigation of Its Photoelectrochemical Capability. J. Phys. Chem. C 2007, 111, 11836–11842. [Google Scholar] [CrossRef]
  35. Zhao, W.; Ma, W.; Chen, C.; Zhao, J.; Shuai, Z. Efficient Degradation of Toxic Organic Pollutants with Ni2O3/TiO2-XBx under Visible Irradiation. J. Am. Chem. Soc. 2004, 126, 4782–4783. [Google Scholar] [CrossRef]
  36. Netskina, O.V.; Kochubey, D.I.; Prosvirin, I.P.; Malykhin, S.E.; Komova, O.V.; Kanazhevskiy, V.V.; Chukalkin, Y.G.; Bobrovskii, V.I.; Kellerman, D.G.; Ishchenko, A.V.; et al. Cobalt-Boron Catalyst for NaBH4 Hydrolysis: The State of the Active Component Forming from Cobalt Chloride in a Reaction Medium. Mol. Catal. 2017, 441, 100–108. [Google Scholar] [CrossRef]
  37. Wang, L.; Yue, H.; Hua, Z.; Wang, H.; Li, X.; Li, L. Highly Active Pt/NaxTiO2 Catalyst for Low Temperature Formaldehyde Decomposition. Appl. Catal. B 2017, 219, 301–313. [Google Scholar] [CrossRef]
  38. Yang, M.; Liu, J.; Lee, S.; Zugic, B.; Huang, J.; Allard, L.F.; Flytzani-Stephanopoulos, M. A Common Single-Site Pt(II)-O(OH)x- Species Stabilized by Sodium on “Active” and “Inert” Supports Catalyzes the Water-Gas Shift Reaction. J. Am. Chem. Soc. 2015, 137, 3470–3473. [Google Scholar] [CrossRef]
  39. Cai, T.; Chen, X.; Johnson, J.K.; Wu, Y.; Ma, J.; Liu, D.; Liang, C. Understanding and Improving the Kinetics of Bulk Carbonation on Sodium Carbonate. J. Phys. Chem. C 2020, 124, 23106–23115. [Google Scholar] [CrossRef]
  40. Kong, D.; Wang, Y.; Huang, S.; Von Lim, Y.; Zhang, J.; Sun, L.; Liu, B.; Chen, T.; Valdivia y Alvarado, P.; Yang, H.Y. Surface Modification of Na2Ti3O7 Nanofibre Arrays Using N-Doped Graphene Quantum Dots as Advanced Anodes for Sodium-Ion Batteries with Ultra-Stable and High-Rate Capability. J. Mater. Chem. A Mater. 2019, 7, 12751–12762. [Google Scholar] [CrossRef]
  41. Zárate, R.A.; Fuentes, S.; Wiff, J.P.; Fuenzalida, V.M.; Cabrera, A.L. Chemical Composition and Phase Identification of Sodium Titanate Nanostructures Grown from Titania by Hydrothermal Processing. J. Phys. Chem. Solids 2007, 68, 628–637. [Google Scholar] [CrossRef]
  42. Dozzi, M.V.; Artiglia, L.; Granozzi, G.; Ohtani, B.; Selli, E. Photocatalytic Activity vs Structural Features of Titanium Dioxide Materials Singly Doped or Codoped with Fluorine and Boron. J. Phys. Chem. C 2014, 118, 25579–25589. [Google Scholar] [CrossRef]
  43. Zhong, L.; Kropp, T.; Baaziz, W.; Ersen, O.; Teschner, D.; Schlögl, R.; Mavrikakis, M.; Zafeiratos, S. Correlation between Reactivity and Oxidation State of Cobalt Oxide Catalysts for CO Preferential Oxidation. ACS Catal. 2019, 9, 8325–8336. [Google Scholar] [CrossRef]
  44. Kucheyev, S.O.; Van Buuren, T.; Baumann, T.F.; Satcher, J.H.; Willey, T.M.; Meulenberg, R.W.; Felter, T.E.; Poco, J.F.; Gammon, S.A.; Terminello, L.J. Electronic Structure of Titania Aerogels from Soft X-ray Absorption Spectroscopy. Phys. Rev. B Condens. Matter Mater. Phys. 2004, 69, 245102. [Google Scholar] [CrossRef]
  45. Qiao, L.; Xiao, H.Y.; Meyer, H.M.; Sun, J.N.; Rouleau, C.M.; Puretzky, A.A.; Geohegan, D.B.; Ivanov, I.N.; Yoon, M.; Weber, W.J.; et al. Nature of the Band Gap and Origin of the Electro-/Photo-Activity of Co3O4. J. Mater. Chem. C Mater. 2013, 1, 4628–4633. [Google Scholar] [CrossRef]
  46. Wang, W.; Wu, M.; Han, P.; Liu, Y.; He, L.; Huang, Q.; Wang, J.; Yan, W.; Fu, L.; Wu, Y. Understanding the Behavior and Mechanism of Oxygen-Deficient Anatase TiO2 toward Sodium Storage. ACS Appl. Mater. Interfaces 2019, 11, 3061–3069. [Google Scholar] [CrossRef]
  47. Tanuma, S.; Powell, C.J.; Penn, D.R. Calculations of Electron Inelastic Mean Free Paths. V. Data for 14 Organic Compounds over the 50–2000 EV Range. Surf. Interface Anal. 1994, 21, 165–176. [Google Scholar] [CrossRef]
  48. Briggs, D.; Seah, M.P. (Eds.) Practical Surface Analysis, Auger and X-ray Photoelectron Spectroscopy; Wiley: Hoboken, NJ, USA, 1996; Volume 1, ISBN 0471953407. [Google Scholar]
Figure 1. (a) Co 2p3/2, (b) Ti 2p, and (c) valence band (VB) NAP-XPS spectra of Co/m-TiO2 (black lines) and Co/TiO2 (red lines) catalysts measured under 5 mbar H2 at 200 °C (dark-colored lines) and 2.5 mbar CO2:H2 at 350 °C (light-colored lines). Spectra are normalized to the same height to facilitate peak shape comparison. The excitation photon energies for each spectrum (in blue) are selected so as to give photoelectrons with kinetic energies around 180 eV. The estimated analysis depth of the presented spectra is about 2.5 nm (calculated as 3 times the inelastic mean free path of the photoelectrons).
Figure 1. (a) Co 2p3/2, (b) Ti 2p, and (c) valence band (VB) NAP-XPS spectra of Co/m-TiO2 (black lines) and Co/TiO2 (red lines) catalysts measured under 5 mbar H2 at 200 °C (dark-colored lines) and 2.5 mbar CO2:H2 at 350 °C (light-colored lines). Spectra are normalized to the same height to facilitate peak shape comparison. The excitation photon energies for each spectrum (in blue) are selected so as to give photoelectrons with kinetic energies around 180 eV. The estimated analysis depth of the presented spectra is about 2.5 nm (calculated as 3 times the inelastic mean free path of the photoelectrons).
Nanomaterials 13 02672 g001
Figure 2. B 1s core level spectra of Co/m-TiO2 catalyst measured under (a) 5 mbar H2 at 200 °C and (b) 2.5 mbar CO2:H2 at 350 °C. Na 1s spectra of Co/m-TiO2 catalyst under (c) 5 mbar H2 at 200 °C and (d) 2.5 mbar CO2:H2 at 350 °C. Spectra are normalized to the same height to facilitate peak shape comparison. The Na 1s spectra were recorded at 3 different excitation photon energies corresponding to 3 different analysis depths (a.d.) (indicated in blue).
Figure 2. B 1s core level spectra of Co/m-TiO2 catalyst measured under (a) 5 mbar H2 at 200 °C and (b) 2.5 mbar CO2:H2 at 350 °C. Na 1s spectra of Co/m-TiO2 catalyst under (c) 5 mbar H2 at 200 °C and (d) 2.5 mbar CO2:H2 at 350 °C. Spectra are normalized to the same height to facilitate peak shape comparison. The Na 1s spectra were recorded at 3 different excitation photon energies corresponding to 3 different analysis depths (a.d.) (indicated in blue).
Nanomaterials 13 02672 g002
Figure 3. (a) O 1s and (b) C 1s spectra of Co/TiO2 and Co/m-TiO2 catalyst measured under 5 mbar H2 at 200 °C and 2.5 mbar CO2:H2 at 350 °C. The O 1s spectra are normalized to the same height while those of C 1s are not.
Figure 3. (a) O 1s and (b) C 1s spectra of Co/TiO2 and Co/m-TiO2 catalyst measured under 5 mbar H2 at 200 °C and 2.5 mbar CO2:H2 at 350 °C. The O 1s spectra are normalized to the same height while those of C 1s are not.
Nanomaterials 13 02672 g003
Figure 4. Normalized (a) Co L3,2-edge and (b) Ti L3,2-edge NEXAFS spectra of Co/m-TiO2 (black lines) and Co/TiO2 (red lines) catalysts measured under 5 mbar H2 at 200 °C (dark-colored lines) and 2.5 mbar CO2:H2 at 350 °C (light-colored lines).
Figure 4. Normalized (a) Co L3,2-edge and (b) Ti L3,2-edge NEXAFS spectra of Co/m-TiO2 (black lines) and Co/TiO2 (red lines) catalysts measured under 5 mbar H2 at 200 °C (dark-colored lines) and 2.5 mbar CO2:H2 at 350 °C (light-colored lines).
Nanomaterials 13 02672 g004
Figure 5. Co 2p3/2, Ti 2p, B 1s, Na 1s, valence band (VB) NAP-XPS, and Co and Ti L3,2-edge spectra of Co/m-TiO2 (blue lines) and Co/TiO2 (magenta lines) catalysts measured in 1 mbar O2 at 350 °C. The excitation photon energies for each NAP-XPS spectrum (shown in blue) are selected so as to produce photoelectrons with kinetic energies around 180 eV. The estimated information depth is about 2.5 nm (calculated as 3 times the inelastic mean free path of the photoelectrons). The presented spectra are normalized to the same height.
Figure 5. Co 2p3/2, Ti 2p, B 1s, Na 1s, valence band (VB) NAP-XPS, and Co and Ti L3,2-edge spectra of Co/m-TiO2 (blue lines) and Co/TiO2 (magenta lines) catalysts measured in 1 mbar O2 at 350 °C. The excitation photon energies for each NAP-XPS spectrum (shown in blue) are selected so as to produce photoelectrons with kinetic energies around 180 eV. The estimated information depth is about 2.5 nm (calculated as 3 times the inelastic mean free path of the photoelectrons). The presented spectra are normalized to the same height.
Nanomaterials 13 02672 g005
Figure 6. Comparison of atomic concentrations determined from NAP-XPS spectra (analysis depth 2.5 nm) acquired in H2, CO2:H2, and O2 atmospheres at the temperatures indicated at the bottom of the graph for (a) Co/TiO2 and (b) Co/ TiO2 catalysts. For comparison, the bulk atomic concentrations from ICP-OES measurements are presented. Above the black bars, the atomic concentration of cobalt is indicated.
Figure 6. Comparison of atomic concentrations determined from NAP-XPS spectra (analysis depth 2.5 nm) acquired in H2, CO2:H2, and O2 atmospheres at the temperatures indicated at the bottom of the graph for (a) Co/TiO2 and (b) Co/ TiO2 catalysts. For comparison, the bulk atomic concentrations from ICP-OES measurements are presented. Above the black bars, the atomic concentration of cobalt is indicated.
Nanomaterials 13 02672 g006
Figure 7. The atomic concentration calculated based on NAP-XPS and NAP-HAXPES spectra as a function of the analysis depth (defined as 3 times the IMFP) measured at 200 °C in 5 mbar H2 (a,c) and (b,d) at 350 °C in 1 mbar O2 for Co/TiO2 (a,b) and (c,d) Co/m-TiO2 catalysts.
Figure 7. The atomic concentration calculated based on NAP-XPS and NAP-HAXPES spectra as a function of the analysis depth (defined as 3 times the IMFP) measured at 200 °C in 5 mbar H2 (a,c) and (b,d) at 350 °C in 1 mbar O2 for Co/TiO2 (a,b) and (c,d) Co/m-TiO2 catalysts.
Nanomaterials 13 02672 g007
Figure 8. Illustration of the proposed surface arrangement of Co/TiO2 and Co/m-TiO2 catalysts in H2 and CO2:H2 atmospheres as well as in subsequent exposure to O2.
Figure 8. Illustration of the proposed surface arrangement of Co/TiO2 and Co/m-TiO2 catalysts in H2 and CO2:H2 atmospheres as well as in subsequent exposure to O2.
Nanomaterials 13 02672 g008
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

Salusso, D.; Scarfiello, C.; Efimenko, A.; Pham Minh, D.; Serp, P.; Soulantica, K.; Zafeiratos, S. Direct Evidence of Dynamic Metal Support Interactions in Co/TiO2 Catalysts by Near-Ambient Pressure X-ray Photoelectron Spectroscopy. Nanomaterials 2023, 13, 2672. https://doi.org/10.3390/nano13192672

AMA Style

Salusso D, Scarfiello C, Efimenko A, Pham Minh D, Serp P, Soulantica K, Zafeiratos S. Direct Evidence of Dynamic Metal Support Interactions in Co/TiO2 Catalysts by Near-Ambient Pressure X-ray Photoelectron Spectroscopy. Nanomaterials. 2023; 13(19):2672. https://doi.org/10.3390/nano13192672

Chicago/Turabian Style

Salusso, Davide, Canio Scarfiello, Anna Efimenko, Doan Pham Minh, Philippe Serp, Katerina Soulantica, and Spyridon Zafeiratos. 2023. "Direct Evidence of Dynamic Metal Support Interactions in Co/TiO2 Catalysts by Near-Ambient Pressure X-ray Photoelectron Spectroscopy" Nanomaterials 13, no. 19: 2672. https://doi.org/10.3390/nano13192672

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