Next Article in Journal
Improvement of Echinacea purpurea and Ganoderma lucidum Extracts with Cell Model on Influenza A/B Infection
Previous Article in Journal
Cannabis sativa as an Herbal Ingredient: Problems and Prospects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N2 as an Efficient IR Probe Molecule for the Investigation of Ceria-Containing Materials

by
Kristina K. Chakarova
1,*,
Mihail Y. Mihaylov
1,*,
Bayan S. Karapenchev
1,2,
Iskra Z. Koleva
2,
Georgi N. Vayssilov
2,
Hristiyan A. Aleksandrov
1,2 and
Konstantin I. Hadjiivanov
1
1
Institute of General and Inorganic Chemistry, Bulgarian Academy of Sciences, 1113 Sofia, Bulgaria
2
Faculty of Chemistry and Pharmacy, University of Sofia, 1126 Sofia, Bulgaria
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(15), 3608; https://doi.org/10.3390/molecules29153608
Submission received: 2 July 2024 / Revised: 26 July 2024 / Accepted: 29 July 2024 / Published: 30 July 2024

Abstract

:
Ceria and ceria-based catalysts are very important in redox and acid-base catalysis. Nanoceria have also been found to be important in biomedical applications. To design efficient materials, it is necessary to thoroughly understand the surface chemistry of ceria, and one of the techniques that provides such information about the surface is the vibrational spectroscopy of probe molecules. Although the most commonly used probe is CO, it has some disadvantages when applied to ceria and ceria-based catalysts. CO can easily reduce the material, forming carbonate-like species, and can be disproportionate, thus modifying the surface. Here, we offer a pioneering study of the adsorption of 15N2 at 100 K, demonstrating that dinitrogen can be more advantageous than CO when studying ceria-based materials. As an inert gas, N2 is not able to oxidize or reduce cerium cations and does not form any surface anionic species able to modify the surface. It is infrared and transparent, and thus there is no need to subtract the gas phase spectrum, something that often increases the noise level. Being a weaker base than CO, N2 has a negligible induction effect. By using stoichiometric nano-shaped ceria samples, we concluded that 15N2 can distinguish between surface Ce4+ sites on different, low index planes; with cations on the {110} facets and on some of the edges, Ce4+15N2 species with IR bands at 2258–2257 cm−1 are formed. Bridging species, where one of the N atoms from the molecule interacts with two Ce4+ cations, are formed on the {100} facets (2253–2252 cm−1), while the interaction with the {111} facets is very weak and does not lead to the formation of measurable amounts of complexes. All species are formed by electrostatic interaction and disappear during evacuation at 100 K. In addition, N2 provides more accurate information than CO on the acidity of the different OH groups because it does not change the binding mode of the hydroxyls.

Graphical Abstract

1. Introduction

Materials based on ceria are highly effective for a wide range of applications, such as heterogeneous catalysis, chemical sensing, and biomedicine. In particular, in catalysis, ceria plays a key role as an active support or component in various processes for automotive exhaust gas aftertreatment, the oxidation of volatile organic compounds, CO2 conversion, low-temperature water-gas shift (WGS) reaction, organic compound synthesis, and more [1,2,3,4].
The importance of ceria for these applications stems from its exceptional surface (and partly bulk) properties, mainly related to the ability of Ce ions to easily switch between Ce4+ and Ce3+ oxidation states [1,3]. However, some organic reactions, such as dehydration and ketonization, are catalyzed by acid-base sites, while others, such as addition, substitution, isomerization, and ring-opening reactions, require both acid-base and redox centers [5].
Due to its great significance, CeO2 has been highly studied, and the development of active materials based on CeO2 represents a key current topic in science [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17]. A search in Scopus (keywords: ceria or cerium (di)oxide and surface) shows a sharp increase in interest in this field since the beginning of the century: 548 documents in 2000 and 16,800 documents in 2023.
Stoichiometric CeO2 possesses a cubic fluorite structure with eightfold coordination for cerium ions and tetrahedral coordination for oxygen anions. According to the large size of the cations, CeO2 shows medium Lewis acidity and relevant surface basicity. Under reducing conditions, it releases oxygen, creating oxygen vacancies within the preserved fluorite structure, and in an oxidizing atmosphere, it is easily oxidized back to CeO2 [6,7]. Note that the creation of oxygen vacancies can modify the acid-base properties of ceria.
The surface properties of CeO2 depend on the size and shape of the crystallites [8,9,10]. Smaller particles are characterized by a larger active surface area, more edges, steps, and vertices, and favor reduction. Recently, to clarify the relationship between activity and structure, special attention has been given to ceria nanoparticles with specific shapes [11,12,13,14,15,16,17], as they preferentially expose certain low-index planes. Thus, ideal nanocubes are enclosed by the {100} facets, while ideal nanooctahedra are enclosed by the {111} facets. It has been found that, compared to the {111} facet, the {100} and {110} facets are more active in CO oxidation [10,12,18,19,20,21], H2 oxidation, WGS [8,22], and the reverse WGS reaction [23]. It has also been reported that the {100} facet is more active in the dehydrogenation of 2-propanol to acetone, while the {111} facet demonstrated higher activity in propanol dehydration [24].
The design of effective materials based on CeO2 requires a deep understanding of the surface chemistry of CeO2, and one of the techniques that provides rich information about the surface is vibrational spectroscopy [25,26]. This technique can directly observe surface hydroxyl groups and some impurities, such as residual carbonates or nitrates. Specifically, for cerium dioxide and materials based on cerium dioxide, vibrational spectroscopy can also be used to monitor its degree of reduction through the Ce3+ spin-orbital 2F5/22F7/2 electronic transition at around 2147–2110 cm−1 [6,7,27,28,29,30,31,32].
Additional data on the nature of various surface sites can be obtained using probe molecules. Among the different probes, CO is the most informative and widely used [25]. However, the use of CO to characterize CeO2-based materials has some issues. CO can easily reduce CeO2, forming carbonates, and can even be disproportionate, thus modifying the surface [33]. Nonetheless, reactive adsorption can be largely suppressed at low temperatures. When forming carbonyls with Ce4+ cations, CO is bound by σ and electrostatic bonds, which leads to a blue shift of ν(C−O). There is no consensus in the literature regarding the assignment of different carbonyl bands, especially the oxidation state of cerium in the complexes. The acidity of the hydroxyl groups is also under debate.
Recently, we published a comprehensive study on CO adsorption on stoichiometric CeO2, combining IR spectroscopy and DFT calculations [34]. It was revealed that ν(C−O) is sensitive to the localization of the Ce4+ adsorption sites and decreases in the order edge/apex, {110}, {100}, and {111} facets. However, significant band shifts with increasing coverage, due to the inductive effect, the formation of dicarbonyls, and associated bridging carbonyls, hinder a quantitative assessment of these centers. Additionally, the results indicated that CO adsorption could induce reconstructions in some hydroxyl groups, thus compromising the measurement of proton acidity.
In search of a more appropriate probe molecule, we selected dinitrogen. Although ν(N−N) of dinitrogen is not IR-active in the gas phase, it becomes activated upon adsorption due to symmetry lowering. 15N2 was chosen in order to avoid any hindrance from gaseous CO2. The use of N2 as a probe molecule to test surface acidity at low temperatures has been proposed for different systems [35,36,37,38,39,40,41,42,43]. However, to the best of our knowledge, dinitrogen has not yet been applied as a probe for CeO2-based systems. Compared to CO, dinitrogen offers certain advantages. There is no need to subtract the spectrum of the gas phase. As an inert gas, N2 cannot oxidize or reduce cerium cations, and it does not form any surface anionic species capable of modifying the surface. Being a weaker base than CO, N2 is not expected to have a noticeable induction effect or cause surface reconstructions.
Preliminary studies have also indicated that dinitrogen is also a suitable probe for detecting Ce3+ sites. Investigations on this subject are now in progress and will allow for a comprehensive understanding of the surface chemistry of ceria.
In this work, we demonstrate the potential of dinitrogen as an IR probe molecule for studying stoichiometric ceria nanoparticles. Dinitrogen adsorption measurements were conducted in parallel with those of CO, allowing for direct comparison. The interpretation of the experimental data was also supported by theoretical modeling of the possible dinitrogen complexes and by calculating their wavenumber and binding energy.

2. Results and Discussion

2.1. Basic Characteristics of the Samples

In this work, we studied three ceria samples: nanocubes (CeO2-NC), nanopolyhedra (CeO2-NP), and nanorods (CeO2-NR). The different shapes of the crystallites are achieved by varying the conditions of the hydrothermal synthesis, the concentration of NaOH, and the reaction temperature. The main characteristics of these samples are reported elsewhere [34], and the most important are summarized in Table 1. Energy dispersive X-ray (EDX) analysis did not show the presence of residual sodium [34].
Briefly, the CeO2-NC sample consisted of particles mainly enclosed by the {100} facets, although some {110} and {111} facets were exposed to a lesser extent. The CeO2-NP sample was obtained by calcination of CeO2-NC at 923 K. In agreement with earlier reports [34,44,45], this treatment led to a strong increase in the relative exposure of the {111} facets at the expense of the {100} facets. The CeO2-NR sample exposed mainly the {110} facets, followed by {111} and {100}. It was characterized by the smallest particles and had the highest specific surface area. The main particle size of CeO2-NC was 27 nm, and the surface area was 31 m2 g−1. These values only slightly decreased after calcination at 923 K to produce CeO2-NP.
Therefore, depending on the shape of the crystallites, different low-index facets predominated in each of our samples. The models of the different facets we used are described below. In all cases, the nanoparticles have a significant number of centers located at edges, corners, and defects.

2.2. Lewis Acidic Ce4+ Sites

Before the adsorption of probe molecules, the samples were subjected to activation consisting of successive thermal treatments in O2 and a vacuum. This process aimed to remove water, organic impurities, carbonates, etc., from the surface and to ensure the stoichiometric state of the cerium samples. After activation, virtually all the cerium is in the Ce4+ form [46], as evidenced by the absence of the Ce3+ electronic transition band at approximately 2115 cm−1 [6,19], the yellow color [47], and the lack of the formation of reactive oxygen species (O2, O22−) upon O2 adsorption [34,46]. However, small amounts of oxyhydroxide and carbonates remain. Subsequently, the samples were brought into contact with heavy water and gradually dehydrated and dehydroxylated by evacuation at elevated temperatures to create bare Ce4+ Lewis acid sites.
Adsorption of 15N2 was first studied on the CeO2-NC sample evacuated at different temperatures (Figure 1A). An increase in evacuation temperature leads to gradual sample dehydroxylation and the creation of bare Ce4+ Lewis acid sites [34].
The introduction of 15N2 (10 mbar) to the sample evacuated at 298 K leads only to the appearance of a very weak band at 2252 cm−1, which is attributed to OH−15N2 interaction [36,40,41]. Indeed, the band develops alongside the shifting of hydroxyl bands (for more details, see below). The situation is similar with the sample evacuated at 373 K, but, in this case, another weak band at 2258 cm−1 is discernible. This later band significantly increases in intensity after preliminary evacuation at 473 K, while the OH−15N2 band fades. When the sample was evacuated at 673 K, the band at 2257 cm−1 further developed, and a strong band at 2252 cm−1 appeared. Both bands were even slightly more intense in a sample evacuated at 773 K.
The results described show that the spectra of dinitrogen adsorbed on stoichiometric CeO2-NC are relatively simple and consist of two bands (at 2257 and 2252 cm−1) the intensity of which strongly depends on the sample dehydroxylation degree. In fact, there are two species with different stabilities associated with the 2252 cm−1 band. The band at 2257 cm−1 seems to have a component at 2258 cm−1. We noted that the band maxima practically do not depend on the coverage (see Figure S1), which suggests (i) the lack of lateral and vibrational interaction between the adsorbed molecules and (ii) no measurable formation of geminal species. In any case, the results are somewhat surprising, because the use of CO as a probe molecule has clearly indicated the existence of at least three kinds of Ce4+ sites with different acidities [34]. In order to correctly assign the bands, we compared the results with results on CO adsorption. As already noted, the CO adsorption experiments were performed immediately after the experiments on N2 adsorption.
Now, we consider the results obtained with CO as a probe (Figure 1B), which are consistent with our previous report [34]. Because CO is more strongly adsorbed than 15N2 and, at high coverage, forms geminal structures, we examined the spectra registered at intermediate (ca. 50%) coverage. There are three main bands in the spectra. The first band is at 2175–2171 cm−1 and it first appears with the sample evacuated at 373 K, reaching maximal intensity with the sample evacuated at 473 K. This band is attributed to CO adsorbed on Ce4+ sites from the {110} facets (component at 2171 cm−1) and Ce4+ on edges (component at 2175 cm−1) [34]. The second band is at 2162 cm−1 and it becomes well discernible with the sample evacuated at 673 K. This band is associated with the {100} facets, which retain OH groups at relatively high temperatures. Finally, the third band is at ca. 2155 cm−1 and is of significant intensity even with the sample evacuated at ambient temperature. This band was attributed to CO adsorbed on the {111} facets, which are not hydroxylated and hold only weakly adsorbed water molecules at ambient conditions.
Comparison between the spectra of adsorbed 15N2 (Figure 1A) and CO (Figure 1B) allows for the determination of the following conclusions:
  • Dinitrogen, unlike CO, does not bind to the weakly acidic surface Ce4+ sites located on the {111} facets. This follows from the fact that there is no 15N2 band in Figure 1A changing in concert with the CO band at 2155–2152 cm−1 (Figure 1B).
  • 15N2 adsorbed on the {110} facets and some edges is characterized by a band at 2258–2257 cm−1. Indeed, this band follows the changes of the CO band at 2175–2171 cm−1. We tentatively assign the component at 2158 cm−1 to dinitrogen on the edges and the component at 2257 cm−1 to dinitrogen on the {110} facets.
  • The band at 2252 cm−1 (2253 cm−1 at low coverage) is attributed to 15N2 on the {100} facets. It gains significant intensity when the sample is evacuated at 673 K or above, exactly the same as the carbonyl band at 2162 cm−1. A minor component of this band, easily disappearing after evacuation, characterizes OH−15N2 adducts.
To obtain further confirmation of the proposed assignments, we compared the spectra of 15N2 adsorbed on three samples, CeO2-NC, CeO2-NP, and CeO2-NR, which differ in their exposed faces (Figure 2A). To ensure a high dehydroxylation degree, the samples were evacuated at 773 K. As expected, the band at 2252 cm−1, due to the dinitrogen adsorbed on {100} facets, is most intense with CeO2-NC, and weak with the CeO2-NP sample, consistent with the development of the inert {111} at the expense of the {100} facets. The band at 2257 cm−1 is most intense with the CeO2-NR sample, in agreement with the high exposure of the {110} facets. These observations are in full agreement with the results on CO adsorption (Figure 2B).
Although the general picture of 15N2 adsorption on ceria seems to be clear, some additional details can be extracted from the analysis of the spectra recorded at decreasing coverage (Figure S1). The maximum of the band at 2257 cm−1 appears to be practically coverage-independent. However, a slight blue shift, less than 1 cm−1, is observed for the band at 2252 cm−1 with coverage decreasing. This shift can be justified by the fact that, at high coverage, the band is superimposed on the band of the OH−15N2 complexes at ca. 2252 cm−1. However, it seems that a negligible shift also occurs with the component due to 15N2 polarized by Ce4+ sites. In any case, the wavenumber of the band at intermediate coverages, after destroying the OH−15N2 species, is close to 2253 cm−1.

2.3. DFT Modeling of N2 Adsorption on Ceria Models

We theoretically investigated the adsorption of N2 on Ce4+ cations located on the surface of four ceria models: the CeO2(111) surface model with step, the ideal CeO2(100) and CeO2(110) surface models, as well as the Ce40O80 nanoparticle (see Table 2). The obtained structures for dinitrogen adsorption on CeO2(100) and CeO2(110) surfaces are shown in Figure 3 and Figure 4, respectively. For direct comparison with experimental values, all frequencies in the text correspond to 15N2.
When N2 is adsorbed to a four-coordinated Ce4+ cation located at the corner of the Ce40O80 nanoparticle model, its binding energy (BE) is −0.30 eV, while the Ce4+−N distance is 309 pm. The calculated N−N vibrational wavenumber, 2254 cm−1, is only 2 cm−1 higher than the corresponding calculated value for the gas phase 15N2 molecule, at 2252 cm−1. The BE per N2 ligand for the Ce4+(N2)2 complex, −0.27 eV, is slightly lower than the value for the monoligand complex. Similarities between both complexes were also observed for the Ce4+−N distances, which are only 1–2 pm longer in the diligand complex than in the monoligand complex. The calculated frequencies for the Ce4+(N2)2 complex are 2258 and 2254 cm−1.
The binding energy of N2 is −0.26 ÷ −0.27 eV when the ligand is adsorbed at a top position to a Ce4+ cation located at the terrace of the CeO2(111) surface (Figure S2). The calculated N−N vibrational wavenumber, 2255 cm−1, is slightly shifted by 2–3 cm−1 to higher frequencies with respect to the corresponding value for the 15N2 in the gas phase, at 2252 cm−1. Most probably, the low binding energy of N2 is the reason that we do not detect an experimental band when N2 is adsorbed to CeO2(111). Stronger adsorption was found when N2 interacts with the low-coordinated Ce4+ cation at the edge of the step, as the BE value is −0.31 eV, while the calculated N−N wavenumber is 2261 cm−1.
As recently reported, the CeO2(100) surface is fully hydroxylated at ambient conditions [34,48], and, upon dehydroxylation, half of the oxygen anions are removed as part of the water molecules [34]. For this reason, for this surface, we created two models where we moved half of the surface O to the other (initially cerium-terminated) surface. In the first model, the surface O centers were moved along the <100> direction. In the second, oxygen atoms were moved in checkerboard style. Both models provide very similar results, as the N2 prefers to interact with two Ce4+ cations in a bridge coordination (Figure 3a and Figure S3). In this position, the BE of N2 is ~−0.45 eV, which is twice as large as an absolute value as the BE of N2 at top coordination, at −0.20 eV. Despite the different stabilities, the N−N vibrational wavenumber is essentially the same in both coordination modes, 2253 cm−1, and does not change much from the corresponding value for the 15N2 molecule in the gas phase.
We also calculated a full monolayer of N2 molecules initially located in the bridge or top positions. The most stable structure is again when all N2 molecules are in bridge positions (Figure 3c), as the binding energy per N2 molecule, −0.43 eV, is the same as the corresponding value for the structure with only one N2 adsorbate, manifesting the lack of significant lateral interactions between the adsorbates. When all N2 molecules are initially located in top coordination, they change their coordination to bridge during the geometry optimization.
The adsorption complex of one N2 ligand with top coordination to the CeO2(110) surface model (Figure 4) has a BE of −0.30 eV, and the calculated wavenumber, 2261 cm−1, is higher by 9 cm−1 than the corresponding value for the gas phase 15N2 molecule.

2.4. Acidity of Hydroxyl Groups

The low-temperature adsorption of CO and N2 was also used to determine the acidity of hydroxyl groups on ceria. In order to obtain high-quality spectra, we examined deuterated samples because the noise level is much lower in the ν(OD) region as compared to ν(OH). Acidity was assessed based on the red shift of the ν(OD) modes induced by the probe molecules [25,26]. The greater the shift, the higher the acidity. It should be noted that the CO-induced shift is approximately 2.5 times larger than that induced by N2 due to its higher basicity [26,40]. For these experiments, we studied CeO2-NC evacuated at 473 and 773 K, respectively.
Generally, the OH/OD groups on ceria are not expected to exhibit high acidity [8,49]. For CeO2-NC, two main OD bands were observed around 2745 and 2705 cm−1, attributed to the Ce4+-associated terminal and bridging OD groups, respectively. Additionally, at lower frequencies, a series of weak bands characteristic of hydrogen-bonded OD groups in cerium oxyhydroxide were detected. It should be noted that cooling the sample to 100 K causes a slight blue shift of the OD bands.
Figure 5 demonstrates the effect of CO and N2 adsorption on the OD spectra of CeO2-NC.
CO adsorption on the sample evacuated at 773 K (Figure 5B, spectra d–f) results in a shift of the bands of terminal and bridging deuteroxyls to a single band around 2682 cm−1. Thus, the observed values of Δν(OD) are −61 and −22 cm−1, corresponding to −83 and −30 cm−1 for Δν(OH), respectively. For comparison, the shift of the SiOH modes on silica is −90 cm−1 [26]. Consistent with expectations, the measured acidity for bridging hydroxyls is quite low, lower than the reported value for silanols on silica. However, the shift for terminal OD groups is close to that for silanols, pointing out relatively higher acidity. We explain this discrepancy by the assumption that terminal OD groups adopt a bridging configuration upon complexation with CO [34]. As a result, the CO-induced shift of the bands attributed to terminal OD groups consists of two components: one due to structural transformation and the other due to complexation with CO. Therefore, the actual acidity of terminal OH/OD groups is lower than the measured one. A similar process has been proposed previously for silanols on amorphous silica-alumina [50].
It should also be noted that deuterated hydroxyls associated with cerium oxyhydroxide impurities (bands at 2614, 2603, and 2594 cm−1) are not affected by CO, as these groups are already involved in a D-bonding interaction, which is stronger than their interaction with CO.
The changes in OD spectra following adsorbed dinitrogen show that both bands, of the terminal and bridging OD groups, shift very slightly, ~10 cm−1, to two new bands (Figure 5A, lower set of spectra d–f). This corresponds to a CO-induced shift of approximately 25 cm−1, similar to that observed with CO for bridging deuteroxyls. However, for terminal OD groups, we measured with CO a significantly larger shift of around 80 cm−1, explained by the change in the OD binding mode, i.e., their transformation into bridging OD groups. Evidently, 15N2, being a weaker base, cannot induce similar transformation, thus allowing terminal groups to retain their initial configuration.
Finally, we examined a sample evacuated at 473 K and exhibiting higher OD coverage. Following CO adsorption, two shifted bands are observed at 2681 and 2650 cm−1 (Figure 5B, upper set of spectra a–c). The additional shifted band indicates the presence of hydroxyl groups with enhanced acidity, which disappear at higher evacuation temperatures. A similar effect is observed during dinitrogen adsorption (Figure 5A, upper set of spectra a–c). However, in this case, we clearly observed that terminal groups shift by −10 cm−1 to one band, while bridging groups shift to two other bands, by −10 cm−1 and −15 cm−1, respectively. Hence, we can confidently conclude that the hydroxyls with increased acidity are of the bridging type.

2.5. Discussion

The Lewis acidity of ceria develops at the expense of dehydration and the dehydroxylation of the surface. At ambient conditions, the {100} and {110} facets are fully hydroxylated [34,48], while the {111} facet is hydrated [34,51]. During evacuation, even at ambient temperature, the {111} facet is dehydrated, thus leaving exposed 7-coordinated Ce4+ cations, which are the weakest Lewis acid sites on stoichiometric ceria. With CO, these sites form carbonyls characterized by a band at 2155–2152 cm−1 and easily disappear during pumping, even at 100 K. The low value of ν(CO) and its low stability are consistent with the weak adsorbent-adsorbate interaction. Dinitrogen is a weaker base than CO [35,40] and it appears that it is not able to form complexes with the Ce4+ sites from the {111} plane, even at 100 K. In principle, it is possible that some highly symmetric and infrared inactive adducts are formed. However, our DFT calculations do not support such a possibility.
Evacuation at higher temperatures leads first to dehydroxylation of the {110} plane via a recombination of the terminal and triply-bridged OH groups and the appearance of 6-coordinated Ce4+ cations with enhanced acidity. We also propose that, under similar conditions, a fraction of 6-coordinated cations with similar electrophilicity, but situated on edges, is formed. With CO, these sites form linear complexes (ca. 2171 cm−1 for planes and 2175 cm−1 for edges), which are converted, at high coverage, into geminal species. However, with the weaker base 15N2, these Ce4+ sites form exclusively linear species characterized by a band at 2258–2257 cm−1, with the higher wavenumber component likely characterizing edge sites. No evidence for geminal species was found.
A further increase in the evacuation temperature above 573 K leads to dehydroxylation of the {100} plane via a recombination of doubly-bridged OH groups. The Ce4+ formed is 6-coordinated, but the geometric arrangement favors the formation of bridging complexes. As a result, CO adsorption leads to the appearance of carbonyls bridging two Ce4+ sites by their carbon atom. The maximum of the carbonyl band is strongly coverage-dependent and is detected at 2169 cm−1, at very low coverage, and at 2162 cm−1, at intermediate coverage. Note that, in this case, the interaction is essentially electrostatic, and the adducts formed are very different in nature from the classic bridging carbonyls on metal surfaces where the rehybridization of CO occurs. With 15N2, these sites also form bridging adducts, absorbing at 2252 cm−1, and the band maximum is practically coverage-independent due to the weak induction effect of dinitrogen.
In addition to the Ce4+ sites, stoichiometric ceria are characterized by surface OH groups that, under certain conditions, can cover significant fractions of the {100} and {110} planes. Both probe molecules, CO and 15N2, could be used for the estimation of protonic acidity. In both cases, the bands due to the probe molecule interacting with OH groups are of relatively low intensity and are even weaker in the case of 15N2 due to its low basicity. Good agreement was found between the acidity of the bridging hydroxyls, as measured by the two probe molecules. However, it appears that, in this case, the use of 15N2 is advantageous because it allows for measuring the acidity of the terminal OH groups. As we proposed earlier, the failure of CO to measure this acidity is likely due to the fact that it, being a relatively strong base, causes the transformation of linear to bridging OH groups when carbonyl adducts are formed.
For convenience, the observed and calculated frequencies of CO and 15N2 adsorbed on different surface strictures of ceria are summarized in Table 3.
In summary, the results indicate that dinitrogen shows some similarities and some differences from CO as a probe molecule for testing ceria surfaces. The main disadvantage of 15N2 seems to be its impossibility to detect Ce4+ sites on the {111} ceria face. However, this could also be regarded as an advantage because 15N2 probes only the strongest Lewis acid sites. Another drawback of dinitrogen is the relatively low intensity of its IR band. However, this could be a problem only when studying materials with a low content of cerium. One could also speculate that the spectral interval of the detected bands is too narrow, but this is compensated by the negligible coverage dependence of the band maxima.
On the contrary, it seems that 15N2 has many advantages as a probe molecule for Ce-containing materials as compared to CO. First, no reactive adsorption that can modify the surface (reduction, formation of carbonates) occurs. Due to weak basicity, 15N2 affects the surface very weakly, and, as a result, the bands of adsorbed species are practically coverage-independent. Moreover, no geminal species are formed at high coverage. All this allows for an easy comparison between the results obtained in different laboratories. In contrast to CO, 15N2 gives reliable results on the acidity of the different OH groups on ceria.
Finally, we would like to note that this study was performed with stoichiometric ceria. However, to efficiently use one probe molecule, it is necessary to know how it interacts with reduced Ce3+ sites. Investigations in this respect are now in progress in our laboratory. Generally, the complexes of 15N2 on reduced ceria are detected at lower frequencies and with lower intensity as compared to the stoichiometric samples.

3. Materials and Methods

The ceria samples used in this work were investigated and characterized in our previous study [34]. Ce(NO3)3.6H2O (Fluka, Buchs, Switzerland, 99% purity) and NaOH (Merck, Darmstadt, Germany, 99% purity) were used for the synthesis. The CeO2-NC and CeO2-NR samples were prepared using the hydrothermal method [19,52]. Briefly, 85 mL of an aqueous solution of 5 g of Ce(NO3)3.6H2O were added to 150 mL of 36 wt. % aqueous NaOH solution with vigorous stirring. Then, the reaction mixture was transferred to an autoclave. CeO2-NC and CeO2-NR were obtained after aging in the autoclave for 24 h at 453 and 373 K, respectively. The suspensions were then centrifuged, the precipitates thoroughly washed with deionized water, dried at 393 K, and finally calcined in air at 673 K for 2 h. The CeO2-NP sample was obtained by calcination of a fraction of the CeO2-NC sample at 923 K for 1 h. This is in line with literature reports showing surface reconstruction of ceria nanocubes at high temperatures, finally leading to the formation of {111} facets [44,45].
The IR spectra were recorded with a Thermo Scientific Nicolet 6700 FTIR spectrometer (Madison, WI, USA) using an MCT-A detector. Each spectrum was obtained by the accumulation of 64 scans at a spectral resolution of 2 cm−1 and a precision of 0.01 cm−1. Sample powders were pressed into self-supporting pellets (ca. 10 mg cm−2). The latter were treated in situ in a home-made IR cell connected to a vacuum-adsorption apparatus with a residual pressure below 10−3 Pa. The cell allowed measurements between ambient temperature and 100 K.
First, the samples were activated by heating in 100 mbar O2 at 773 K for 30 min, followed by 30 min of evacuation at the same temperature. To obtain deuteroxylated ceria, the samples were exposed to D2O vapor at room temperature and then evacuated. This procedure was repeated several times. Finally, before the adsorption experiment, the samples were evacuated (dehydrated/dehydroxylated) at a certain elevated temperature for 30 min.
Adsorption of the two probe molecules (15N2 and CO) was performed successively at 100 K. First, dinitrogen adsorption was conducted at an equilibrium pressure of 10 mbar, followed by evacuation. Then, carbon monoxide was adsorbed at an equilibrium pressure of 5 mbar. To ensure good thermal conductivity, He (2 mbar) was added to the system before the introduction of CO or 15N2. Before use, 15N2, CO, and He were additionally purified by passing them through a liquid nitrogen trap.
The adsorption experiments were performed using the following gases and adsorbates: CO (Merck, Darmstadt, Germany, >99.5%), 15N2 (Sigma-Aldrich, St. Louis, MO, USA, 98 at. %), O2 (Messer, Bad Soden, Germany, 99.999%), He (Messer, Bad Soden, Germany, 99.999%), and D2O (Cambridge Isotope Laboratories, Inc., Tewksbury, MA, USA, 99.9%).
Theoretical modeling was based on periodic Density Functional Theory (DFT) calculations. We used Vienna Ab initio Simulation Package (VASP) [53,54,55] with PW91 [56] exchange-correlation functional and dispersion correction of D2 type [57]. The projector-augmented wave (PAW) method was used to describe the core-valence electron interactions and the calculations were performed in the Γ point only. We employed a plane wave basis with an energy cut-off of 415 eV. All atomic coordinates were optimized until the atomic forces became less than 2 × 10−4 eV/pm.
We employed four models of ceria, in which the distance between the ceria moieties in neighboring cells was at least 1.0 nm:
CeO2(111) surface slab model with a step—half of the surface three-layer is deleted. The unit cell consists of 60 Ce and 120 O atoms, with a size of 1.322 × 1.984 nm2 (α = 60°; β = γ = 90°).
CeO2(110) surface model, consisting of Ce96O192, as the cell parameters are a = 2.1900 nm, b = 1.5486 nm, c = 2.1678 nm, α = β = γ = 90°.
CeO2(100) surface model, consisting of Ce64O128, as the cell parameters are a = 2.1900 nm, b = 1.0950 nm, c = 2.1581 nm, α = β = γ = 90°. Due to the polarity of the surface, we created an O-terminated surface as half of the oxygen atoms were moved from the top surface to the bottom one.
Ce40O80 nanoparticle model is situated in a parallelepiped unit cell with the dimensions of 2.2 × 1.9 × 1.9 nm.
The binding energy of N2 (BE) to Ce4+ cations was calculated as the difference between the energy of the adsorption complex on the one hand, and the energies of the pristine ceria model and the isolated dinitrogen molecule on the other. With this definition, the Ce4+-N2 interaction is energetically favorable when the values are negative. The calculated frequencies for 14N2 are scaled by 0.9668, representing the ratio between the experimental Raman frequency, 2331 cm−1 [58], and the calculated value of the isolated 14N2, 2411 cm−1. The corresponding frequency values for 15N2 molecule were obtained using an isotope ratio of 1.035.

4. Conclusions

Dinitrogen appears to be a suitable IR probe molecule for testing the surface of ceria and ceria-based materials. To avoid any hindrance to gaseous CO2, it is recommended to use the 15N2 isotopologue. When adsorbed on stoichiometric ceria, 15N2 does not interact with the weakly acidic Ce4+ on the (111) planes, but forms well-defined species with the Ce4+ sites from the (110) and (100) planes, which are detected in the IR spectra at 2258–2257 and 2253–2252 cm−1, respectively. 15N2 also provides information about the protonic acidity of the OH/OD groups by shifting their maxima to lower frequencies. Note that CO fails to measure the acidity of the terminal OH groups because it causes a change in their binding mode.
Compared to CO as a probe molecule, N2 has some advantages. It does not reduce the Ce4+ cations and does not form anionic species (as carbonates formed by CO), which can modify the surface. 15N2 has a weak induction effect, and the maxima of the bands of adsorbed 15N2 are practically coverage-independent. Finally, dinitrogen can be used for selective determination of protonic acidity.

Supplementary Materials

Supplementary data to this article can be found online at https://www.mdpi.com/article/10.3390/molecules29153608/s1. Figure S1: IR spectra of 15N2 (different coverages) adsorbed at 100 K on CeO2-NC, pre-evacuated at 773 K; Figure S2: Adsorption complexes of N2 with Ce4+ cations on the CeO2(111) model; Figure S3: Electron density difference between the linear and bridge coordination of N2 on the (100) CeO2 surface.

Author Contributions

Conceptualization, K.I.H.; methodology, K.K.C. and G.N.V.; investigation, K.K.C., B.S.K. and I.Z.K.; writing—original draft preparation, M.Y.M.; writing—review and editing, K.K.C., M.Y.M., K.I.H., G.N.V. and H.A.A.; visualization, K.K.C., M.Y.M., B.S.K. and I.Z.K.; funding acquisition, M.Y.M. and K.I.H. All authors have read and agreed to the published version of the manuscript.

Funding

The experimental and partly the theoretical parts of this work were financially supported by the Bulgarian Science Fund (Project numbers KП-06-ДB-1/2021 and KΠ-06-H-59/5/2021).

Data Availability Statement

All data generated during this study are provided in the manuscript and in the Supporting information.

Acknowledgments

I.Z.K. is grateful to the European Union—NextGenerationEU, through the National Recovery and Resilience Plan of the Republic of Bulgaria, project No BG-RRP-2.004-0008 for the financial support. Thanks are also due to the Nestum facility of Sofia Tech Park for providing computational resources.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Trovarelli, A.; Fornasiero, P. (Eds.) Catalysis by Ceria and Related Materials, 2nd ed.; Imperial College Press: London, UK, 2013. [Google Scholar]
  2. Abdelouahab-Reddama, Z.; El Mail, R.; Coloma, F.; Sepúlveda-Escribano, A. Platinum supported on highly-dispersed ceria on activated carbon for the total oxidation of VOCs. Appl. Catal. A 2015, 494, 87–94. [Google Scholar] [CrossRef]
  3. Aneggi, E.; Boaro, M.; Leitenburg, C.D.; Dolcetti, G.; Trovarelli, A. Insights into the redox properties of ceria-based oxides and their implications in catalysis. J. Alloys Compd. 2006, 408, 1096–1102. [Google Scholar] [CrossRef]
  4. Trovarelli, A.; de Leitenburg, C.; Boaro, M.; Dolcetti, G. The utilization of ceria in industrial catalysis. Catal. Today 1999, 50, 353–367. [Google Scholar] [CrossRef]
  5. Vivier, L.; Duprez, D. Ceria-based solid catalysts for organic chemistry. ChemSusChem 2010, 3, 654–678. [Google Scholar] [CrossRef] [PubMed]
  6. Binet, C.; Daturi, M.; Lavalley, J.-C. IR study of polycrystalline ceria properties in oxidised and reduced states. Catal. Today 1999, 50, 207–225. [Google Scholar] [CrossRef]
  7. Mihaylov, M.Y.; Ivanova, E.Z.; Vayssilov, G.N.; Hadjiivanov, K.I. Revisiting ceria-NOx interaction: FTIR studies. Catal. Today 2020, 357, 613–620. [Google Scholar] [CrossRef]
  8. Agarwal, S.; Lefferts, L.; Mojet, B.L. Ceria nanocatalysts: Shape dependent reactivity and formation of OH. ChemCatChem 2013, 5, 479–489. [Google Scholar] [CrossRef]
  9. Cao, Y.; Zhao, L.; Gutmann, T.; Xu, Y.; Dong, L.; Buntkowsky, G.; Gao, F. Getting Insights into the influence of crystal plane effect of shaped ceria on its catalytic performances. J. Phys. Chem. C 2018, 122, 20402–20409. [Google Scholar] [CrossRef]
  10. Zhang, M.; Li, J.; Li, H.; Li, Y.; Shen, W. Morphology-dependent redox and catalytic properties of CeO2 nanostructures: Nanowires, nanorods and nanoparticles. Catal. Today 2009, 148, 179–183. [Google Scholar]
  11. Konsolakis, M.; Lykaki, M. Facet-dependent reactivity of ceria nanoparticles exemplified by CeO2-based transition metal catalysts: A critical review. Catalysts 2021, 11, 452. [Google Scholar] [CrossRef]
  12. Qiao, Z.-A.; Wu, Z.; Dai, S. Shape-controlled ceria-based nanostructures for catalysis applications. ChemSusChem 2013, 6, 1821–1833. [Google Scholar] [CrossRef] [PubMed]
  13. Spanier, J.E.; Robinson, R.D.; Zhang, F.; Chan, S.-W.; Herman, I.P. Size-dependent properties of CeO2-y nanoparticles as studied by Raman scattering. Phys. Rev. B 2001, 64, 245407. [Google Scholar] [CrossRef]
  14. Song, G.; Cheng, N.; Zhang, J.; Huang, H.; Yuan, Y.; He, X.; Luo, Y.; Huang, K. Nanoscale cerium oxide: Synthesis, biocatalytic mechanism, and applications. Catalysts 2021, 11, 1123. [Google Scholar] [CrossRef]
  15. Trovarelli, A.; Llorca, J. Ceria catalysts at nanoscale: How do crystal shapes shape catalysis? ACS Catal. 2017, 7, 4716–4735. [Google Scholar] [CrossRef]
  16. Wang, X.; Li, M.; Wu, Z. In situ spectroscopic insights into the redox and acid-base properties of ceria catalysts. Chin. J. Catal. 2021, 42, 2122–2140. [Google Scholar] [CrossRef]
  17. Wu, Z.; Mann, A.K.P.; Li, M.; Overbury, S.H. Spectroscopic investigation of surface-dependent acid–base property of ceria nanoshapes. J. Phys. Chem. C 2015, 119, 7340–7350. [Google Scholar] [CrossRef]
  18. Wu, Z.; Li, M.; Overbury, S.H. On the structure dependence of CO oxidation over CeO2 nanocrystals with well-defined surface planes. J. Catal. 2012, 285, 61–73. [Google Scholar] [CrossRef]
  19. Mai, H.-X.; Sun, L.-D.; Zhang, Y.-W.; Si, R.; Feng, W.; Zhang, H.-P.; Liu, H.-C.; Yan, C.-H. Shape-selective synthesis and oxygen storage behavior of ceria nanopolyhedra, nanorods, and nanocubes. J. Phys. Chem. B 2005, 109, 24380–24385. [Google Scholar] [CrossRef]
  20. Bezkrovnyi, O.S.; Kraszkiewicz, P.; Ptak, M.; Kepinski, L.; Bezkrovnyi, O.S.; Kraszkiewicz, P.; Ptak, M.; Kepinski, L. Thermally induced reconstruction of ceria nanocubes into zigzag {111}-nanofacetted structures and its influence on catalytic activity in CO oxidation. Catal. Commun. 2018, 117, 94–98. [Google Scholar] [CrossRef]
  21. Aneggi, E.; Llorca, J.; Boaro, M.; Trovarelli, A. Surface-structure sensitivity of CO oxidation over polycrystalline ceria powders. J. Catal. 2005, 234, 88–95. [Google Scholar] [CrossRef]
  22. Désaunay, J.; Bonura, G.; Chiodo, V.; Freni, S.; Couzinié, J.-P.; Bourgon, J.; Ringuedé, A.; Labat, F.; Adamo, C.; Cassir, M. Surface-dependent oxidation of H2 on CeO2 surfaces. J. Catal. 2013, 297, 193–201. [Google Scholar] [CrossRef]
  23. Kovacevic, M.; Agarwal, S.; Mojet, B.L.; van Ommen, J.G.; Lefferts, L. The effects of morphology of cerium oxide catalysts for dehydrogenation of ethylbenzene to styrene. Appl. Catal. A 2015, 505, 354–364. [Google Scholar] [CrossRef]
  24. Sudduth, B.; Yun, D.; Sun, J.; Wang, Y. Facet-dependent selectivity of CeO2 nanoparticles in 2-propanol conversion. J. Catal. 2021, 404, 96–108. [Google Scholar] [CrossRef]
  25. Hadjiivanov, K.I.; Vayssilov, G.N. Characterization of oxide surfaces and zeolites by carbon monoxide as an IR probe molecule. Adv. Catal. 2002, 47, 307–511. [Google Scholar]
  26. Hadjiivanov, K.I. Identification and characterization of surface hydroxyl groups by infrared spectroscopy. Adv. Catal. 2014, 57, 99–318. [Google Scholar]
  27. Wu, W.; Savereide, L.M.; Notestein, J.; Weitz, E. In-situ IR spectroscopy as a probe of oxidation/reduction of Ce in nanostructured CeO2. Appl. Surf. Sci. 2018, 445, 548–554. [Google Scholar] [CrossRef]
  28. Afrin, S.; Bollini, P. On the utility of Ce3+ spin-orbit transitions in the interpretation of rate data in ceria catalysis: Theory, validation, and application. J. Phys. Chem. C 2023, 127, 234–247. [Google Scholar] [CrossRef]
  29. Mihaylov, M.Y.; Ivanova, E.Z.; Aleksandrov, H.A.; Petkov, P.S.; Vayssilov, G.N.; Hadjiivanov, K.I. FTIR and density functional study of NO interaction with reduced ceria: Identification of N3 and NO2− as new intermediates in NO conversion. Appl. Catal. B 2015, 176–177, 107–119. [Google Scholar] [CrossRef]
  30. Mihaylov, M.Y.; Ivanova, E.Z.; Aleksandrov, H.A.; Petkov, P.S.; Vayssilov, G.N.; Hadjiivanov, K.I. Formation of N3 during interaction of NO with reduced ceria. Chem. Commun. 2015, 51, 5668–5671. [Google Scholar] [CrossRef]
  31. Li, H.; Zhang, P.; Li, G.; Lu, J.; Wu, Q.; Gu, Y. Stress measurement for nonstoichiometric ceria films based on Raman spectroscopy. J. Alloys Compd. 2016, 682, 132–137. [Google Scholar] [CrossRef]
  32. Bozon-Verduraz, F.; Bensalem, A. IR studies of cerium dioxide: Influence of impurities and defects. J. Chem. Soc. Faraday Trans. 1994, 90, 653–657. [Google Scholar] [CrossRef]
  33. Vassileva, E.; Varimezova, B.; Hadjiivanov, K. Column solid-phase extraction of heavy metal ions on a high surface area CeO2 as a pre-concentration method for trace determination. Anal. Chim. Acta 1996, 336, 141–150. [Google Scholar] [CrossRef]
  34. Chakarova, K.; Zdravkova, V.; Karapenchev, B.; Nihtianova, D.; Ivanova, E.; Aleksandrov, H.; Koleva, I.; Panayotov, D.; Mihaylov, M.; Vayssilov, G.; et al. Evolution of Ce4+ Lewis acidity during dehydroxylation of ceria nanoparticles with different morphology: An integrated FTIR, DFT and HRTEM study. J. Catal. 2024, 433, 115463. [Google Scholar] [CrossRef]
  35. Neyman, K.M.; Strodel, P.; Ruzankin, S.P.; Schlensog, N.; Knözinger, H. N2 and CO molecules as probes of zeolite acidity: An infrared spectroscopy and density functional investigation. Catal. Lett. 1995, 31, 273–285. [Google Scholar] [CrossRef]
  36. Geobaldo, F.; Lamberti, C.; Ricchiardi, G.; Bordiga, S.; Zecchina, A.; Turnes Palomino, G.; Otero Arean, C. N2 Adsorption at 77 K on H-mordenite and alkali-metal-exchanged mordenites: An IR study. J. Phys. Chem. 1995, 99, 11167–11177. [Google Scholar] [CrossRef]
  37. Larin, A.V.; Vercauteren, D.P.; Lamberti, C.; Bordiga, S.; Zecchina, A. Interaction between probe molecules and zeolites. Part II: Interpretation of the IR spectra of CO and N2 adsorbed in NaY and NaRbY. Phys. Chem. Chem. Phys. 2002, 4, 2424–2433. [Google Scholar] [CrossRef]
  38. Sakata, Y.; Kinoshita, N.; Domen, K.; Onishi, T.J. Infrared studies on dinitrogen and dihydrogen adsorbed over TiO2, at low temperatures. Chem. Soc. Faraday Trans. I 1987, 83, 2765–2772. [Google Scholar] [CrossRef]
  39. Valenzano, L.; Civalleri, B.; Chavan, S.; Palomino, G.T.; Arean, C.O.; Bordiga, S. Computational and experimental studies on the adsorption of CO, N2, and CO2 on Mg-MOF-74. J. Phys. Chem. C 2010, 114, 11185–11191. [Google Scholar] [CrossRef]
  40. Wakabayashi, F.; Kondo, J.N.; Domen, K.; Hirose, C. Direct comparison of N2 and CO as IR-spectroscopic probes of acid sites in H-ZSM-5 zeolite. J. Phys. Chem. 1995, 99, 10573–10580. [Google Scholar] [CrossRef]
  41. Zecchina, A.; Otero Arean, C.; Turnes Palomino, G.; Geobaldo, F.; Lamberti, C.; Spoto, G.; Bordiga, S. The vibrational spectroscopy of H2, N2, CO and NO adsorbed on the titanosilicate molecular sieve ETS-10. Phys. Chem. Chem. Phys. 1999, 1, 1649–1657. [Google Scholar] [CrossRef]
  42. Chakarova, K.; Hadjiivanov, K. FTIR study of N2 and CO adsorption on H-D-FER. Micropor. Mesopor. Mater. 2013, 177, 59–65. [Google Scholar] [CrossRef]
  43. Chakarova, K.; Andonova, S.; Dimitrov, L.; Hadjiivanov, K. FTIR study of CO and N2 adsorption on [Ge]FAU zeolites in their Na- and H-forms. Micropor. Mesopor. Mater. 2016, 220, 188–197. [Google Scholar] [CrossRef]
  44. Aneggi, E.; Wiater, D.; Leitenburg, C.; Llorca, J.; Trovarelli, A. Shape-dependent activity of ceria in soot combustion. ACS Catal. 2014, 4, 172–181. [Google Scholar] [CrossRef]
  45. Yang, C.; Capdevila-Cortada, M.; Dong, C.; Zhou, Y.; Wang, J.; Yu, X.; Nefedov, A.; Heißler, S.; López, N.; Shen, W.; et al. Surface refaceting mechanism on cubic ceria. J. Phys. Chem. Lett. 2020, 11, 7925–7931. [Google Scholar] [CrossRef] [PubMed]
  46. Chakarova, K.; Drenchev, N.; Mihaylov, M.; Hadjiivanov, K. Interaction of O2 with reduced ceria nanoparticles at 100–400 K: Fast oxidation of Ce3+ ions and dissolved H2. Catalysts 2024, 14, 45. [Google Scholar] [CrossRef]
  47. Sun, C.; Li, H.; Chen, L. Nanostructured ceria-based materials: Synthesis, properties, and applications. Energy Environ. Sci. 2012, 5, 8475–8505. [Google Scholar] [CrossRef]
  48. Gill, L.; Beste, A.; Chen, B.; Li, M.; Mann, A.K.P.; Overbury, S.H.; Hagaman, E.W. Fast MAS 1H NMR study of water adsorption and dissociation on the (100) surface of ceria nanocubes: A fully hydroxylated, hydrophobic ceria surface. J. Phys. Chem. C 2017, 121, 7450–7465. [Google Scholar] [CrossRef]
  49. Farra, R.; Wrabetz, S.; Schuster, M.E.; Stotz, E.; Hamilton, N.G.; Amrute, A.P.; Pérez-Ramírez, J.; López, N.; Teschner, D. Understanding CeO2 as a Deacon catalyst by probe molecule adsorption and in situ infrared characterisations. Phys. Chem. Chem. Phys. 2013, 15, 3454–3465. [Google Scholar] [CrossRef]
  50. Trombetta, M.; Busca, G.; Rossini, S.; Piccoli, V.; Cornaro, U.; Guercio, A.; Catani, R.; Willey, R.J. FT-IR studies on light olefin skeletal isomerization catalysis III. Surface acidity and activity of amorphous and crystalline catalysts belonging to the SiO2-Al2O3 system. J. Catal. 1998, 179, 581–596. [Google Scholar] [CrossRef]
  51. Henderson, M.A.; Perkins, C.L.; Engelhard, M.H.; Thevuthasan, S.; Peden, C.H.F. Redox properties of water on the oxidized and reduced surfaces of CeO2(111). Surf. Sci. 2003, 526, 1–18. [Google Scholar] [CrossRef]
  52. Zabilskiy, M.; Djinovic, P.; Tchernychova, E.; Tkachenko, O.P.; Kustov, L.M.; Pintar, A. Nanoshaped CuO/CeO2 materials: Effect of the exposed ceria surfaces on catalytic activity in N2O decomposition reaction. ACS Catal. 2015, 5, 5357–5365. [Google Scholar] [CrossRef]
  53. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef] [PubMed]
  54. Version VASP.5.3. Available online: https://www.vasp.at/ (accessed on 29 July 2024).
  55. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  56. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687, Erratum in Phys. Rev. B 1993, 48, 4978. [Google Scholar] [CrossRef] [PubMed]
  57. Grimme, S.J. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. Comp. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  58. Rasetti, F. Incoherent scattered radiation in diatomic molecules. Phys. Rev. 1929, 34, 367. [Google Scholar] [CrossRef]
Figure 1. FTIR spectra of 15N2 and CO adsorbed at 100 K on CeO2-NC pre-evacuated at different temperatures, as marked on the spectra. (A) Adsorbed 15N2 at maximum coverage (10 mbar equilibrium pressure). (B) Adsorbed CO at intermediate coverage (for details, see text). The region of the carbonyl bands corresponding to 15N2 bands is highlighted.
Figure 1. FTIR spectra of 15N2 and CO adsorbed at 100 K on CeO2-NC pre-evacuated at different temperatures, as marked on the spectra. (A) Adsorbed 15N2 at maximum coverage (10 mbar equilibrium pressure). (B) Adsorbed CO at intermediate coverage (for details, see text). The region of the carbonyl bands corresponding to 15N2 bands is highlighted.
Molecules 29 03608 g001
Figure 2. FTIR spectra of 15N2 and CO adsorbed at 100 K on differently shaped ceria nanoparticles pre-evacuated at 773 K. (A) Adsorbed 15N2 at maximum coverage (10 mbar equilibrium pressure). (B) Adsorbed CO at intermediate coverage (for details, see text). The region of the carbonyl bands corresponding to the 15N2 bands is highlighted. The corresponding second derivative is shown above the spectrum.
Figure 2. FTIR spectra of 15N2 and CO adsorbed at 100 K on differently shaped ceria nanoparticles pre-evacuated at 773 K. (A) Adsorbed 15N2 at maximum coverage (10 mbar equilibrium pressure). (B) Adsorbed CO at intermediate coverage (for details, see text). The region of the carbonyl bands corresponding to the 15N2 bands is highlighted. The corresponding second derivative is shown above the spectrum.
Molecules 29 03608 g002
Figure 3. Optimized complexes of an N2 molecule in (a) bridge and (b) top coordination to the CeO2(100) surface model. (c) A full monolayer of N2 molecules adsorbed in bridge coordination to the CeO2(100) surface model. CeO2(100) model was created with surface O centers moved along the “100” direction. Color coding: Ce—yellow, O—red, and N—blue.
Figure 3. Optimized complexes of an N2 molecule in (a) bridge and (b) top coordination to the CeO2(100) surface model. (c) A full monolayer of N2 molecules adsorbed in bridge coordination to the CeO2(100) surface model. CeO2(100) model was created with surface O centers moved along the “100” direction. Color coding: Ce—yellow, O—red, and N—blue.
Molecules 29 03608 g003
Figure 4. Optimized complexes of an N2 molecule in (a) bridge and (b) top coordination to the CeO2(110) surface model. Color coding: Ce—yellow, O—red, and N—blue.
Figure 4. Optimized complexes of an N2 molecule in (a) bridge and (b) top coordination to the CeO2(110) surface model. Color coding: Ce—yellow, O—red, and N—blue.
Molecules 29 03608 g004
Figure 5. FTIR spectra in the OD region, registered after adsorption of 15N2 (A) and CO (B) on CeO2-NC pre-evacuated at 473 K (upper spectra a–c) and 773 K (lower spectra d–f): spectra taken immediately after adsorption (a, d), after a certain time (b, e), and after short evacuation (c, f).
Figure 5. FTIR spectra in the OD region, registered after adsorption of 15N2 (A) and CO (B) on CeO2-NC pre-evacuated at 473 K (upper spectra a–c) and 773 K (lower spectra d–f): spectra taken immediately after adsorption (a, d), after a certain time (b, e), and after short evacuation (c, f).
Molecules 29 03608 g005
Table 1. Some basic characteristics of the samples studied.
Table 1. Some basic characteristics of the samples studied.
SampleParticle ShapeAverage Particle Size [nm] *SBET [m2 g−1]Pore
Volume
[cm3 g−1]
Average Pore
Diameter
[nm]
1CeO2-NCcubes27.2310.1722
2CeO2-NPpolyhedra29.2200.1926
3CeO2-NRrods6.31100.4617
* Average particle size determined according to the Scherrer’s equation.
Table 2. Binding energies (in eV), interatomic distances (in pm), and calculated vibrational frequencies (in cm−1) of N2 molecule adsorbed on Ce4+ ions of various ceria-based systems: Ce40O80 nanoparticle, CeO2(100) (with checkered (ch) and linear arrangements of the surface oxygen centers), CeO2(110), and CeO2(111) surface model with step. The shift with respect to isolated N2 molecule, Δν, is also shown.
Table 2. Binding energies (in eV), interatomic distances (in pm), and calculated vibrational frequencies (in cm−1) of N2 molecule adsorbed on Ce4+ ions of various ceria-based systems: Ce40O80 nanoparticle, CeO2(100) (with checkered (ch) and linear arrangements of the surface oxygen centers), CeO2(110), and CeO2(111) surface model with step. The shift with respect to isolated N2 molecule, Δν, is also shown.
StructureBEd(Ce-N)d(N-N)Δd(N-N)ν(14N2)ν(15N2)Δν
Ce40O80 apex top−0.30309111.0−0.1233322542
CeO2(100)ch—bridge−0.45309111.0−0.1233122520
CeO2(100)—bridge−0.43312; 313111.0−0.1233222531
CeO2(110)—top−0.30304111.0−0.1234022619
CeO2(111)_edge-top−0.31304110.9−0.2234022619
CeO2(111)—top−0.27 ÷ −0.26297111.0−0.1233422553
Table 3. Summary of binding energies (in eV) and infrared bands (in cm−1) of CO and 15N2 probe molecules characteristic of different ceria surfaces.
Table 3. Summary of binding energies (in eV) and infrared bands (in cm−1) of CO and 15N2 probe molecules characteristic of different ceria surfaces.
StructureBE(CO) aBE(N2)ν(CO)expν(CO)calc aν(15N2)expν(15N2)calcν(14N2)exp. b
Apex—top−0.39−0.3021752183225822542337
Edge (111)—top−0.40−0.3121752180225822612337
Ce(100)—bridge−0.62−0.4321622155225322532332
Ce(110)—top−0.37−0.3021712165225722612336
Ce(111)—top−0.38−0.2721552149-2255-
a Data from Ref. [34]. b Calculated on the basis of the experimental values for ν(15N2).
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

Chakarova, K.K.; Mihaylov, M.Y.; Karapenchev, B.S.; Koleva, I.Z.; Vayssilov, G.N.; Aleksandrov, H.A.; Hadjiivanov, K.I. N2 as an Efficient IR Probe Molecule for the Investigation of Ceria-Containing Materials. Molecules 2024, 29, 3608. https://doi.org/10.3390/molecules29153608

AMA Style

Chakarova KK, Mihaylov MY, Karapenchev BS, Koleva IZ, Vayssilov GN, Aleksandrov HA, Hadjiivanov KI. N2 as an Efficient IR Probe Molecule for the Investigation of Ceria-Containing Materials. Molecules. 2024; 29(15):3608. https://doi.org/10.3390/molecules29153608

Chicago/Turabian Style

Chakarova, Kristina K., Mihail Y. Mihaylov, Bayan S. Karapenchev, Iskra Z. Koleva, Georgi N. Vayssilov, Hristiyan A. Aleksandrov, and Konstantin I. Hadjiivanov. 2024. "N2 as an Efficient IR Probe Molecule for the Investigation of Ceria-Containing Materials" Molecules 29, no. 15: 3608. https://doi.org/10.3390/molecules29153608

APA Style

Chakarova, K. K., Mihaylov, M. Y., Karapenchev, B. S., Koleva, I. Z., Vayssilov, G. N., Aleksandrov, H. A., & Hadjiivanov, K. I. (2024). N2 as an Efficient IR Probe Molecule for the Investigation of Ceria-Containing Materials. Molecules, 29(15), 3608. https://doi.org/10.3390/molecules29153608

Article Metrics

Back to TopTop