Next Article in Journal
Li4(OH)3Br-Based Shape Stabilized Composites for High-Temperature TES Applications: Selection of the Most Convenient Supporting Material
Next Article in Special Issue
Efficient Visible-Light Photocatalysis of TiO2-δ Nanobelts Utilizing Self-Induced Defects and Carbon Doping
Previous Article in Journal
Impacts of Amplitude and Local Thermal Non-Equilibrium Design on Natural Convection within NanoflUid Superposed Wavy Porous Layers
Previous Article in Special Issue
Facile Strategy for Mass Production of Pt Catalysts for Polymer Electrolyte Membrane Fuel Cells Using Low-Energy Electron Beam
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermal CO Oxidation and Photocatalytic CO2 Reduction over Bare and M-Al2O3 (M = Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au) Cotton-Like Nanosheets

1
Department of Chemistry, Yeungnam University, Gyeongsan 38541, Korea
2
Department of Chemistry, Chungnam National University, Daejeon 34134, Korea
3
Department of Chemical Engineering and Applied Chemistry, Chungnam National University, Daejeon 34134, Korea
*
Author to whom correspondence should be addressed.
These authors equally contributed to this work.
Nanomaterials 2021, 11(5), 1278; https://doi.org/10.3390/nano11051278
Submission received: 30 March 2021 / Revised: 8 May 2021 / Accepted: 9 May 2021 / Published: 13 May 2021

Abstract

:
Aluminum oxide (Al2O3) has abundantly been used as a catalyst, and its catalytic activity has been tailored by loading transition metals. Herein, γ-Al2O3 nanosheets were prepared by the solvothermal method, and transition metals (M = Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au) were loaded onto the nanosheets. Big data sets of thermal CO oxidation and photocatalytic CO2 reduction activities were fully examined for the transition metal-loaded Al2O3 nanosheets. Their physicochemical properties were examined by scanning electron microscopy, high-resolution transmission electron microscopy, X-ray diffraction crystallography, and X-ray photoelectron spectroscopy. It was found that Rh, Pd, Ir, and Pt-loading showed a great enhancement in CO oxidation activity while other metals negated the activity of bare Al2O3 nanosheets. Rh-Al2O3 showed the lowest CO oxidation onset temperature of 172 °C, 201 °C lower than that of bare γ-Al2O3. CO2 reduction experiments were also performed to show that CO, CH3OH, and CH4 were common products. Ag-Al2O3 nanosheets showed the highest performances with yields of 237.3 ppm for CO, 36.3 ppm for CH3OH, and 30.9 ppm for CH4, 2.2×, 1.2×, and 1.6× enhancements, respectively, compared with those for bare Al2O3. Hydrogen production was found to be maximized to 20.7 ppm during CO2 reduction for Rh-loaded Al2O3. The present unique pre-screening test results provided very useful information for the selection of transition metals on Al2O3-based energy and environmental catalysts.

1. Introduction

Aluminum oxide (Al2O3) has extensively been used as a heterogeneous catalyst in diverse catalytic reactions of CO oxidation [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19], CO2 reduction [20,21,22,23], CO2 methanation/hydrogenation [20,24,25,26,27], and preferential oxidation of CO [28,29]. The efforts to increase the catalytic activity of the metal oxide have been devoted to the modification of the metal surface by loading of transition metals in groups of 9 (Co, Rh, and Ir), 10 (Ni, Pd, and Cu), and 11 (Cu, Ag, and Au). The morphology of a metal oxide support has also been a key factor for the enhancement of the catalytic activity [30]. It was reported that the role of an overlayer metal becomes different when the support metal oxide is different [2]. The relative catalytic activities of overlayer metals also become different when the catalytic application areas are different for the application to CO oxidation using transition metal-loaded Al2O3 catalysts.
Chen et al. prepared Pt nanoparticles (NPs) on Al2O3 and observed 100% CO conversion at −20 °C [1]. For the extremely high activity compared with a commercial Pt/Al2O3, based on the experimental and the density functional theory (DFT) calculations, they proposed that CO was initially adsorbed on Pt(OH) kink sites and reacted with OH to release gaseous CO2. Afterward, OH was regenerated by activation of O2 on terrace sites. Lou and Liu studied CO oxidation of single Pt atoms dispersed on Fe2O3 (highly reducible), ZnO (reducible), and γ-Al2O3 (irreducible) supports, and observed that the catalytic activity was in the order of Pt/γ-Al2O3 < Pt/ZnO < Pt/Fe2O3 [2], where the highly reducible support showed the highest catalytic activity. Chen et al. tested Pt/Al2O3 for preferential oxidation (PROX) of CO in H2 [29]. They concluded that CO conversion and CO2 selectivity reached up to 100% in a wide range of −30 °C to 120 °C. The high performance was attributed to a combination of Pt(OH) and metallic Pt on the Al2O3 support. Therefore, the adsorption of CO and the activation of O2 were optimally tuned to maximize the performance. For monodispersed single Pt atoms on θ-Al2O3, Moses-DeBusk et al. found that the CO oxidation did not follow a conventional Langmuir-Hinshelwood mechanism [11]. The Pt atom was first oxygenated, and then CO was bound to form a carbonate (CO3), which dissociated to generate gaseous CO2 [11]. Yang et al. employed the DFT calculation to investigate the relative CO oxidation for single-atom catalysts of Ni/γ-Al2O3 and Pd/γ-Al2O3 [7]. They reported that Ni showed an unexpectedly higher CO oxidation activity than the Pd. Ananth et al. synthesized Ag2O/γ-Al2O3 and (Ag2O + RuO2)/γ-Al2O3 catalysts and tested the CO oxidation performances to show that the catalytic activity was increased by the addition of RuO2 [6]. Han et al. reported a high CO oxidation activity at 30 °C for NiO (≤1 nm) on mesoporous Al2O3 prepared using atomic layer deposition [16]. The deactivation was found to be lowered with increasing the pre-annealing temperature.
For the application of Al2O3 to CO2 reduction, Zhao et al. synthesized Au/Al2O3/TiO2 nanocomposites, where the atomic-layer Al2O3 was sandwiched between the two layers [21]. They tested the photocatalytic CO2 reduction activity and observed CO (major) and CH4 (minor) as products. It was concluded that the charge transfer and surface charge recombination were highly influenced by Al2O3 interlayer thickness. Therefore, the maximum photocatalytic activity (37 μmol/g of CO and 2 μmol/g of CH4) was obtained by achieving optimum Al2O3 thickness (5 Å). Kwak et al. performed a temperature-programmed CO2 reduction with H2 on Ru/Al2O3 catalysts and observed CO and CH4 formation yields with activation energies of 82 kJ/mol and 62 kJ/mol, respectively [20]. It was found that CO formation selectivity was increased with increasing Ru metal dispersion but decreased with increasing Ru clustering and concluded that CO was not an intermediate species for CH4 formation. Chein and Wang tested CO2 methanation activities using Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 catalysts [27] and found that the hybrid bimetallic Ru-Ni showed higher performance than the monometallic catalysts.
Although numerous detailed in-depth studies have been performed using transition metal-loaded Al2O3 catalysts, no systematic comparison studies have been reported among diverse (M = Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au) transition metal-loaded Al2O3 catalysts prepared by the same synthesis method. Motivated by this, we synthesized transition metal-loaded Al2O3 nanosheets and evaluated thermal CO oxidation activity as well as photocatalytic CO2 reduction activity. Consequently, the roles of overlayer transition metals were comparatively investigated in two totally different application reactions. Thereby, the present pre-screening test results provided useful information on the quick-selection of catalysts for thermal CO oxidation and photocatalytic CO2 reduction.

2. Materials and Methods

2.1. Catalysts Synthesis Procedures

For the synthesis of Al precursor nanosheets, 1 mmol of aluminum nitrate nonahydrate (Al(NO3)3 9H2O, 98%, Sigma-Aldrich, St. Louis, MO, USA), 0.008 g of polyethylene glycol (PEG, Mn = 4000, Sigma-Aldrich, St. Louis, MO, USA), and 20 μL of oleic acid (≥99%, Sigma-Aldrich, St. Louis, MO, USA) were fully dissolved by magnetic stirring in a mixed solvent of 10 mL of deionized water and 15 mL of ethanol (99.9%, Samchun Chem., Gyounggi, Korea) for 20 min. After that, the solution was transferred into a Teflon-lined stainless-steel autoclave reactor, which was then tightly capped for sealing. The tightly capped reactor was placed in an oven setting at 200 °C for 12 h. After the thermal reaction, the reactor was naturally cooled to room temperature, and the finally obtained white precipitates were collected by washing with deionized water and ethanol repeatedly by centrifugation at 3600 rpm. The collected wet powder was fully dried in an oven setting at 80 °C for 24 h. To obtain Al2O3 nanosheets, the dried powder sample was thermally annealed at 600 °C for 2 h.
For transition metal (M = Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au) loadings, 50 mg of Al2O3 nanosheets were fully dispersed in 20 mL ethanol, followed by adding 2 mol% of metal ions. The chemicals for metal ions were cobalt(II) nitrate hexahydrate (≥98%, Sigma-Aldrich, St. Louis, MO, USA), nickel(II) nitrate hexahydrate (≥98.5%, Sigma-Aldrich, St. Louis, MO, USA), copper(II) nitrate trihydrate (99%, Daejung, Gyounggi, Korea), rhodium(III) chloride hydrate (≥99.9%, Sigma-Aldrich, St. Louis, MO, USA), palladium(II) chloride (99.8%, Sigma-Aldrich, St. Louis, MO, USA), silver nitrate solution (0.1 N, Samchun Pure Chem., Gyounggi, Korea), platinum (III) chloride (≥99.9%, Sigma-Aldrich, St. Louis, MO, USA), iridium (III) chloride hydrate (≥99.9%, Sigma-Aldrich, St. Louis, MO, USA), and gold(III) chloride trihydrate (≥99.9%, Sigma-Aldrich, St. Louis, MO, USA). After complete mixing, the solvent was slowly evaporated by gentle heating (50 °C) while stirring. The dried M-loaded Al2O3 nanosheets were, again, thermally annealed at 600 °C for 2 h.

2.2. Sample Characterization

The surface morphologies of the Al-precursor, Al2O3 nanosheets, and M-loaded Al2O3 nanosheets were examined using a scanning electron microscope (SEM, Hitachi S-4800, Hitach Ltd., Tokyo, Japan) at conditions of 10 kV and 10 mA. X-ray crystallographic diffraction patterns were recorded using a PANalytical X’Pert Pro MPD diffractometer (PANalytical, Almelo, Netherland) with Cu Kα radiation (40 kV and 30 mA). Transmission electron microscopic (TEM) and high-resolution TEM (HRTEM) were obtained for bare Al2O3 nanosheets and the selected Ni- and Rh-loaded Al2O3 nanosheets using an FEI Tecnai G2 F20 TEM (Hillsboro, OR, USA) operated at 300.0 kV. X-ray photoelectron spectra were taken using a Thermo-VG Scientific K-alpha+ spectrometer (Thermo VG Scientific, Waltham, MA, USA) with a hemispherical energy analyzer. Attenuated total reflection Fourier-transform infrared spectroscopy (FT-IR) was employed using a Nicolet iS 10 FT-IR spectrometer (Thermo Scientific Korea, Seoul, Korea). The Brunauer-Emmett-Teller (BET) surface areas were measured using a ChemBET TPR/TPD analyzer (Quantachrome Instruments Corp., Boynton Beach, FL, USA) equipped with a thermal conductivity detector.

2.3. Thermal CO Oxidation and Photocatalytic CO2 Reduction Experiments

For thermal CO oxidation reactions, 20 mg of a catalyst was initially loaded into a U-shape quartz tube. After that, the tube was positioned in a temperature-programmed furnace. The temperature heating rate was 20 °C/min, and the flowing gas was CO(1.0%)/O2(2.5%)/N2 at a flow rate of 40 mL/min. The gas products from the outlet of the tube were monitored using a quadrupole mass spectrometer (RGA200, Stanford Research Systems, Sunnyvale, CA, USA). After the first run to a maximum temperature of 500 °C, the sample cell was naturally cooled to a room temperature of 25 °C. After that, the second run was performed at room temperature.
For photocatalytic CO2 reduction experiments, 3 mg of a catalyst was fully dispersed on a quartz disc (an area of 15.9 cm2) and placed in a stainless-steel reactor (volume ~40 mL) with additional deionized water (20 μL) beside the disc. After that, the reactor was tightly closed with a quartz window (0.3 cm thick and 4.5 cm diameter) on top. Afterward, pure (99.999%) CO2 gas was fully flushed and filled with the gas. For the photocatalytic CO2 reduction test, the reactor with the quartz window was placed under UVC (200–280 nm) lamps (a power density of 5.94 mW/cm2) for 12 h. After the UVC irradiation time, 0.5 mL of gas was taken and injected into a YL 6500 gas chromatograph (GC, Young In Chromass Co., Ltd., Seoul, Korea). For the analysis of the CO, CH3OH, CH4, and H2 products, the GC system was equipped with HP-Plot Q-PT column (Agilent Technologies, Inc., Santa Clara, CA, USA), 40/60 Carboxen-1000 column (Sigma-Aldrich, St. Louis, MO, USA), a Ni catalyst methanizer assembly, a thermal conductivity detector, and a flame ionization detector.

3. Results and Discussion

Figure 1(a,a1) show the SEM image of the as-prepared Al-precursor with a morphology of cotton-like nanostructures. Figure 1b shows the sample after thermal annealing at 600 °C, abbreviated as bare Al2O3. It appears that the morphology showed no significant change, but the nanosheets became somewhat compacted. The Brunauer-Emmett-Teller (BET) surface area of bare Al2O3 was measured to be 154.4 m2/g. The corresponding transmission electron microscope (TEM) image clearly showed the morphology of nanosheets. For the high-resolution TEM image of bare Al2O3, clear lattice fringes were seen, and the lattice spacing was estimated to be 0.197 nm. This was well-matched to the (002) crystal plane of cubic phase gamma-Al2O3. This was further discussed in detail below. The structure projection of the (002) and (022) planes for Al2O3 are shown in Figure 1(b1) for visual understanding.
Figure 2 shows the SEM and TEM images of selected Ni- and Rh-loaded Al2O3 nanosheets. The SEM images of other M-loaded Al2O3 nanosheets are provided in the Supporting Information, Figure S1. The SEM image (Figure 2a) of Rh-Al2O3 showed small nanoparticles embedded on the nanosheets. The nanoparticles appeared as a result of Rh particle formation. The color of burlywood was clearly different from the white color for bare Al2O3. On the other hand, the SEM image (Figure 2b) of Ni-Al2O3 nanosheets showed only cotton-like nanosheets. It was difficult to discriminate Ni species from the bare Al2O3 support. However, the color clearly changed from white to pale blue upon Ni-loading. The photos and optical microscope images of the M-loaded Al2O3 nanosheets are provided in the Supporting Information, Figures S2 and S3, respectively. Although the SEM images showed no clear metal embedment, the color change was a clear indication of metal-loading on the Al2O3 support. The metal-loading was also confirmed by the X-ray photoelectron spectroscopy (XPS) data, discussed below.
The TEM images (Figure 2(a1)) of Rh-Al2O3 clearly showed NPs (with a size of ~20 nm) embedded onto the nanosheets. For the HRTEM image (Figure 2(b2)) of an Rh-NP, clear lattice fringes were observed, and the distances were estimated to be 0.263 nm and 0.254 nm. These distances matched well with the (114) and (200) crystal planes of orthorhombic (Pbca) Rh2O3 (ICSD ref. 98-000-9206), respectively. This indicated that Rh was embedded not in the metallic form but rather in the oxide form. This was further confirmed by the XPS data below. The fast Fourier transform (FFT) pattern of the HRTEM image reflected the crystallinity of the Rh oxide. The TEM images (Figure 2(b1)) of Ni-Al2O3 nanosheets showed only nanosheet morphology, consistent with the corresponding SEM image (Figure 2b).
For the HRTEM image in Figure 2(b2), the lattice fringes with distances of 0.227 nm and 0.196 nm matched well with the (111) and (002) crystal planes of the cubic phase γ-Al2O3. The lattices showed poor crystallinity compared with those of bare Al2O3, seen in Figure 1(b3). Interestingly, some areas (dotted circles) showed very poor crystallinity, and these appeared like amorphous particles. This is likely an indication of Ni embedment on the Al2O3 nanosheets. Very similarly for Co-Al2O3 nanosheets, although particles were not clearly seen in the TEM image (Supporting Information, Figure S4), the corresponding HRTEM image showed the areas with very poor crystallinity. The areas appeared like Co-embedment in the Al2O3 support.
The BET surface areas of Ni-Al2O3 and Rh-Al2O3 nanosheets were measured to be 151.2 m2/g and 153.8 m2/g, respectively. The surface areas were very similar to that of bare Al2O3. This indicated that the surface area was not significantly impacted by the metal-loading.
Figure 3 displays the X-ray diffraction patterns of bare Al2O3 and M-loaded Al2O3 nanosheets. For the XRD patterns of bare Al2O3, two distinctive peaks were observed at 2θ = 45.9° and 67.0°. These two peaks could be assigned to the (002) and (022) planes of cubic phase (Fm-3m) γ-Al2O3, (ICSD ref. 98-003-0267), respectively. The XRD result was in good consistency with the HRTEM result of the bare Al2O3 nanosheet. For the XRD profiles of M-loaded Al2O3 nanosheets, two peaks were commonly observed, as expected. Interestingly, Co, Ni, Cu, Rh, and Ag-loaded samples showed no significant extra peaks in the corresponding XRD profiles. These results indicated that the metals were loaded with an amorphous oxide state (discussed below in XPS) or embedded very uniformly without forming good crystal phases. In addition, because the metal amount was only 2 mol%, the XRD patterns could not be clearly observed when the phase was an amorphous oxide form. As seen in the HRTEM images of Ni-Al2O3 and Co-Al2O3 nanosheets discussed above (Figure 2(b2) and Figure S4, respectively), the particle-like areas showed very poor crystallinity. On the other hand, Pd, Ir, Pt, and Au-loaded samples showed new peaks in the corresponding XRD profiles.
For the XRD patterns of Pd-Al2O3 nanosheets, several peaks at 2θ = 33.8°, 42.0°, 54.7°, 60.1°, and 71.5° showed good matches with the (011), (110), (112), (013), and (121) crystal planes of tetragonal (p 42/mmc) PdO (ICSD ref. 98-002-9281), respectively [13]. For Ir-Al2O3 nanosheets, several strong XRD peaks were observed at 2θ = 27.9°, 34.6°, 39.9°, 53.9°, 57.9°, 58.3°, 66.0°, 69.0°, and 73.0°, with good matches with the (110), (011), (020), (121), (220), (002), (130), (112), and (031) crystal planes of tetragonal (p 42/mnm) IrO2 (ICSD ref. 98-008-4577), respectively. For Pt-Al2O3 nanosheets, three major peaks were observed at 2θ = 39.8°, 46.2°, and 67.5°, assigned to the (111), (002,) and (022) crystal planes of the cubic (Fm-3m) crystal phase of metallic Pt (ICSD ref. 98-007-6153), respectively. For Au-Al2O3 nanosheets, three strong peaks were observed at 2θ = 38.1°, 44.3°, and 64.5°, assigned to the (111), (002), and (022) crystal planes of the cubic (Fm-3m) crystal phase of Au (ICSD ref. 98-061-1624), respectively.
XPS was employed to confirm the loading of the transition metals and examine the oxidation states. Figure 4 shows Co 2p, Ni 2p, Cu 2p, Rh 3d, Pd 3d, Ag 3d, Ir 4d, Pt 4d, and Au 4d of Co-Al2O3, Ni-Al2O3, Cu-Al2O3, Rh-Al2O3, Pd-Al2O3, Ag-Al2O3, Ir-Al2O3, Pt-Al2O3, and Au-Al2O3 nanosheets, respectively. XPS valence band spectra (Figure 4, right panel) are also displayed for the corresponding samples. The survey, Al 2p, O 1s, and C 1s profiles are provided in the Supporting Information, Figure S5. All the binding energies (BEs) were referenced to the C 1s XPS peak at 284.8 eV. The survey spectra commonly showed the elements of Al, O, and C (surface impurities), as expected. The XPS peaks of the loaded transition metals were very weakly observed.
For Co-Al2O3 nanosheets, Co 2p1/2 and Co 2p3/2 XPS peaks were observed at binding energies (BEs) of 797.4 eV and 781.6 eV, respectively, with a spin-orbit splitting of 15.8 eV. This could be attributed to Co2+ of CoO and Co(OH)2 [31,32]. The corresponding satellite peaks for Co2+ were clearly observed around 803 eV and 786 eV. For Ni-Al2O3 nanosheets, Ni 2p1/2 and Ni 2p3/2 XPS peaks were observed at binding energies (BEs) of 873.4 eV and 856.1 eV, respectively, with a spin-orbit splitting of 17.3 eV. This could be attributed to Ni2+ of NiO and Ni(OH)2 [15,31,32]. The corresponding satellite peaks for Ni2+ were clearly observed around 880 eV and 862 eV. For Cu-Al2O3 nanosheets, Cu 2p1/2 and Cu 2p3/2 XPS peaks were observed at binding energies (BEs) of 952.4 eV and 932.7 eV, respectively, with a spin-orbit splitting of 19.7 eV. This could be attributed to Cu2+ of CuO and Cu(OH)2 [32,33]. The corresponding satellite peak for Cu2+ was clearly observed around 942 eV. For Co, Ni, and Cu, no metallic XPS peaks were observed. On the basis of XRD, HRTEM, and XPS data, Co, Ni, and Cu appeared to be embedded as an amorphous oxide form.
For Rh-Al2O3 nanosheets, Rh 3d3/2 and Rh 3d5/2 XPS peaks were observed at BEs of 314.3 eV and 309.7 eV, respectively, with a spin-orbit splitting of 4.6 eV. The XPS BEs were attributed to an oxidation state of Rh3+ [32,34]. As discussed above, the lattice distances in the HRTEM image confirmed orthorhombic Rh2O3. An additional weak shoulder peak was seen around 308 eV for Rh 3d5/2 peak. This could be due to metallic Rh [32,34]. On the basis of the XPS and HRTEM data, Rh-species appeared to be consistent with Rh@Rh2O3 core-shell type structure.
For Pd-Al2O3 nanosheets, Pd 3d3/2 and Pd 3d5/2 XPS peaks were observed at BEs of 341.7 eV and 336.2 eV, respectively, with a spin-orbit splitting of 5.5 eV. The XPS peaks were attributed to an oxidation state of Pd2+ [32,34,35]. There was a good coincidence between the oxidation state of the XPS and the XRD profiles of tetragonal PdO. A weak shoulder of the Pd 3d5/2 peak was seen around 335.5 eV, plausibly due to metallic Pd [32]. For Ag-Al2O3 nanosheets, Ag 3d3/2 and Ag 3d5/2 XPS peaks were observed at BEs of 374.5 eV and 368.6 eV, respectively. This was attributed to metallic Ag [6,32,36]. The shoulder XPS peak at 367.6 eV for Ag 3d5/2 was plausibly due to AgO [36]. On the basis of the XPS profile for each M-Al2O3 sample, it could be concluded that the transition metal was loaded on the Al2O3 support.
For Ir-Al2O3 nanosheets, Ir 4d3/2 and Ir 4d5/2 XPS peaks were observed at BEs of 313.6 eV and 297.7 eV, respectively. These peaks were assigned to the Ir4+ oxidation state [37], in good coincidence with the XRD profiles of tetragonal IrO2, shown above. A weak shoulder of Ir 4d5/2 peak was seen around 295 eV, plausibly due to metallic Ir [37]. For Pt-Al2O3 nanosheets, Pt 4d3/2 and Pt 4d5/2 XPS peaks were observed at BEs of 332.9 eV and 314.9 eV, respectively, with a spin-orbit splitting of 18.0 eV. The XPS peaks were attributed to metallic Pt [34,35], which was well-fitting with the XRD result of metallic Pt. For Au-Al2O3 nanosheets, the Au 4f7/2 and Au 4f5/2 XPS peaks were observed at BEs of 87.3 eV and 83.6 eV, respectively, with a spin-orbit splitting of 3.7 eV. The XPS peaks were attributed metallic Au [32]. This result was in good agreement with the XRD profiles of metallic Au, shown above.
For the Al 2p XPS profiles (Supporting Information, Figure S5), a broad peak was commonly observed around 74.1 eV, attributed to Al of the Al2O3 support [32,38]. An additional peak at 75.0 eV was observed and attributed to the Al of surface Al-OH species [32,38]. For the O 1s XPS profiles (Supporting Information, Figure S5), a broad peak was commonly observed around 530.9 eV due to lattice O of Al2O3 support. A broad shoulder at 532.5 eV was attributed to oxygen defects and surface OH/H2O species [39].
The valence band (VB) spectra are shown in Figure 4 to further examine electronic structures. For the VB of bare Al2O3 nanosheets, two broad features were seen around 9 eV and 6 eV, attributed bonding 2pσ (mixed with Al 3s, Al 3p, and Al 3d) and antibonding 2pπ of the oxygen [40]. For VB spectra of M-Al2O3 nanosheets, the density of states (DOS) was observed to be closer to the Fermi level. Especially, Rh, Pd, Ir, and Pt showed more clearly new features near 2 eV below the Fermi level, attributed to the Rh 4d, Pd 4d, Ir 5d, and Pt 5d, respectively. This could be related to the higher CO oxidation activities for these metals, discussed below. However, the DOS profiles showed no explicit relationship with the photocatalytic CO2 reduction activity. The detailed roles of the overlayer elements could be understood with the aid of density functional theory.
Temperature-programmed CO oxidation profiles (Supporting Information, Figure S6) were obtained to examine thermal CO oxidation catalytic activities for bare Al2O3 and M-Al2O3 nanosheets. To evaluate the catalytic activities of the catalysts, Figure 5a,b display the CO oxidation onset temperatures for the first and the second runs, respectively. Table 1 summarizes the onset temperatures (TM-Al2O3,onset) and the temperature difference (TM-Al2O3,2nd − TM-Al2O3,1st) between the first and the second runs. The onset temperatures of Ir-, Pt-, Pd-, and Rh-loaded Al2O3 nanosheets were observed to be much lower than those of Au-, Ag-, Cu-, Co-, and Ni-loped Al2O3 nanosheets. The group 11 (Au, Ag, and Cu) and the period 4 (Co, Ni, and Cu) elements showed much poor catalytic activity on the Al2O3 support. Additionally, the onset temperatures of Au and Ag-loaded Al2O3 nanosheets were unexpectedly even higher than expected [6,10,15]. In other words, the Au- and Ag-loaded Al2O3 nanosheets showed poorer CO oxidation activity. In the first run, the Rh-Al2O3 nanosheets showed the lowest onset of 135 °C, while the Ni-Al2O3 nanosheets showed the highest onset of 490 °C. The temperature difference between the two samples was estimated to be 335 °C. In the second run, the Rh-Al2O3 nanosheets also showed the lowest onset of 172 °C while the Ni-Al2O3 nanosheets showed the highest onset of 480 °C. The temperature difference was estimated to be 308 °C. Pd, Ir, and Pt showed the CO oxidation onsets at 207 °C, 217 °C, and 216 °C, respectively, in the second run. For highly dispersed (or single atom state) 0.2 wt % Pt on mesoporous Al2O3 support, Zhang et al. reported CO oxidation onset at ~200 °C, which was in good coincidence with the present result [5]. These results clearly indicated that the CO oxidation activity was highly influenced by the nature of overlayer metal species.
Figure 5c shows the CO oxidation profiles for the first and the second runs of the selected samples (bare Al2O3, Ni-Al2O3, and Rh-Al2O3 catalysts). As seen in the Figure 5, the CO oxidation onset of Rh-Al2O3 occurred much earlier than that of bare Al2O3. The onsets of Rh-Al2O3 in the first and the second runs were observed to be 251 °C and 201 °C lower than those of bare Al2O3, respectively. However, the onset temperatures became much higher upon loading Ni.
To examine the difference in catalytic activity between the first and the second runs, Figure 5d plots the temperature differences (TM-Al2O3,2nd − TM-Al2O3,1st) in the CO oxidation onsets between the first and the second runs. In the first run, the CO oxidation reactions were performed with the as-prepared samples. In the second run, the CO oxidation reactions were performed with samples, which were already participated in the first run. Therefore, the surface states (or the catalytic-active sites) were expected to be different for the samples in the first and the second runs. The values (TM-Al2O3,2nd − TM-Al2O3,1st) are summarized in Table 1. The positive value (Figure 5d) indicated that the CO oxidation started at a higher temperature in the second run. In other words, the CO oxidation catalytic activity became lower in the second run.
For Co- and Ni-Al2O3 nanosheets, the onset temperatures in the second run were observed to be slightly lower than those in the first run. However, the other samples commonly showed higher onset temperatures in the second run, compared with the first run. This indicated that, for the latter, the catalytic activity became somewhat lower after the first run. The lower catalytic activity appeared to be mainly due to a change in crystallinity and lower catalytic-active sites.
To evaluate the roles of the transition metals in catalytic activities, compared with bare Al2O3, Figure 5e,f show the relative CO oxidation onsets (TAl2O3 − TM-Al2O3), compared with those of the first and the second runs of the bare Al2O3, respectively. The values are summarized in Table 2. In the first runs, the TAl2O3,1st − TM-Al2O3,1st values of Co and Ni showed positive, and others showed negative values. In the second runs, Co, Ni, Cu, Ag, and Au showed positive, and others showed negative values. On the basis of Figure 5e,f, the catalytic activity became poorer upon loading Co and Ni, compared with bare Al2O3. Unexpectedly, the Au, Ag, and Cu (group 11) showed somewhat higher activities in the first run but showed poorer catalytic activity in the second run, compared with the bare Al2O3 nanosheet. The Rh, Pd, Ir, and Pt showed much higher (with lowering of onset temperatures between 156 °C and 261 °C) CO oxidation activity in the first and second runs. Conclusively, the CO oxidation activity showed the order of Ni < Co < Au < Cu < Ag < Pd < Pt < Ir < Rh in the first run, and Ni < Au < Ag < Cu < Co < Ir < Pt ≈ Pd < Rh in the second run.
For CO oxidation, a simplified mechanism is described below;
M-Al2O3 + CO (g) → CO(ad)-M-Al2O3      adsorption of CO
M-Al2O3 + 1/2O2 (g) → O(ad)-M-Al2O3   dissociative adsorption of O2
CO (g) + O(ad)-M-Al2O3 → M-Al2O3 + CO2 (g)
CO (g) + HO-M-Al2O3 → M-Al2O3-H + CO2 (g)
CO (g) + O2(ad)-M-Al2O3 → CO3(ad)-M-Al2O3   carbonate formation
CO3(ad)-M-Al2O3 → O(ad)-M-Al2O3 + CO2 (g)
The CO oxidation mechanism was explained by the Langmuir-Hinshelwood mechanism [12,13] and the non-Langmuir-Hinshelwood mechanism [4,11], depending on the overlayer transition metals. In reaction (1), CO was adsorbed on metal site, and in reaction (2), oxygen was dissociatively adsorbed on the surface. In reaction (3), gaseous CO and surface O reacted to release CO2 [12,13]. When moisture was present in the reaction, the surface OH group was plausibly formed and CO might also react with the surface metal hydroxide to form the CO2 in reaction (4) [15]. On the basis of the FT-IR spectra (Supporting Information, Figure S7), surface OH groups were observed in the as-prepared samples. Therefore, reaction (4) was likely involved in the first run of CO2 formation. If H was not desorbed as H2O, the surface H was recycled as shown in reaction (4). Otherwise, the H was removed from the surface as gaseous H2O, and, thus, the reaction (4) was diminished in the second run. Reaction (5) was also reported for surfaces such as Pt/Al2O3 [4,11]. In reaction (5), CO was adsorbed on oxygenated metal atoms to initially form carbonate. Then, the carbonate dissociated to generate CO2 in reaction (6).
Photocatalytic CO2 reduction products were examined for bare Al2O3 and M-Al2O3 nanosheets and are displayed in Figure 6 [39,41,42]. Major CO2 reduction products were observed to be carbon monoxide (CO), methanol (CH3OH), and methane (CH4) with an order: CH4 < CH3OH < CO. CO was the most dominantly produced species. CH3OH showed a higher production amount compared with CH4. Hydrogen (H2) was additionally observed as a photocatalytic water splitting product during CO2 reduction. Figure 6a plots all of the product amounts (μmol/mol = ppm) for bare Al2O3 and M-Al2O3 nanosheets. As a quick glance, Ag-Al2O3 nanosheets showed the highest amounts of CO2 reduction products: 237.3 ppm for CO, 36.3 ppm for CH3OH, and 30.9 ppm for CH4, and Rh-Al2O3 nanosheets showed the highest H2 production (20.7 ppm). For the bare Al2O3 nanosheets in Figure 6b, CO, CH3OH, and CH4 were observed to be 107.5 ppm, 29.6 ppm, and 19.5 ppm, respectively. No H2 was detected. CO reduction yields (μmol/mol) in different groups of 9, 10, and 11, and with different units (μmol/g), are provided in the Supporting Information, Figures S8 and S9, respectively.
For bare Al2O3, the selectivities for CO, CH3OH, and CH4 were estimated to be 68.6%, 18.9%, and 12.5%, respectively. Upon Co- and Cu-loading, CH4 and H2 showed meaningful (>25%) enhancements. However, the amounts of CO and CH3OH showed no critical change. CO, CH3OH, and CH4 productions were enhanced by 28%, 17%, and 24% upon Ni-loading. CO was increased by 2.2× upon loading Ag in Figure 6c. CH3OH and CH4 were also increased by 1.23× and 1.58×, respectively, upon loading Ag. Rh and Pd-loadings had a smaller effect on the CO production relative to the bare support. CH3OH and CH4 productions were not meaningfully enhanced by Rh- and Pd-loadings. Instead, interestingly the H2 production was commonly observed in these metal-loadings. For Ir, Pt, and Au elements in period 6, CO productions were all decreased by metal-loadings. CH3OH productions were somewhat increased by 19% and 16% upon loading of Pt and Au, respectively. The CH4 production was only increased upon loading Pt relative to the bare substrate.
For H2 production, Ag, Pd, and Rh (in period 5) metals commonly showed H2 productions with amounts of 2.1 ppm, 3.0 ppm, and 20.7 ppm, respectively. For the metals of Co, Ni, and Cu (in period 4), the H2 production amounts were observed to be 1.9 ppm, 0 ppm, and 3.0 ppm, respectively. That is, Ni showed no H2 production. The metals of Au, Pt, and Ir in period 6 commonly showed no H2 production at all. The Rh-Al2O3 nanosheets predominantly showed the highest H2 production with an amount of 20.7 ppm.
The photocatalytic CO2 reduction mechanism is generally written as xCO2 + yH+ + ze → CaHbOc products + dH2O [41,42]. Electrons (e) and holes (h+) were generated under UVC irradiation in reaction (7). H+ ion was generated via the reactions in (8)–(11). The generation of electrons was an important factor for the multielectron processes. The mechanisms for the productions of CO (in reaction (12)), CH3OH (in reaction (13)), and CH4 (in reaction (14)) are written as below and shown in Figure 6 [38,39].
Al oxides + UVC → Al oxides (e + h+)
H2O → H+ + OH
OH + h+ → •OH
•OH + H2O + 3h+ → O2 + 3H+
H+ + e → 1/2H2
CO2 + 2H+ + 2e → CO + H2O, −0.530 V vs. standard hydrogen electrode (SHE)
CO2 + 6H+ + 6e → CH3OH + H2O, −0.380 V vs. SHE
CO2 + 8H+ + 8e → CH4 + 2H2O, −0.240 V vs. SHE
These reaction channels were closely spaced in free energy change, and, thus, the hydrogen production channel (H+ + e → 1/2H2, −0.42 V vs. SHE) occurred competitively. In the mechanism, CO2 was initially adsorbed to form COOH. The COOH was then attacked by H+ and e to generate gaseous CO. The CO production channel was only enhanced by loading Ag or Ni on Al2O3 support. CH3OH production was likely formed when surface COad underwent step-wise hydrogenation. This production was enhanced by loading Ni, Rh, Ag, Pt, or Au on Al2O3 support. CH4 production was formed via C–O bond scission of hydrogenated ≡C‒OH and new C‒H bond formation. This production was somewhat enhanced by loading Co, Ni, Cu, Ag, or Pt. The present pre-screening tests need further investigations to understand the detailed roles of the overlayer elements, with the aid of density functional theory.

4. Conclusions

In summary, γ-Al2O3 nanosheets were prepared by the solvothermal method followed by thermal calcination at 600 °C for 2 h. Transition metals (M = Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au) were loaded on Al2O3 nanosheet supports, and their thermal CO oxidation and photocatalytic CO2 reduction activities were fully tested.
The thermal CO oxidation activity showed the order of Ni < Co < Au < Cu < Ag << Pd < Pt < Ir < Rh in the first run, and Ni < Au < Ag < Cu < Co << Ir < Pt ≈ Pd < Rh in the second run. The Au, Ag, Co, Ni, and Cu elements reduced the catalytic activity on the Al2O3 support. CO oxidation activity was greatly enhanced by the loading of Ir, Pt, Pd, and Rh elements. Rh-Al2O3 nanosheets showed the highest CO oxidation activity with onset temperatures of 135 °C and 172 °C for the first and the second runs, respectively.
Photocatalytic CO2 reduction experiments were also performed to show that CO, CH3OH, and CH4 were common products with an order of CH4 (14.6–30.9 ppm range) < CH3OH (23.0–36.3 ppm range) << CO (76.5–237.3 ppm range). The highest performance was achieved after Ag-loadings with yields of 237.3 ppm for CO, 36.3 ppm for CH3OH, and 30.9 ppm for CH4, corresponding to 2.2×, 1.2×, and 1.6× enhancements, respectively, compared with those for the bare Al2O3. CO production was substantially decreased by the loading of Pd and Pt. Hydrogen production was enhanced by Rh-loadings with a yield of 20.7 ppm. Conclusively, Rh-Al2O3 and Ag-Al2O3 showed the best thermal CO oxidation and photocatalytic CO2 reduction performances, respectively, among Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au element loadings.
The present pre-screening test results could be a useful quick guide for the selection of overlayer transition metals in the groups of 9, 10, and 11 when Al2O3 is used as a support catalyst material. It also enriched the understanding of the role of an overlayer transition metal.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano11051278/s1, Figure S1: Scanning electron microscope (SEM) images for M-loaded Al2O3 nanosheets, Figure S2: Photos for M-loaded Al2O3 nanosheets, Figure S3: Optical microscope images for Al2O3 and M-loaded Al2O3 nanosheets, Figure S4: Transmission electron microscopic (TEM) and high-resolution TEM (HRTEM) images of Co-Al2O3 nanosheets, Figure S5: First and second CO oxidation profiles for Al2O3 and M-loaded Al2O3 nanosheets, Figure S6: Survey, C 1s, Al 2p, and O 1s profile for bare and M-Al2O3 nanosheets, Figure S7: FT-IR spectra for Al2O3 and M-loaded Al2O3 nanosheets before and after CO oxidation, Figure S8: CO2 reduction CO, CH4, and CH3OH yields (μmol/mol) over bare and M-loaded Al2O3 nanosheets, group 9: (Co, Rh and Ir)-Al2O3, group 10: (Ni, Pd and Pt)-Al2O3, and group 11: (Ir, Pt and Au)-Al2O3, Figure S9: CO, CH4, and CH3OH yields (μmol/g) for over bare and M-loaded Al2O3 nanosheets, (Co, Ni and Cu)-Al2O3, (Rh, Pd and Ag)-Al2O3, (Ir, Pt and Au)-Al2O3.

Author Contributions

H.J.Y. performed the material synthesis, SEM, TEM, XRD, CO oxidation, and data analysis; J.H.Y. performed the CO2 reduction and data analysis; S.J.P. performed the XPS experiments and data analysis; Y.S. designed the experiments and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was finally supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MEST; 2016R1D1A3B04930123).

Data Availability Statement

Data are available in the main text.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, Y.; Feng, Y.X.; Li, L.; Liu, J.Y.; Pan, X.L.; Liu, W.; Wei, F.F.; Cui, Y.T.; Qiao, B.T.; Sun, X.C.; et al. Identification of active sites on high-performance Pt/Al2O3 catalyst for cryogenic CO oxidation. ACS Catal. 2020, 10, 8815–8824. [Google Scholar] [CrossRef]
  2. Lou, Y.; Liu, J. CO oxidation on metal oxide supported single Pt atoms: The role of the support. Indus. Eng. Chem. Res. 2017, 56, 6916–6925. [Google Scholar] [CrossRef]
  3. Hazlett, M.J.; Moses-Debusk, M.; Parks, J.E.; Allard, L.F.; Epling, W.S. Kinetic and mechanistic study of bimetallic Pt-Pd/Al2O3 catalysts for CO and C3H6 oxidation. Appl. Catal. B 2017, 202, 404–417. [Google Scholar] [CrossRef] [Green Version]
  4. Newton, M.A.; Ferri, D.; Smolentsev, G.; Marchionni, V.; Nachtegaal, M. Kinetic Studies of the Pt Carbonate-Mediated, Room-Temperature Oxidation of Carbon Monoxide by Oxygen over Pt/Al2O3 Using Combined, Time-Resolved XAFS, DRIFTS, and Mass Spectrometry. J. Am. Chem. Soc. 2016, 138, 13930–13940. [Google Scholar] [CrossRef]
  5. Zhang, Z.; Zhu, Y.; Asakura, H.; Zhang, B.; Zhang, J.; Zhou, M.; Han, Y.; Tanaka, T.; Wang, A.; Zhang, T.; et al. Thermally stable single atom Pt/m-Al2O3 for selective hydrogenation and CO oxidation. Nat. Commun. 2017, 8, 16100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Ananth, A.; Jeong, R.H.; Boo, J.-H. Preparation, characterization and CO oxidation performance of Ag2O/γ-Al2O3 and (Ag2O+RuO2)/γ-Al2O3 catalysts. Surfaces 2020, 3, 251–264. [Google Scholar] [CrossRef]
  7. Yang, T.; Fukuda, R.; Hosokawa, S.; Tanaka, T.; Sakaki, S.; Ehara, M. A theoretical investigation on CO oxidation by single-atom catalysts M1 /γ-Al2O3 (M = Pd, Fe, Co, and Ni). ChemCatChem 2017, 9, 1222–1229. [Google Scholar] [CrossRef] [Green Version]
  8. Gao, H. CO oxidation mechanism on the γ-Al2O3 supported single Pt atom: First principle study. Appl. Surf. Sci. 2016, 379, 347–357. [Google Scholar] [CrossRef]
  9. Li, Q.; Wei, Y.; Sa, R.; Ma, Z.; Wu, K. A novel Pd3O9@Al2O3 catalyst under a hydroxylated effect: High activity in the CO oxidation reaction. Phys. Chem. Chem. Phys. 2015, 17, 32140–32148. [Google Scholar] [CrossRef] [PubMed]
  10. Li, Z.-Y.; Yuan, Z.; Li, X.-N.; Zhao, Y.-X.; He, S.-G. CO oxidation catalyzed by single gold atoms supported on aluminum oxide clusters. J. Am. Chem. Soc. 2014, 136, 14307–14313. [Google Scholar] [CrossRef]
  11. Moses-DeBusk, M.; Yoon, M.; Allard, L.F.; Mullins, D.R.; Wu, Z.; Yang, X.; Veith, G.; Stocks, G.M.; Narula, C.K. CO oxidation on supported single Pt atoms: Experimental and ab initio density functional studies of CO interaction with Pt atom on θ-Al2O3 (010) surface. J. Am. Chem. Soc. 2013, 135, 12634–12645. [Google Scholar] [CrossRef] [PubMed]
  12. Zhou, Y.; Wang, Z.; Liu, C. Perspective on CO oxidation over Pd-based catalysts. Catal. Sci. Technol. 2015, 5, 69–81. [Google Scholar] [CrossRef]
  13. Satsuma, A.; Osaki, K.; Yanagihara, M.; Ohyama, J.; Shimizu, K. Activity controlling factors for low-temperature oxidation of CO over supported Pd catalysts. Appl. Catal. B 2013, 132–133, 511–518. [Google Scholar] [CrossRef]
  14. Haneda, M.; Todo, M.; Nakamura, Y.; Hattori, M. Effect of Pd dispersion on the catalytic activity of Pd/Al2O3 for C3H6 and CO oxidation. Catal. Today 2017, 281, 447–453. [Google Scholar] [CrossRef]
  15. Gavril, D. CO oxidation on nanosized Au/Al2O3 by surface hydroxyl groups and in the absence of O2, studied by inverse gas chromatography. Catal. Today 2015, 244, 36–46. [Google Scholar] [CrossRef]
  16. Han, S.W.; Kim, D.H.; Jeong, M.-G.; Park, K.J.; Kim, Y.D. CO oxidation catalyzed by NiO supported on mesoporous Al2O3 at room temperature. Chem. Eng. J. 2016, 283, 992–998. [Google Scholar] [CrossRef]
  17. Kwak, J.H.; Hu, J.; Mei, D.; Yi, C.-W.; Kim, D.H.; Peden, C.H.F.; Allard, L.F.; Szanyi, J. Coordinatively Unsaturated Al3+ Centers as Binding Sites for Active Catalyst Phases of Platinum on γ-Al2O3. Science 2009, 325, 1670–1673. [Google Scholar] [CrossRef]
  18. Shen, Y.; Lu, G.; Guo, Y.; Wang, Y. A synthesis of high-efficiency Pd-Cu-Clx/Al2O3 catalyst for low temperature CO oxidation. Chem. Commun. 2010, 46, 8433–8435. [Google Scholar] [CrossRef]
  19. Gribov, E.N.; Zavorotynska, O.; Agostini, G.; Vitillo, J.G.; Ricchiardi, G.; Spoto, G.; Zecchina, A. FTIR spectroscopy and thermodynamics of CO and H2 adsorbed on γ-, δ- and α-Al2O3. Phys. Chem. Chem. Phys. 2010, 12, 6474–6482. [Google Scholar] [CrossRef]
  20. Kwak, J.H.; Kovarik, L.; Szanyi, J. CO2 reduction on supported Ru/Al2O3 catalysts: Cluster size dependence of product selectivity. ACS Catal. 2013, 3, 2449–2455. [Google Scholar] [CrossRef]
  21. Zhao, H.; Zheng, X.; Feng, X.; Li, Y. CO2 reduction by plasmonic Au nanoparticle-decorated TiO2 photocatalyst with an ultrathin Al2O3 interlayer. J. Phys. Chem. C 2018, 122, 18949–18956. [Google Scholar] [CrossRef]
  22. Zhao, H.; Chen, J.; Rao, G.; Deng, W.; Li, Y. Enhancing photocatalytic CO2 reduction by coating an ultrathin Al2O3 layer on oxygen deficient TiO2 nanorods through atomic layer deposition. Appl. Surf. Sci. 2017, 404, 49–56. [Google Scholar] [CrossRef] [Green Version]
  23. Wu, S.; Li, Y.; Zhang, Q.; Hu, Q.; Wu, J.; Zhou, C.; Zhao, X. Formation of NiCo alloy nanoparticles on Co doped Al2O3 leads to high fuel production rate, large light-to-fuel efficiency, and excellent durability for photothermocatalytic CO2 reduction. Adv. Energy Mater. 2020, 10, 2002602. [Google Scholar] [CrossRef]
  24. Méndez-Mateos, D.; Barrio, V.L.; Requies, J.M.; Cambra, J.F. Effect of the addition of alkaline earth and lanthanide metals for the modification of the alumina support in Ni and Ru catalysts in CO2 methanation. Catalysts 2021, 11, 353. [Google Scholar] [CrossRef]
  25. Mihet, M.; Lazar, M.D. Methanation of CO2 on Ni/γ-Al2O3: Influence of Pt, Pd or Rh promotion. Catal. Today 2018, 306, 294–299. [Google Scholar] [CrossRef]
  26. Liang, C.; Hu, X.; Wei, T.; Jia, P.; Zhang, Z.; Dong, D.; Zhang, S.; Liu, Q.; Hu, G. Methanation of CO2 over Ni/Al2O3 modified with alkaline earth metals: Impacts of oxygen vacancies on catalytic activity. Int. J. Hydrog. Energy 2019, 44, 8197–8213. [Google Scholar] [CrossRef]
  27. Chein, R.-Y.; Wang, C.-C. Experimental study on CO2 methanation over Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 catalysts. Catalysts 2020, 10, 1112. [Google Scholar] [CrossRef]
  28. Lou, Y.; Liu, J. A highly active Pt–Fe/γ-Al2O3 catalyst for preferential oxidation of CO in excess of H2 with a wide operation temperature window. Chem. Commun. 2017, 53, 9020–9023. [Google Scholar] [CrossRef] [Green Version]
  29. Chen, Y.; Lin, J.; Li, L.; Pan, X.; Wang, X.; Zhang, T. Local structure of Pt species dictates remarkable performance on Pt/Al2O3 for preferential oxidation of CO in H2. Appl. Catal. B 2021, 282, 119588. [Google Scholar] [CrossRef]
  30. Han, D.; Lee, D. Morphology controlled synthesis of γ-Al2O3 nano-crystallites in Al@Al2O3 core–shell micro-architectures by interfacial hydrothermal reactions of Al metal substrates. Nanomaterials 2021, 11, 310. [Google Scholar] [CrossRef] [PubMed]
  31. Lee, S.; Kang, J.S.; Leung, K.T.; Kim, S.K.; Sohn, Y. Magnetic Ni-Co alloys induced by water gas shift reaction, Ni-Co oxides by CO oxidation and their supercapacitor applications. Appl. Surf. Sci. 2016, 386, 393–404. [Google Scholar] [CrossRef]
  32. NIST Standard Reference Database 20, Version 4.1. 2012. Available online: https://srdata.nist.gov/xps/ (accessed on 1 March 2021).
  33. Akhavan, O.; Azimirad, R.; Safa, S.; Hasani, E. CuO/Cu(OH)2 Hierarchical nanostructures as bactericidal photocatalysts. J. Mater. Chem. 2011, 21, 9634–9640. [Google Scholar] [CrossRef]
  34. Renéme, Y.; Dhainaut, F.; Trentesaux, M.; Ravanbakhsh, B.; Granger, P.; Dujardin, C.; Gengembre, L.; De Cola, P.L. XPS investigation of surface changes during thermal aging of natural gas vehicle catalysts: Influence of Rh addition to Pd. Surf. Interface Anal. 2010, 42, 530–535. [Google Scholar] [CrossRef]
  35. Ivanova, A.S.; Slavinskaya, E.M.; Gulyaev, R.V.; Zaikovskii, V.I.; Stonkus, O.A.; Danilova, I.G.; Plyasova, L.M.; Polukhina, I.A.; Boronin, A.I. Metal-support interactions in Pt/Al2O3 and Pd/Al2O3 catalysts for CO oxidation. Appl. Catal. B 2010, 97, 57–71. [Google Scholar] [CrossRef]
  36. Kang, J.G.; Sohn, Y. Interfacial nature of Ag nanoparticles supported on TiO2 photocatalysts. J. Mater. Sci. 2012, 47, 824–832. [Google Scholar] [CrossRef]
  37. Freakley, S.J.; Ruiz-Esquius, J.; Morgan, D.J. The X-ray photoelectron spectra of Ir, IrO2 and IrCl3 revisited. Surf. Interface Anal. 2017, 49, 794–799. [Google Scholar] [CrossRef]
  38. Yoon, H.J.; Kim, S.K.; Lee, S.W.; Sohn, Y. Stalagmite Al(OH)3 growth on aluminum foil surface by catalytic CO2 reduction with H2O. Appl. Surf. Sci. 2018, 450, 85–90. [Google Scholar] [CrossRef]
  39. Yang, J.H.; Park, S.J.; Rhee, C.K.; Sohn, Y. Photocatalytic CO2 reduction and electrocatalytic H2 evolution over Pt(0,II,IV)-loaded oxidized Ti sheets. Nanomaterials 2020, 10, 1909. [Google Scholar] [CrossRef]
  40. Filatova, E.O.; Konashuk, A.S. Interpretation of the changing the band gap of Al2O3 depending on its crystalline form: Connection with different local symmetries. J. Phys. Chem. C 2015, 119, 20755–20761. [Google Scholar] [CrossRef]
  41. Sun, Z.; Ma, T.; Tao, H.; Fan, Q.; Han, B. Fundamentals and challenges of electrochemical CO2 reduction using two-dimensional materials. Chem 2017, 3, 560–587. [Google Scholar] [CrossRef] [Green Version]
  42. Sohn, Y.; Huang, W.; Taghipour, F. Recent progress and perspectives in the photocatalytic CO2 reduction of Ti-oxide-based nanomaterials. Appl. Surf. Sci. 2017, 396, 1696–1711. [Google Scholar] [CrossRef]
Figure 1. Scanning electron microscope (SEM) (a,a1,b), transmission electron microscope (TEM) (b2), high resolution TEM (b3) images of the as-synthesized Al-precursor (a,a1) and Al2O3 (b,b2,b3), and the structure projection (b1) of the (002) and (022) planes for cubic phase γ-Al2O3.
Figure 1. Scanning electron microscope (SEM) (a,a1,b), transmission electron microscope (TEM) (b2), high resolution TEM (b3) images of the as-synthesized Al-precursor (a,a1) and Al2O3 (b,b2,b3), and the structure projection (b1) of the (002) and (022) planes for cubic phase γ-Al2O3.
Nanomaterials 11 01278 g001
Figure 2. SEM (a,b), TEM (a1,b1), HRTEM (a2,b2) images of selected Rh-Al2O3 (a,a1,a2) and Ni-Al2O3 (b,b1,b2) nanosheets. Insets of Figure 2(a2) show the fast Fourier transform (FFT) pattern of the HRTEM image, and the structure projection of the (114) and (200) planes for Rh2O3.
Figure 2. SEM (a,b), TEM (a1,b1), HRTEM (a2,b2) images of selected Rh-Al2O3 (a,a1,a2) and Ni-Al2O3 (b,b1,b2) nanosheets. Insets of Figure 2(a2) show the fast Fourier transform (FFT) pattern of the HRTEM image, and the structure projection of the (114) and (200) planes for Rh2O3.
Nanomaterials 11 01278 g002
Figure 3. XRD profiles bare Al2O3 and M-loaded Al2O3 nanosheets.
Figure 3. XRD profiles bare Al2O3 and M-loaded Al2O3 nanosheets.
Nanomaterials 11 01278 g003
Figure 4. Co 2p, Ni 2p, Cu 2p, Rh 3d, Pd 3d, Ag 3d, Ir 4d, Pt 4d, Au 4d, and VB profiles of Co-Al2O3, Ni-Al2O3, Cu-Al2O3, Rh-Al2O3, Pd-Al2O3, Ag-Al2O3, Ir-Al2O3, Pt-Al2O3, and Au-Al2O3 nanosheets.
Figure 4. Co 2p, Ni 2p, Cu 2p, Rh 3d, Pd 3d, Ag 3d, Ir 4d, Pt 4d, Au 4d, and VB profiles of Co-Al2O3, Ni-Al2O3, Cu-Al2O3, Rh-Al2O3, Pd-Al2O3, Ag-Al2O3, Ir-Al2O3, Pt-Al2O3, and Au-Al2O3 nanosheets.
Nanomaterials 11 01278 g004
Figure 5. CO oxidation onsets for the first (a) and second (b) runs of M-loaded Al2O3 nanosheets. Selected first and second run CO oxidation profiles with temperature (c) for bare, Rh- and Ni-Al2O3 nanosheets. Differences in CO oxidation onsets for the first and the second runs (d) and relative CO oxidation onset temperatures relative to that of bare nanosheets (e,f).
Figure 5. CO oxidation onsets for the first (a) and second (b) runs of M-loaded Al2O3 nanosheets. Selected first and second run CO oxidation profiles with temperature (c) for bare, Rh- and Ni-Al2O3 nanosheets. Differences in CO oxidation onsets for the first and the second runs (d) and relative CO oxidation onset temperatures relative to that of bare nanosheets (e,f).
Nanomaterials 11 01278 g005
Figure 6. CO2 reduction CO, CH4, and CH3OH yields (μmol/mol) over bare and M-loaded Al2O3 nanosheets (a), (Co, Ni, and Cu)-Al2O3 (b), (Rh, Pd, and Ag)-Al2O3 (c), (Ir, Pt, and Au)-Al2O3 (d), and CO2 reduction mechanism.
Figure 6. CO2 reduction CO, CH4, and CH3OH yields (μmol/mol) over bare and M-loaded Al2O3 nanosheets (a), (Co, Ni, and Cu)-Al2O3 (b), (Rh, Pd, and Ag)-Al2O3 (c), (Ir, Pt, and Au)-Al2O3 (d), and CO2 reduction mechanism.
Nanomaterials 11 01278 g006
Table 1. CO oxidation onset temperatures (TM-Al2O3,onset) in the first and second runs. Differences in CO oxidation onset temperatures (TM-Al2O3,2nd − TM-Al2O3,1st) between the first and second runs.
Table 1. CO oxidation onset temperatures (TM-Al2O3,onset) in the first and second runs. Differences in CO oxidation onset temperatures (TM-Al2O3,2nd − TM-Al2O3,1st) between the first and second runs.
Group
#9
First
Run
Second
Run
Diff.Group
#10
First
Run
Second
Run
Diff.Group
#11
First
Run
Second
Run
Diff.
Co390385−5Ni490480−10Cu36038929
Rh13517237Pd2012076Ag34039757
Ir19521722Pt1972069Au37541540
Table 2. Differences in CO oxidation onset temperatures (TAl2O3,onset − TM-Al2O3,onset) in the first and second runs, compared with that of bare Al2O3 nanosheets. The CO oxidation onset temperatures of bare Al2O3 were 386 °C and 373 °C for the first and the second runs, respectively.
Table 2. Differences in CO oxidation onset temperatures (TAl2O3,onset − TM-Al2O3,onset) in the first and second runs, compared with that of bare Al2O3 nanosheets. The CO oxidation onset temperatures of bare Al2O3 were 386 °C and 373 °C for the first and the second runs, respectively.
Group
#9
First
Run
Second
Run
Group
#10
First
Run
Second
Run
Group
#11
First
Run
Second
Run
Co412Ni104107Cu−2616
Rh−251−201Pd−185−166Ag−4624
Ir−191−156Pt−189−167Au−1142
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yoon, H.J.; Yang, J.H.; Park, S.J.; Sohn, Y. Thermal CO Oxidation and Photocatalytic CO2 Reduction over Bare and M-Al2O3 (M = Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au) Cotton-Like Nanosheets. Nanomaterials 2021, 11, 1278. https://doi.org/10.3390/nano11051278

AMA Style

Yoon HJ, Yang JH, Park SJ, Sohn Y. Thermal CO Oxidation and Photocatalytic CO2 Reduction over Bare and M-Al2O3 (M = Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au) Cotton-Like Nanosheets. Nanomaterials. 2021; 11(5):1278. https://doi.org/10.3390/nano11051278

Chicago/Turabian Style

Yoon, Hee Jung, Ju Hyun Yang, So Jeong Park, and Youngku Sohn. 2021. "Thermal CO Oxidation and Photocatalytic CO2 Reduction over Bare and M-Al2O3 (M = Co, Ni, Cu, Rh, Pd, Ag, Ir, Pt, and Au) Cotton-Like Nanosheets" Nanomaterials 11, no. 5: 1278. https://doi.org/10.3390/nano11051278

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop