Dorothea
Gömpel
a,
Muhammad Nawaz
Tahir
*bc,
Mujeeb
Khan
d,
Syed Farooq
Adil
d,
Mohammed Rafi
Shaik
d,
Mufsir
Kuniyil
d,
Abdulrahman
Al-Warthan
d and
Wolfgang
Tremel
*a
aChemistry Department, Johannes Gutenberg-Universität Mainz, Duesbergweg 10-14, D-55128 Mainz, Germany. E-mail: tremel@uni-mainz.de
bInterdisciplinary Research Center for Hydrogen Technologies and Carbon Management (IRC-HTCM), King Fahd University of Petroleum & Minerals KFUPM, Dahran 31261, Saudi Arabia. E-mail: muhammad.tahir@kfupm.edu.sa
cInterdisciplinary Research Center for Hydrogen and Energy Storage (IRC-HES), King Fahd University of Petroleum and & Minerals, Dahran 31261, Saudi Arabia
dDepartment of Chemistry, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Kingdom of Saudi Arabia
First published on 8th January 2024
Vanadium oxides are promising oxidation catalysts because of their rich redox chemistry. We report the synthesis of VO2 nanocrystals with VO2(B) crystal structure. By varying the mixing ratio of the components of a binary ethanol/water mixture, different VO2 nanocrystal morphologies (nanorods, -urchins, and -sheets) could be made selectively in pure form. Polydisperse VO2(B) nanorods with lengths between 150 nm and a few micrometers were formed at large water:
ethanol ratios between 4
:
1 and 3
:
2. At a water
:
ethanol ratio of 1
:
9 VO2 nanosheets with diameters of ∼50–70 nm were formed, which aggregated to nano-urchins with diameters of ∼200 nm in pure ethanol. The catalytic activity of VO2 nanocrystals for the oxidation of alcohols was studied as a function of nanocrystal morphology. VO2 nanocrystals with all morphologies were catalytically active. The activity for the oxidation of benzyl alcohol to benzaldehyde was about 30% higher than that for the oxidation of furfuryl alcohol to furfural. This is due to the substrate structure. The oxidation activity of VO2 nanostructures decreases in the order of nanourchins > nanosheets > nanorods.
The high catalytic activity of V-oxides can be attributed to the ability of vanadium to easily switch between its V3+, V4+ and V5+ oxidation states under the respective reaction conditions22 and to the presence of unsaturated surface coordination sites, which represent oxygen-deficient defects that allow reactions involving bulk oxygen atoms.9,23–26 Oxide surfaces undergo restructuring and can exchange oxygen atoms with the environment in ways that are difficult to predict.27,28 VO2 in particular is a catalyst for the desulfurization of dibenzothiophene,29 the oxidative dehydrogenation of propane30 or the electrochemical reduction of trinitrotoluene.31
Nanocrystals can serve as model catalysts with defined surface structure to study the structure–property relationships of powder catalysts. Apart from the oxidation state and crystallinity of vanadium oxide catalysts there are a variety of variables such as structure, surface area, morphology, surface wetting and interface interactions.32 Colloidal synthesis allows the controlled preparation of metal oxide nanocrystals with defined and uniform morphology. Different facets of a nanocrystal have different surface energies, surface structures, and chemical reactivity, which can significantly determine catalytic performance. By controlling the morphology and size of the nanocrystals, it is possible to tune the exposed facets and enhance the catalytic activity and selectivity for a given reaction.22,33,34
The activities of nanostructured VOx catalysts have been reported as a function of crystal morphology, crystal size and surface area.35,36 Most of these studies, however, have been carried out with Magneli-type vanadium oxides such as VnO2n+1 (V3O7, V4O9, and V6O13),9 whereas morphology-dependent catalytic properties of VO2 NPs have rarely been reported.29 Several routes have been reported for the synthesis of VO2 and other vanadium oxide nanocrystals.37–40 The most common ones are based on hydro- or solvothermal techniques.41 The choice of the solvent controls the reactivity, solubility and the diffusion of the precursors.42 In binary solvent mixtures these factors can be tuned accurately and independently. This allows the preparation of nanocrystals with different morphologies, e.g., spheres,43–45 rods,46–49 sheets,50–53 and urchins54,55 only by varying the volume ratio of the solvents.56
Sheet- and urchin-like morphologies are the most promising options for large surface areas, but the morphological effects can be offset by the particle size. Often the crystallites are several hundred nanometers in diameter. This leads in total to a significant reduction of the surface area. Therefore, a synthesis of nanometer-sized urchins and sheets is highly desirable. This requires the reactivity during the reaction to be fine-tuned to prevent uncontrolled aggregation into larger particles.
Here we describe the solvo-/hydrothermal synthesis of VO2(B) nanorods and VO2 nanosheets and -urchins. Therefore, the effect of solvent, especially the impact of chain length of aliphatic alcohols and the composition of binary mixtures on morphology and phase of VO2 nanoparticles was explored systematically. The catalytic activity of VO2 nanoparticles with different morphology and active surface area was tested for the oxidation of benzyl alcohol to benzaldehyde and furfuryl alcohol to furfural as a model reaction (Scheme 1). To the best of our knowledge, this is the first investigation of the catalytic properties of VO2 nanoparticles in oxidation reactions as a function of particle size and morphology.
![]() | ||
Scheme 1 Preparation of VO2 nanoparticles with different morphologies and their catalytic applications in the oxidation of furfuryl alcohol. |
The electron micrographs in Fig. 1 show that the morphology of the VO2 nanoparticles is determined by the composition of the solvent mixture. The ratio of water to ethanol was systematically varied between 0% (pure water) and 100% (pure ethanol). Nanorods with very different aspect ratios formed in water (Fig. 1a). The length of the rods varied between 150 nm and several micrometers, while the width was between 20 and 200 nm. At water to ethanol ratios of 4:
1 (20 mL
:
5 mL, Fig. 1b) and 3
:
2 (15
:
10 mL, Fig. 1c), the morphology was similar (nanorods), but the rods were significantly smaller and less polydisperse than when synthesized in pure water. Their width was in the range between 20 and 80 nm and their length between 100 and 220 nm. There was no visible difference in the morphology of the nanorods from water
:
ethanol mixtures between water
:
ethanol ratios of 4
:
1 and 3
:
2. The particles remained rod-like for higher ethanol concentrations. However, their polydispersity increased again (Fig. 1d, water
:
ethanol ratio 2
:
3). At a water
:
ethanol ratio of 1
:
4, the rods agglomerated into disordered bundles (Fig. 1e), and at a ratio of 1
:
9, nanorods no longer formed. The particles became increasingly isotropic but without defined morphology (Fig. 1f).
At high ethanol concentrations (water:
ethanol < 1
:
49), a significant change in morphology occurred, and sheet-like nanocrystals with a narrow size distribution were formed (Fig. 1g and h).
The morphology of some nanoparticles in Fig. 1g is reminiscent of structures obtained at a water–ethanol ratio of 1:
9 (Fig. 1f). In both cases, the diameter of the sheets was about 50–70 nm. The thickness of about 5 nm can be deduced from some sheets that were oriented vertically on the grid accidentally. Nanoparticles synthesized in pure ethanol have urchin-like morphologies with a total diameter of about 200 nm (Fig. 1i), as illustrated in Fig. 2. The wet chemistry methods used for the synthesis of nanoparticles are ideal to control the morphology, size and composition. In particular, hydrothermal/solvothermal methods59,60 where the solubility and re-precipitation of the precursors define the number of nuclei and the growth (size) of particles, can be utilized to control size and morphology. We used pure water (high dielectric constant) as solvent, resulting in a medium of low supersaturation, leading to a small number of nuclei and the formation of larger particles (nanorods). However, a gradual increase of the amount of ethanol (low dielectric constant that provides a medium with high supersaturation), the number of nuclei increases and ultimately results in small size particles sizes. Using pure ethanol, where the number of nuclei is highest, these nuclei combine to form nano-urchins to compensate their surface energy. Basically, a single urchin consists of several sheets that agglomerate in a random manner. Therefore, the crystallinity of the resulting particles is very low as indicated by the very broad diffraction intensities (Fig. 3, vide infra). The driving force for this agglomeration could be the different colloidal stability of the sheets depending on the solvent. Moreover, the sheets that compose the nano-urchins were smaller than free-standing sheets, thereby favoring aggregation. Thus, the composition of the solvent mixture is an effective means of controlling the morphology of vanadium oxide nanoparticles. By varying the water
:
ethanol ratio, three different morphologies – rods, sheets, and urchins (resulting from sheet agglomeration) – could be made in a reproducible fashion. We note that the surfactant F-127 helps to obtain well dispersed (but not porous) nanoparticles as indicated by the TEM images in Fig. S1.† The morphology of the nanoparticles was primarily controlled by using different ethanol
:
water ratios.
Fig. 3 shows the PXRD patterns of the products prepared at different water:
ethanol ratios. The diffractograms of the products prepared with water and water
:
ethanol ratios between 5
:
1 and 1
:
4 show the reflections of VO2(B) (JCPDS 812392, Fig. 3a–e). Some reflections display sharper profiles than others. This is indicative of strongly anisotropic crystallites, as the crystallite size is inversely proportional to the full width at half maximum (fwhm).61,62 Anisotropic crystallite sizes are in good agreement with the rod-like morphology of the nanoparticles (Fig. 3a–e), where the reflections should be sharper along the lattice direction associated with the long axis of the rods. For the rods obtained in pure water, the 00l reflections are weak. For the product obtained at a water
:
ethanol ratio of 1
:
9 (Fig. 3f), there or virtually no resolved reflections i.e., the intensities of the X-ray diffraction patterns are extremely broad (compared to the diffractograms in Fig. 3a–e). The principal reflection is roughly at the same 2θ value as that for VO2(B), indicating that the main product is also VO2(B), albeit with very low crystallinity, i.e., no long range order.
The PXRD patterns of the sheet-like products prepared with an ethanol excess (water–ethanol ratio < 1:
49, Fig. 3g–i) also show broad reflections, indicating a low crystallinity. However, the weak reflection at 2θ = 15.7° in Fig. 3g and h, which is not present in the sample prepared from pure ethanol, does not match the diffraction pattern of VO2(B). Its position roughly corresponds to the stacking separation of layers in the structure of layered vanadium oxides (like V2O5), although no other reflections of V2O5 are present. This might be compatible with the formation of a second phase which may be present as a surface (mono)layer on the crystallites of VO2(B), but might also be due to the sheet-like structure. The remaining broad intensities match in essence the diffraction pattern of VO2(B), indicating that it is still the principal phase, although the reflections are also very broad and weak. The diffractogram in Fig. 3i is again not well defined, and the broad reflections are in harmony with the morphology of the particles seen in the TEM images (Fig. 3g–i). These products are made up of nanostructures with thin layers forming sheets or urchins. Typical for layered structures (stacked in a random manner) are broad reflections with partial overlap. The overall crystallinity of the samples is very low, resulting in weak scattering (and broad/weak reflections). Due to the poor crystallinity and the lamellar morphology it was difficult to clearly assign individual reflections and to assign the crystalline phase unambiguously. The change in morphology from rod- to sheet-like structures observed in the TEM would be compatible with a change in crystal structure of VO2(B) to a sheet-like structure and possibly the formation of a surface layer containing fragments of the V2O5 structure. This assumption would also be in harmony with a Raman band at ∼990 cm−1 (Fig. 5, vide infra) that indicates the presence of polymeric V2O563,64 containing distorted VO5 pyramids sharing edges and corners as structural motif, while VO2 (whose structure is based on edge-sharing VO6 octahedra) is responsible for the remaining bands.65
This might arise from a “monolayer” of V2O5 on the surface of the VO2 nanocrystals (that does not show up in the X-ray diffractograms due to the absence of long-range order). Although the XPS spectra (Fig. 7, vide infra) could indicate that vanadium remains in oxidation state +4 (V 2p at 516.9 eV) consistent with a VO2 composition the V 2p signals of V4+ and V5+ are to close (516.3, 517.3 eV)66 to allow for a clear distinction.
The VO2 nanorods, -sheets and -urchins were characterized by Raman spectroscopy (Fig. 5). The Raman spectrum of the nanorods shows bands at 283 and 406 cm−1 that can be assigned to V–O bending vibrations. The band at 477 cm−1 belongs to the V–O–V bending vibration, and the broad peak at 526 cm−1 corresponds to the stretching mode of triply coordinated V3–O. The V2–O band is located at 695 cm−1 whereas the V–O stretch appears at 878 cm−1. Additionally, there is another band at 929 cm−1 which is attributed to a (local, i.e., defect-related) V4+O moiety.67
The strong Raman peak at 995 cm−1 might be caused by the V–O stretching mode of a V2O5 surface layer.63,64,68,69 Still, the Raman and X-ray data corroborate that the principle phase of the rods is VO2. The different signal-to-noise ratios between sheets and urchins on one hand and rods on the other hand are due to different filters. The nanosheets and -urchins decompose under the measurement conditions for the rods. Therefore, the radiation intensity is lower for the sheets and urchins, while no peaks were observed for the rods. Since the local structure of the phases crystallizing in sheet and urchin form is not complete from the PXRD and the vibrational spectra, it remains difficult to assign the bands and phase identity unambiguously. Both, urchins and sheets, exhibit their strongest band at the same wavenumber (855 cm−1). This position is typical of a V–O stretch. Additionally, a weak band appears at 366 cm−1, which is due to V–O bending modes.70 The Raman spectra suggest in agreement with the PXRD and IR data that the sheets and urchins have the same structure, because they show the same characteristic bands. The difference in fwhm of the VO2 urchins and sheets is due to their respective morphology. The urchins show a higher degree of disorder as the petals are randomly oriented which in turn leads to a strong peak broadening in the Raman spectrum.
To further characterize the nanosheets, their composition was examined by energy dispersive X-ray spectroscopy (EDX, Fig. 6a). EDX confirms the presence of vanadium, oxygen, carbon and copper in the sample. The copper signal is from the copper grid on which the sample was measured. In addition, the carbon signal is likely caused in part by (i) the carbon coating on the grid. The carbon signal further originates (ii) from the organic components that serve as capping agents for the nanoparticles. The FT-IR spectrum of the sheets shows the presence of alkyl groups (Fig. 4). Therefore, the layers are composed of vanadium and oxygen, while part of the carbon signal likely originates from the ligands covering the nanoparticles. Fig. 6b shows a dark field image of the nanosheets with a size between 50 and 70 nm. Here, upright sheets with a thickness of 5 nm can be seen. Some electron diffraction patterns of the sheets are shown in Fig. 6c and d. Fig. 6c shows the patterns within the sheet layer. The sheet spacings are 6.6 Å and 23.8 Å (compatible with the VO2(B) structure10) with systematic extinctions along the short axis. A diffraction pattern from the side of a platelet is shown in Fig. 6d, which shows spacings of 14.7 Å. Overall, the reflections are very broad, and the crystallinity of the associated phase is very low. Therefore, it was not possible to determine the cell parameters of the VO2 sheets by electron diffraction, especially since the material is radiation sensitive. The sheets (with the organic surface ligands) decompose under the beam. Either bubble formation or hole etching occurs as in Fig. 6e (red circle), which prevents a structure determination by automated diffraction tomography (ADT).71–73
Finally, the oxidation state of the vanadium atoms in the vanadium oxide nanoparticles was determined by XPS spectroscopy. In the XPS spectra (Fig. 7), the oxidation state of the respective atoms determines the peak position via the electron binding energies (or electron work functions). The nanorods, -urchins, and -sheets show the vanadium 2p peaks at 516.4 eV. The assignment of oxidation states (V4+vs. V5+) based on XPS data is not unambiguous, because the V 2p signals of the different oxides may appear at similar energies (516.3, 517.3 eV) and the value for the nanorods, -urchins, and -sheets (516.4 eV) is in between. Moreover, the oxygen 1s peak (530 eV) which could not be deconvoluted agrees for the three particle types, i.e., nanorods, -urchins, and -sheets contain vanadium in the same oxidation states. Since the phase identity of the crystalline fraction of the nanorods is assured from the PXRD diffractograms and the Raman spectra, (Fig. 3 and 5) the urchins and -sheets also contain V4+, i.e., VO2. The positions of the vanadium 2p (516.9 eV) and oxygen 1s orbitals (530 eV) agree with literature data for the respective bands.74,75
![]() | ||
Fig. 5 Normalized Raman spectra of the VO2 nanorods (black), nano-sheets (red) and nano-urchins (green). |
![]() | ||
Fig. 6 (a) TEM EDX, (b) darkfield image, (c and d) electron diffraction patterns and (e) darkfield image showing radiation damage of the VO2 sheets. |
Here we investigate the effect of particle morphology on the catalytic performance of the VO2 nanoparticles. All three morphologies, i.e., VO2 nanorods, -sheets, and -urchins were used as catalysts for the oxidation of two model compounds, benzyl alcohol and furfuryl alcohol. The experimental procedure for the oxidation of alcohols and GC chromatogram showing product purity of furfuryl alcohol and benzyl alcohol oxidation using VO2 nano-urchins is shown in the ESI Fig. S2 and S3.† Oxidation reactions were performed under identical conditions reaction temperature 150 °C with O2 gas as oxygen source for comparison. All three VO2 particle morphologies were active for the oxidation of benzyl alcohol to benzyl aldehyde. The results in Table 1 show that the catalytic efficiency is morphology dependent. VO2 nano-urchins showed highest yield (∼100%) and a specific activity of 3.01 mmol g−1 h−1. VO2 nanorods showed the lowest yield of ∼71% with a specific activity of 2.13 mmol g−1 h−1. VO2 nanosheets yielded a slightly lower conversion (87%) than urchins. For the oxidation of furfuryl alcohol to furfural the catalytic activity was significantly lower (Table 2) with a maximum conversion of ∼67% for the urchin morphology. Still, the trend in terms of morphology dependence remained. VO2 urchins showed the highest conversion (67%), VO2 nanosheets a slightly lower conversion of ∼60%. while the lowest conversion (∼33%) was achieved with VO2 nanorods. The lower product conversion during the oxidation of furfuryl alcohol compared to that of benzyl alcohol can may be attributed to the O heteroatom in the furfuryl ring system.
Morphology (VO2 NPs) | Benzyl alcohol → Benzaldehyde |
---|---|
Conversion (%) | |
Nano-urchins | 100 |
Nanosheets | 87 |
Nanorods | 71 |
Morphology (VO2 NPs) | Furfuryl alcohol → Furfural |
---|---|
Conversion (%) | |
Nanourchins | 67 |
Nanosheets | 60 |
Nanorods | 33 |
Surface area analysis (using multipoint BET measurement) shows the correlation between the catalytic activity of the VO2 nanocatalysts and their surface area. VO2 nanorods have a surface area of 21.7 m2 g−1 measured, which is within the typical size range for nanorods.76 The surface area of the urchins and sheets is significantly higher (124.6 and 73.9 m2 g−1, respectively) due to their higher surface-to-volume ratio. Specifically, the surface area of the VO2 urchins is remarkably high compared to the reported surface areas of sheet- and urchin-like particles. Xu et al.77 synthesized urchin-like V2O3 particles with a surface area of 48.57 m2 g−1, while Pan et al.78 obtained flower-like V2O5 with a surface area of 33.64 m2 g−1. The differences in surface area are due to the difference in the overall size of the particles, which was in the nanometer range for the VO2 urchin-type particles, while the diameters of the particles reported in the literature are in the micrometer range. Another factor is the thickness of the “petals”. The thinner the petals, the larger the surface area. Pang et al.79 obtained flower-like V4O9 microparticles with a specific surface area of 107.9 m2 g−1, which can be attributed to the thin layers of 2 nm. The surface areas of nanoparticles are usually much larger than those of bulk vanadium oxides (e.g., V2O5 with 8.4 m2 g−1). When correlating the surface areas with the catalytic performance, the catalytic performance is high with increasing surface area. The highest conversion of alcohol to aldehyde is obtained for urchins and sheets, while the lower active surface area of nanorods entails a lower conversion. For all morphologies there are indications for surface layers with local V2O5-type characteristics (without long-range order). Since these surface layers with local V2O5-type moieties but without translational symmetry are present on all surfaces, they affect the catalytic activity of the VO2 particles in a comparable way. To confirm that enhanced catalytic activity is related to VO2 nanomaterials with traces of V2O5 surface patches (VO2/V2O5 hybrid) and not just due to the surface V2O5 layer, we provide a comparison based on previous reports.81–86 (Table S1†). The results in Table S1† show that, depending on their morphology, VO2 nanomaterials with patches of a V2O5 surface layer exhibit better catalytic activity than pure V2O5 nanoparticles as well as V2O5 nanoparticles on different support surfaces.
The turnover number (TON) and turnover frequency (TOF) for the three different morphologies of VO2 nanocatalysts are also compiled in Tables 1 and 2. For the oxidation of benzyl alcohol to benzaldehyde, the TON and TOF values range from 9.97 to 7.08 and from 2.49 to 1.77 h−1, respectively. For the oxidation of furfuryl alcohol to furfural, the values are 8.00–3.94 and 2.00–0.99 h−1, respectively. Urchin-like VO2 nanoparticles had the highest, VO2 nanorods the lowest TON and TOF values. The results are illustrated in Fig. 8 and 9. The reusability of the VO2 nano-urchins, the best catalyst, was investigated in five cycles. The catalyst was essentially stable (activity decrease from 100% to ∼96%, Fig. 9). The VO2 materials are structurally stable as shown by the X-ray diffractogram (Fig. S4†) of the VOx nanourchins measured after the catalytic application.
![]() | ||
Fig. 8 TON and TOF values for as-synthesized VO2 nanocrystals, (A) benzyl alcohol and (B) furfuryl alcohol. |
Based on the catalytic results obtained, we propose a possible reaction mechanism for the oxidation of alcohols over the VO2 nanocatalysts (Fig. S5†). The adsorption of alcohols on the surface of nanocatalysts occur in the beginning of the catalytic cycle. Subsequently, breaking of O–H bond and loss of a proton may lead to the formation of VO2-alkoxide on the surface of the cluster, which is suggested by the DFT simulations in an earlier study.80 In the next step, the released proton is captured by adsorbed O2 on the catalyst surface, which lead to the cleavage of the C–H bond of α-C. This may lead to the generation of desired carbonyl compound and regeneration of the VO2 surface (IV).
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3dt02605a |
This journal is © The Royal Society of Chemistry 2024 |