Pengfei
Yang
a,
Mark
Douthwaite
b,
Jiahao
Pan
a,
Lirong
Zheng
c,
Song
Hong
a,
David J.
Morgan
b,
Mingyu
Gao
a,
Dianqing
Li
a,
Junting
Feng
*a and
Graham J.
Hutchings
*b
aState Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing, People's Republic of China. E-mail: fengjt@buct.edu.cn
bMax Planck- Cardiff Centre on the Fundamentals of Heterogeneous Catalysis FUNCAT, Cardiff Catalysis Institute, School of Chemistry, Cardiff University, Main Building, Park Place, Cardiff, CF10 3AT, UK. E-mail: hutch@cardiff.ac.uk
cInstitute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, Beijing, People's Republic of China
First published on 7th June 2021
A series of TiO2 anatase materials were synthesized with varying proportions of exposed (101) and (001) facets. Pt was subsequently impregnated onto these materials, which were then exposed to either an aerobic or reductive thermal treatment. The final catalysts were assessed for their performance in the aerobic, solvent free oxidation of octanol. Many kinds of characterization methods were employed to gain insight on how the exposed support facets influenced the catalysis. Thermal reduction of the catalysts led to a loss in the support surface structure, but in the contrary, after calcination, the defined facets remained, which was revealed by Raman and XRD. We further provide evidence that the coordinately unsaturated O2c–Ti5c–O2c sites present in the support promoted the dispersion of Pt after calcination in air, leading to the formation of low-coordinated and electron-deficient Pt clusters, based on high resolution- and spherical- aberration corrected transmission electron microscopy and X-ray absorption fine structure analysis. The calcined catalysts were far more active for both the oxidation of octanol and intermediate octanal, thus resulting in a higher selectivity of octanoic acid. Catalyst performance closely correlated with the proportion of low-coordinated Pt species and the specific surface coordination structure of the support in the catalyst. Subsequent in situ Fourier transfer infrared experiments, conducted using alcohol and aldehydic probe molecules, confirmed that substrate adsorption was strongly influenced by the low-coordinated sites of both the Pt cluster and the exposed surface facets on the support.
Many previous studies have investigated the use of supported metal catalysts for liquid phase alcohol oxidations, under both aqueous and solvated conditions.6 Examples of alcohol oxidation reactions carried out in the absence of solvent however, are scarcer by comparison. This is particularly true for the oxidation of large aliphatic alcohols, such as octanol,7–10 decanol11,12 and 1-dodecanol.11,13 Previously, researchers have employed noble metal supported catalysts for the oxidation of octanol,7–10 but in the vast majority of cases, only modest activities are observed. Although no clear evidence has been demonstrated eluding to why the oxidation of such substrates is so challenging, it is likely to be attributed to both steric hindrance from the alkyl chain and electronic donation from the alkyl chain to the alcohol moiety, making activation of the alcohol more difficult. The rate of oxidation then, is likely to be significantly influenced by the nature of the catalytic active sites.
Over supported metal heterogeneous catalysts, alcohol oxidation proceeds through several sequential steps. First, the alcohol group is deprotonated on the catalyst surface, forming a metal-alkoxide intermediate. This is followed by β-hydride elimination of a hydrogen atom from the α-carbon in the alkoxy intermediate, leaving a surface metal hydride and the associated surface bound formyl intermediate. The latter can either desorb or undergo sequential oxidation to the associated carboxylic acid via geminal diol dehydration.14 Oxidation of the metal hydride with O2, leads to the formation of water, thus, completing the catalytic cycle. The ease in which the preceding steps occur over a specific catalyst, is what ultimately dictates the catalytic performance observed.7,15 Given that electronic donation from the alkyl chain is likely to inhibit activation in large aliphatic alcohols, it is particularly important to develop catalysts which can effectively promote both the alcohol dissociation and β-hydride elimination.
Previously, it has been demonstrated that the morphology of supported metal particles can have a dramatic influence on their catalytic reactivity.16–20 Liu et al.21 previously determined that Pd {111} facets were more effective for the oxidation of 5-hydroxymethyl-2-furfural, than Pd {100} facets. Furthermore, the {111} facet was determined to be highly selective for the oxidation of the intermediate product, 5-hydroxymethyl-2-furancarboxylic acid, to 5-formyl-2-furancarboxylic acid. This was attributed to a reduced energy barrier for this process, proposed to stem from the superior ability of O2 to reduce to peroxide on these sites. In addition to particle morphology, some works have highlighted that exposed facets on the catalyst support can also influence reactivity in catalytic reactions. We recently demonstrated that crystal-plane support effects in Au1Pt3/TiO2 catalysts influenced reaction selectivity in aerobic glycerol oxidation.22 We determined that exposed (001) facets on the TiO2 support promoted the adsorption of the aldehyde functionalities, resulting in a higher reaction selectivity to carboxylic acid products.
From a microstructure perspective, facet-dependent performance is derived from differences in the coordination and arrangement of surface atoms.23 In addition to influencing the adsorption behavior of reactants, the coordination structure of support could have further impact on the catalytic performance by affecting the properties of the supported metals. For example, Liu et al.17 previously demonstrated that a strong electronic interaction existed between Au nanoparticles and Ti5c species on TiO2 (100) facets, promoting activity for the oxidation of CO. Our group also observed a similar effect when investigating the interaction between CeO2 (100) facets and Au1Pt3 nanoparticles.25 We determined that there was a greater electron transfer between the coordinately saturated Ce6c–O2c sites on the support and the Pt component, which promoted glycerol oxidation. Despite the progress made in this field, understanding how different support coordination structures influence the geometric structure of the supported metal particles remains unclear.
Given that the oxidation of large aliphatic alcohols is likely to be inhibited by both steric effects and electron donation from the aliphatic alkyl chain, we consider that such reactions are likely to be highly sensitive to catalyst structure. We therefore synthesize a series of Pt/TiO2 catalysts, the supports of which, possessed different proportions of exposed (101) and (001) facets, to reveal the effects of the coordination structure of supports on the geometric structure of the supported metal particles. Besides, the catalysts were extensively characterized and tested for the aerobic, solvent free oxidation of octanol, our model large aliphatic substrate. The nature of the exposed surface facets were determined to have a profound effect on both the rate of octanol conversion and the reaction selectivity. We also assessed how the atmosphere used for the thermal treatment of the catalyst, prior to employment in the reaction, influenced performance. Exposing these catalysts to a thermal reductive treatment resulted in a significant loss in the supports surface structure, which dramatically influenced the performance of the catalysts towards octanol oxidation. While, undergoing thermal treatment in air, the catalysts well maintained the original coordination structure of the support, which is considered the precondition for affecting the metal geometric structure.
![]() | ||
Fig. 1 XRD patterns (A) and Raman spectra (B) for TiO2 nanocrystals enclosed by different facets. The geometric model and the coordination structure of (C) (101) facet and (D) (001) facet. |
Illustrative representations of the TiO2 (101) and (001) surface structures are presented in Fig. 1(C) and (D), respectively. With the TiO2 (101) facet, there are two types of exposed Ti–O coordination structures, both of which are of equal proportions. One of these has 3 O atoms coordinated to 6 Ti atoms (denoted as O3c–Ti6c) and the other, has 2 O atoms coordinated to 5 Ti atoms (denoted as O2c–Ti5c). In the contrary, with the (001) TiO2, all of the exposed Ti–O coordination structures exist as O2c–Ti5c.20 Raman spectroscopy was subsequently employed to further assess the surface structure of these materials. This technique is known to be highly sensitive and can provide information on molecular bonding from a micro-perspective. As shown in Fig. 1B, all the samples exhibit similar peak positions at 144 cm−1, 394 cm−1, 514 cm−1 and 636 cm−1, which can be assigned to Eg, B1g, A1g and Eg Raman vibration modes, respectively.27 It has been reported previously that Eg, B1g and A1g modes are attributed to a symmetric stretching vibration, a symmetric bending vibration, and an asymmetric bending vibration of O–Ti–O, respectively. With more (001) facet exposed, the relative intensity of the B1g (394 cm−1) and A1g modes (514 cm−1) to the Eg mode (636 cm−1), increases. This indicates that there is a higher proportion of symmetric and asymmetric bending on the surface of (001) facet. This, in turn, can be attributed to an increase in unsaturated O2c–Ti5c–O2c sites on the surface of (001) facet, while both saturated O3c–Ti6c and unsaturated O2c–Ti5c sites are present on the surface of (101) facet.
Next, Pt was impregnated onto the different TiO2 materials and subsequently calcined to produce a series of 1 wt% Pt/TiO2 catalysts. The actual loading of Pt in the series catalysts were characterized by ICP and listed in Table S1.† All the catalysts has the similar Pt loading which is close to 1%. Furthermore, BET surface area measurements of the TiO2 support material and the series Pt/TiO2 are also listed in Table ESI.† The morphology of the supports and supported metal particles were subsequently probed by high resolution transmission electron microscopy (HRTEM) and spherical aberration correction electron microscopy (Cs-TEM). As shown in Fig. 2, the obtained TiO2 support has an obvious disparity in the arrangement of surface atoms. The lattice fringes of 0.24 and 0.35 nm, with and angle of 68.3°, are observed in Fig. 2A and can be assigned to [001] and [101] directions. This confirms that the surface of the anatase (101) facet is exposed, because of the octahedral structure of anatase. In Fig. 2B, two perpendicular lattice fringes of 0.19 and 0.38 nm, are observed and can be assigned to [100] and [010] (both belong to the <100> family of directions). The orthogonal lattice fringes confirm that this image is representative of a ‘top view’ of the synthesized TiO2, whereby, the (001) facet is highly exposed on the top surface of the materials.28–30 This aligns with the data retrieved from characterization of the TiO2 supports by Raman spectroscopy and XRD. Although the calcined catalysts, with different facets exposed, possess a similar size of supported nanoparticles (ca. 1.7 nm in diameter); determined by HRTEM (Fig. S2†), the results from Cs-TEM also shows that many smaller Pt clusters were also present on the surface of TiO2 support. As such, it suggests that the Pt particles supported on the (001) facet are smaller than those supported on the (101) facet. This indicates that the Pt atoms may have a stronger interaction with the coordinatively unsaturated sites of the (001) facet.
![]() | ||
Fig. 2 Cs-TEM images of calcined catalysts of Pt/TiO2-101-air (A and a) and Pt/TiO2-001-air (B and b). |
In addition to calcination, the series of catalyst precursors were also exposed to a thermal reduction. Compared with the Pt cluster formed in calcined catalysts, large Pt nanoparticles are obtain in the reductive atmosphere under the same temperature, as shown in Fig. S3.† Furthermore, the mean size of Pt nanoparticles on (001) facet is 3.24 nm, which is slightly bigger than that on (101) facet (2.50 nm). Previous studies have shown that exposing such precursors to either an aerobic or reductive heat treatment can significantly influence the properties of the resultant catalysts.31,32 Reductive thermal treatments in particular can lead to SMSI effects, which dramatically influence the morphology at the metal–support interface.33 To establish how the thermal treatment influenced the electronic structure of these catalysts, each series of catalysts (both after reduction or calcination) was probed by X-ray photoelectron spectroscopy (XPS).
In Fig. 3(A) and (B), a peak is observed at ca. 70–80 eV, which is labelled as ‘background loss’. This peak is attributed to energy loss structure of the Ti 3s state, and commonly observed in the Pt 4f region when analysing XPS spectra of Pt supported titania catalysts.34–36 From inspection of the Pt 4f region, it is evident that there are two dominant Pt states are observed, indicative of metallic platinum; observed at ca. 70.1 eV and 73.4 eV, assigned to Pt0 4f7/2 and 4f5/2, and oxidized Pt at ca. 71.8 eV and 75.1 eV, assigned to the Pt2+ 4f572 and 4f5/2 states respectively. As expected, metallic Pt species are dominant in each of the catalysts after thermal reduction in H2 at 500 °C, whilst Pt2+ species are dominant in spectra associated with the calcined catalysts. Besides, the proportion of Pt2+ species is also increased from 63% to 82% when the support possesses a higher proportion of exposed anatase (001) facet, as shown in Fig. S4.† It is plausible to suggest that the Pt component has a stronger interaction with (001) facet on which plenty of low-coordinated O atoms are exposed and show more electron-deficient over (001) facet after calcination.
![]() | ||
Fig. 3 XPS of spectra series Pt/TiO2 with different facets exposed (Pt 4f, calcined (A); Pt 4f, reduced (B);). |
ESI† data, which provided information on the charge and local structure of Pt, was collected by X-ray absorption fine structure (XAFS) at the Pt L3-edge (11564 eV) in fluorescence mode. The XANES region can provide information on the oxidation state, through analysis of the white line intensity. In general, the white line intensity in the Pt L3-edge XANES spectra is dependent on the density of unoccupied states, and often, is used to establish the oxidation state of Pt.38,45,46 As shown in Fig. 4A, the white line of the calcined Pt/TiO2 catalyst has a higher intensity relative to the reduced catalysts. Furthermore, the Pt/TiO2-001-air sample exhibits the highest intensity among the calcined catalysts, indicating that this material possesses the greatest proportion of Pt2+, which aligns with the results from the XPS analysis. EXAFS was further utilized to obtain structural parameters, such as atom coordination numbers and atomic spacing. Analogous k3-weighed Fourier-transformed EXAFS spectra are shown in Fig. 4B, and the associated fitting parameters are listed in Table 2. The results show that the calcined Pt/TiO2 catalysts clearly possess Pt–O bonds, with a coordination number of ca. 2.7. The intensity at this radial distribution also correlated with an increase in the proportion of the (001) facet in the support of the calcined catalysts. A peak which is characteristic of Pt–Pt coordination, centred at ca. 2.7 Å, is observed in the radial distribution functions of all the catalysts. However, Pt–Pt bonding is clearly influenced by the atmosphere used for the catalysts thermal treatment and the proportion of (001) facet in the TiO2 support. As the proportion of the exposed (001) facet increases, the intensity of peak indicative of Pt–Pt bonding reduces. The associated coordination number also decreases from ca. 5.2 to 3.6 as the proportion of exposed (001) facet increases in the calcined catalysts. Interestingly however, no difference is observed in the Pt–Pt coordination number in the reduced catalyst, regardless of the proportion of exposed facets in the support material. The differences in Pt–Pt coordination number in the calcined catalyst further underlines that different sized Pt particles are present in these catalysts; a higher proportion of exposed (001) facet evidently promotes Pt dispersion. This in agreement with the results acquired from our Cs-TEM experiments, discussed previously.
![]() | ||
Fig. 4 (A) XANES spectra and (B) the radial distribution functions (RDFs) of the calcined and reduced Pt/TiO2 with different anatase facets exposed. |
Sample | Shell | Na | R (Å) | σ 2 (Å2·10−3) | ΔE0d (eV) | R factor (%) |
---|---|---|---|---|---|---|
a N: coordination numbers; b R: bond distance; c σ 2: Debye–Waller factors; d ΔE0: the inner potential correction. R factor: goodness of fit. S02 were set as 0.84/0.912 for Pt–O/Pt–Pt, which were obtained from the experimental EXAFS fit of reference PtO2/Pt foil by fixing CN as the known crystallographic value and was fixed to all the samples. | ||||||
Pt/TiO2-101-air | Pt–O | 2.7 | 1.99 | 4.3 | 2.2 | 1.0 |
Pt–Pt | 5.2 | 2.76 | 4.3 | 7.6 | ||
Pt/TiO2-101-001-air | Pt–O | 2.6 | 1.98 | 1.8 | 9.6 | 1.1 |
Pt–Pt | 4.5 | 2.77 | 3.7 | 0.5 | ||
Pt/TiO2-001-air | Pt–O | 2.8 | 2.01 | 2.3 | 9.4 | 1.2 |
Pt–Pt | 3.6 | 2.76 | 4.0 | 8.0 | ||
Pt/TiO2-101-H2 | Pt–Pt | 10.0 | 2.73 | 5.3 | 4.1 | 0.7 |
Pt/TiO2-101-001-H2 | Pt–Pt | 10.2 | 2.75 | 5.1 | 6.6 | 0.5 |
Pt/TiO2-001-H2 | Pt–Pt | 10.2 | 2.74 | 5.8 | 5.3 | 0.7 |
From the comprehensive characterization of these catalysts, it's clear that the exposed TiO2 facets have a dramatic influence on both the geometry and electronic structure of the supported Pt particles. Evidently, increasing the proportion of exposed (001) facets in the support increases the quantity of small, electron deficient Pt clusters after calcination. As such, we propose that the unsaturated O2c–Ti5c–O2c sites, which are the sites situated on the exposed surface of the (001) facets, are responsible for the strong interactions between the support and the Pt component.
To assess what changes to the surface structure occurred, through exposing the catalysts to either aerobic or reductive heat treatments, a series of additional Raman and FTIR spectroscopy experiments were conducted (Fig. 5).
From the Raman spectroscopy experiments (Fig. 5A) it is clear that after Pt loading and calcination, there are no notable changes to the exposed surface facets. On the contrary, after Pt loading and a reductive thermal treatment, all the vibrational peaks exhibit a blue shift. Furthermore, the relative intensity of the Eg (636 cm−1) to A1g (514 cm−1) vibration modes changes; the ratio of Eg to A1g increases after reduction, which is likely indicative of changes in the surface coordination structure, promoted through Pt reduction under a H2 atmosphere.
Further information on the geometric nature of Pt in the catalysts was acquired from conducting a series of in situ CO-FTIR experiments (Fig. 5B). It is important to note that all the reduced Pt/TiO2 catalysts (Pt/TiO2–H2) exhibit a strong IR adsorption due to the formation of Ti3+ species during the reductive heat treatment.40–43 Thus, to amplify the signal observed with these samples, 3 mg of the Pt/TiO2-001-H2 catalysts were mixed with 57 mg KBr, pelleted before conducting the experiment. Due to the fact that a comparatively small amount of catalyst was used for these experiments, the signal observed for the Pt/TiO2–H2 catalysts was significantly lower than that observed over the analogous calcined catalysts. For the calcined catalysts, 60 mg of pure catalyst was pelleted and used in the experiment. However, after amplification of the signal in the corresponding spectrum, similar features are observed to those in the Pt/TiO2-101-H2 catalysts spectrum. For the thermally reduced catalysts, Pt/TiO2-101-H2 and Pt/TiO2-001-H2, the peak observed at ca. 2065 cm−1 is characteristic of linear CO adsorption on flat Pt surfaces. The peak observed at ca. 2082 cm−1 is characteristic of linear CO adsorption on low coordinate Pt corner or terrace sites.37–39,44 No other evidence of CO adsorbed to Pt sites were observed over these catalysts, which suggests that both these materials are comprised of Pt particles with similar geometric structures. Interestingly however, the adsorption bands observed over the analogous calcined catalysts are quite different. Over these catalysts, the bands characteristic of linear CO adsorption on flat and edge sites undergoes a red-shift to 2067 cm−1 and 2051 cm−1, indicating that the Pt species in the calcined catalysts are more electron deficient than the analogous Pt species in the reduced catalysts; an observation which is in agreement with our findings from the XPS experiments. It is also evident from these experiments that the calcined Pt/TiO2-001-air catalyst possesses a higher proportion of linearly adsorbed CO on edge sites to flat sites, than the analogous calcined Pt/TiO2-101-air catalyst. This suggests that the Pt/TiO2-101-air catalyst possesses a higher proportion of coordination unsaturated Pt sites, which must be formed on the surface of TiO2 (001). The lower electron density and higher proportion of co-ordinately unsaturated sites are indicative of a strong interaction between the Pt cluster and anatase (001) facet, suggesting that that coordinative unsaturated site O2c–Ti5c–O2c may have a pivotal role in catalytic performance of this material.
To establish whether this was the case, and to gain further insight on the influence of the atmosphere used for the thermal treatment, the series of materials were employed as catalysts for the oxidation of 1-octanol (Fig. 6). These experiments were all conducted under solvent-free conditions.
Interestingly, all the catalysts exposed to an aerobic thermal heat treatment were notably more active than the analogous catalysts which were thermally reduced (Fig. 6A). Given that the XPS and EXAFS data confirmed that the calcined catalysts comprised of significantly more electron-deficient Pt cluster, its logical to propose that this species is more affective for alcohol oxidation than Pt0. The rate of octanol conversion was also different over the series of calcined catalysts (solid lines in Fig. 6A), indicating that the exposed support facets influenced reactivity. The rate of reaction was found to correlate with the proportion of exposed (001) in the support. This further confirms that octanol oxidation is promoted by the low-coordinated and electron-deficient Pt cluster species, as we demonstrated that the composition was highest in catalysts whose support possessed a higher proportion of exposed (001) facets.
A similar trend is observed in the associated reaction selectivity data. As with short chain mono-alcohols, 1-octanol first undergoes oxidation to the corresponding aldehyde (octanal), which can be sequentially oxidized through to the corresponding acid (octanoic acid). An additional esterification side reaction can also occur, leading to the formation octyl octanoate. For all the catalysts, at low conversion the primary product is octanal. However, the rate in which this intermediate is oxidized to octanoic acid differs and appears to be dependent on (i) whether the sample was calcined or reduced and (ii) the nature and distribution of the exposed surface facets. Note: only small quantities of octyl octanoate are produced in these reactions; reaction selectivity <6%. As with the oxidation of octanol, the rate of octanal oxidation to octanoic acid is identical for all of the reduced catalysts and a higher rate of conversion is observed over the calcined catalysts (Fig. 6B and C). The rate of octanal oxidation is also highest with the calcined catalyst possessing the greater proportion of the exposed (001) facet.
To gain further insight on the relationship between catalyst structure and catalytic activity, a series of additional in situ FTIR experiments were conducted using n-propanol as the probe molecule (Fig. 7); propanol was used as the probe molecule for these experiments due to the low vapor pressure of octanol. As with the CO-FTIR experiments (Fig. 5B), a small amount of the Pt/TiO2-001-H2 sample was mixed with KBr to amplify the observed signal. This explains why the associated FTIR spectrum (Fig. 6D) is slightly noisier than the other spectra presented (Fig. 7A–C). In Fig. 7(A), the absorption bands observed at 2964 and 2939 cm−1 are characteristic of asymmetric stretching of –CH3 and –CH2– (νas-CH3 and νas-CH2–) and the adsorption band at 2879 cm−1 is attributed to the symmetric stretching of –CH3 (νs-CH3).45,46 Adsorption bands at 1459, 1402, 1382 and 1364 cm−1 are characteristic of asymmetric deformation vibrations of –CH3 (δas-CH3), asymmetric deformation vibrations of adsorbed α-CH2– (δas-α-CH2–), symmetric deformation vibrations of –CH3 (δs-CH3), and the symmetric deformation vibrations of adsorbed α-CH2– (δs-α-CH2–), respectively. The bands at 1340 and 1228 cm−1 can be assigned to the δ modes of O–H and C–O stretching (νC–O). The adsorption band at 1100 cm−1 can be assigned to the rocking vibration of –CH3 and deformation vibration mode of –C–C–C–O (ω-CH3 and δ-C–C–C–O), while the adsorption band at 1066 and 1056 cm−1 can be assigned to the stretching vibrations of C–C (ν-C–C). Over the calcined Pt/TiO2-001-air catalyst however, the adsorption band indicative of δO–H incurs a red-shift, from 1340 over Pt/TiO2-101-air to 1326 cm−1. This indicates that there is an increase in the length of the O–H bond over this material, which could be indicative of a decrease in the bonds associated dissociation energy. Notably, the δs-α-CH2– band also exhibits a slight red shift from 1364 cm−1 over Pt/TiO2-101-air catalyst to 1358 cm−1 over Pt/TiO2-001-air catalyst, indicating that there may be a change in the length of the α-C–H bond. The relative intensity of the δs-α-CH2- band to δs-CH3 over Pt/TiO2-001-air catalyst is also higher than that of the other catalysts, suggesting there is a greater adsorption of α-C–H bond over the Pt/TiO2-001-air catalyst. According to the literature,14 the enhanced adsorption and the dissociation of the α-C–H bond could contribute to the subsequent elimination during the oxidation process and hence the Pt/TiO2-001-air catalysts achieved a higher conversion of octanol. Furthermore, the formation of metal-alkoxides through hydroxyl group deprotonation, and the subsequent β-hydride elimination, usually occur on the surface of noble metal active sites. Given that the Pt/TiO2-001-air catalyst possesses a greater proportion of low coordinate Pt clusters, we propose that the higher activity of this catalyst is attributed to the greater proportion of low-coordinated Pt sites it possesses.
![]() | ||
Fig. 7 In situ FTIR spectra of propanol adsorption for 30 min followed by desorption at 1, 5, 10, 15 and 20 min on (A) Pt/TiO2-101-air, (B) Pt/TiO2-001-air, (C) Pt/TiO2-101-H2, (D) Pt/TiO2-001-H2. |
Analogous in situ FTIR experiments were also conducted using propionaldehyde as the probe molecule (Fig. 8). Again, the Pt/TiO2-001-H2 catalyst was mixed with KBr to amplify the signal, which is why much of the features observed in the associated spectrum (Fig. 8D) are indicative of physically adsorbed propionaldehyde. However, some bands indicative of chemisorbed species are visible, allowing for qualitative comparisons to be made with the other catalysts (Fig. 8A–C). Many similar adsorption bands were observed over all the catalysts. Adsorption bands indicative of C–H stretching are observed at 2968, 2940 and 2882 cm−1. The bands at ca. 1762 and 1743 cm−1 can be assigned to the CO bond of propionaldehyde. However, notable differences are observed in the spectra of the calcined Pt/TiO2-101-air and Pt/TiO2-001-air catalysts between 1700 and 1300 cm−1. According to the literature47–52 and our previous study,22 the aldehyde group can be adsorbed and dissociated via monodentate or bidentate intermediates on the surface of TiO2 rather than Pt nanoparticles. As shown in Fig. 8A, the adsorption bands at ca. 1680 and 1413 cm−1 are characteristic of an aldehydic monodentate species over (101) facet of TiO2. For Pt/TiO2-001-air catalyst (Fig. 8B), the aldehydic bidentate species are predominant on the surface of (001) some with the adsorption bands at ca. 1691 and 1363 cm−1, while only a few of monodentate species exists with a small shoulder peak at 1409 cm−1. Considering the coordination unsaturated O2c–Ti5c–O2c sites of (001) facet, the bidentate species of aldehyde C
O bond of which the C and O atoms is bonded with O2c and the adjacent Ti5c are formed and highly dissociated on (001) facet. For the monodentate species of adsorbed aldehyde group, only the C atom is bonded with O2c site on (101) facet without the participation of the adjacent Ti due to the different coordination structure with (001). Besides, the adsorption bands at ca. 1463 and 1375 assigned to the asymmetric and symmetric deformation vibrations of –CH3 (δas-CH3 and δs-CH3) are similar over the series catalysts. But the symmetric deformation vibrations of adsorbed α-C–H (δs-α-C-H) bond has a red-shift from 1350 cm−1 over Pt/TiO2-101-air to 1323 cm−1 over Pt/TiO2-001-air, indicating a longer length of C–H bond over Pt/TiO2-001-air catalysts. Similar with the result of propanol FTIR, the prolonged C–H bond would be attributed to its further elimination and hence favours its oxidation to the carboxylic acid products. Furthermore, the intensity of carboxylate species centred at ca. 1626 cm−1 is increased only Pt/TiO2-001-air catalyst, which further indicating that (001) facet facilitates the further oxidation of aldehyde group. According to our previous study,22 although the bidentate species of aldehyde group are adsorbed on the coordination structure of TiO2 support, the oxidation only happens with the participation of metal active sites. Hence, the low-coordinated Pt active sites could be the reason for the prolonged α-C–H bond of aldehyde, as well as the alcohol.
From the results acquired from these in situ FTIR experiments, we propose that low-coordinated Pt clusters promote both the deprotonation of OH group in octanol and the subsequent activation of α-C–H bonds in both octanol and octanal. The O2c–Ti5c–O2c sites in the TiO2 (001) facet also promote the adsorption and dissociation of the CO bond, which processed via a bidentate intermediate. We propose that this is why a higher reaction selectivity to octanoic acid is observed over this catalyst, when compared at iso-conversion.
The in situ FTIR experiments suggest that the activation of both alcohols and aldehydes are promoted by the presence of the exposed (001) facet. Additional kinetic experiments were subsequently conducted to calculate activation energies for both 1-octanol and octanal oxidation over the different calcined catalysts (Fig. S5 and S6†). For both oxidation reactions, the catalyst possessing the support with the most exposed (001) facets exhibited the lowest activation energies, 30.2 and 27.6 KJ mol−1 for octanol and octanal oxidation, respectively. The least active catalyst, the support of which possessed the greatest proportion of (101) facets, exhibited activation energies of 39.9 and 40.8 KJ mol−1 for octanol and octanal oxidation, respectively. Given that there is a greater difference between the apparent activation energies for the oxidation of octanol, it suggests that the exposed surface facets have a greater influence on formyl oxidation, compared to alcohol oxidation.
On the basis of the above results and the previous studies, a possible mechanism is proposed in Scheme 1. At first, the hydroxyl group of 1-octanol is adsorbed and activated on the surface of low-coordinated and electron-deficient Pt cluster supported on the (001) facet, which promotes the deprotonation of the O–H band and forming an active hydroxyl species with the dissociated O species. The α-C–H of octanol is activated by the low-coordinated Pt cluster following with the β-hydride elimination with the effect of OH* species to form the intermediate product octanal. Then, the aldehyde group of octanal adsorbed on the coordination unsaturated O2c–Ti5c–O2c site of (001) to form a bidentate species while the α-C–H bond of adsorbed aldehyde bidentate species is activated by the nearby low-coordinated Pt cluster. The α-carbon atom of aldehyde group of bidentate species are more active than of monodentate species and benefits the attack of the active hydroxyl species, producing the octanoic acid after the C-H cleavage. Herein, it is an obvious synergetic effect of coordination structure O2c–Ti5c–O2c of (001) facet on support and low-coordinated Pt cluster that facilitate the oxidation of 1-octanol to the final octanoic acid products.
![]() | ||
Scheme 1 Possible reaction pathway for the oxidation of 1-octanol over the Pt/TiO2-001-air catalyst. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1cy00686j |
This journal is © The Royal Society of Chemistry 2021 |