Next Article in Journal
Recent Advances in Wearable Optical Sensor Automation Powered by Battery versus Skin-like Battery-Free Devices for Personal Healthcare—A Review
Next Article in Special Issue
Preparation and Photocatalytic Properties of Anatase TiO2 with Hollow Hexagonal Frame Structure
Previous Article in Journal
Novel Nano-Engineered Biomaterials for Bone Tissue Engineering
Previous Article in Special Issue
g-C3N4: Properties, Pore Modifications, and Photocatalytic Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Efficient Photothermal Reduction of CO2 on Pd2Cu Dispersed TiO2 Photocatalyst and Operando DRIFT Spectroscopic Analysis of Reactive Intermediates

by
Munirathinam Elavarasan
1,
Willie Yang
2,
Sethupathi Velmurugan
3,
Jyy-Ning Chen
1,
Thomas C.-K. Yang
1,3,* and
Toshiyuki Yokoi
2,*
1
Department of Chemical Engineering and Biotechnology, National Taipei University of Technology, No. 1, Section 3, Chung-Hsiao East Road, Taipei 106, Taiwan
2
Institute of Innovative Research, Tokyo Institute of Technology, Yokohama 226-8503, Japan
3
Precision Analysis and Material Research Center, National Taipei University of Technology, No. 1, Section 3, Chung-Hsiao East Road, Taipei 106, Taiwan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(3), 332; https://doi.org/10.3390/nano12030332
Submission received: 14 December 2021 / Revised: 8 January 2022 / Accepted: 15 January 2022 / Published: 21 January 2022
(This article belongs to the Special Issue Semiconductor-Based Nanomaterials for Photocatalytic Applications)

Abstract

:
The photocatalytic conversion of CO2 to fuels using solar energy presents meaningful potential in the mitigation of global warming, solar energy conversion, and fuel production. Photothermal catalysis is one promising approach to convert chemically inert CO2 into value-added chemicals. Herein, we report the selective hydrogenation of CO2 to ethanol by Pd2Cu alloy dispersed TiO2 (P25) photocatalyst. Under UV-Vis irradiation, the Pd2Cu/P25 showed an efficient CO2 reduction photothermally at 150 °C with an ethanol production rate of 4.1 mmol g−1 h−1. Operando diffuse reflectance infrared Fourier transform (DRIFT) absorption studies were used to trace the reactive intermediates involved in CO2 hydrogenation in detail. Overall, the Cu provides the active sites for CO2 adsorption and Pd involves the oxidation of H2 molecule generated from P25 and C–C bond formation.

1. Introduction

An anthropogenic climate dramatically increases the CO2 level in the atmosphere, which can lead to the deterioration of the environment. The utilization of CO2 is an alternative route to reduce it in the environment, and it can also be converted into valuable fuels and chemicals [1,2,3,4,5]. The photocatalytic hydrogenation of CO2 is an efficient technique for eco-friendly conversion, but CO2 activation and selective hydrogenation remain a challenge [6,7]. This is due to the high thermodynamic stability and chemical inertness of CO2 [8,9]. Therefore, the activation and hydrogenation of CO2 require enormous energy to break the O=C=O bond and implement the insertion of the C-H bond. Previous research studies show that the hydrogenation of CO2 resulted in numerous other products. Mostly, the CO2 reduction processes resulted in the production of carbon monoxide (CO), a more extremely poisonous gas than CO2. Moreover, it is very rare to produce value added chemicals, such as methane, methanol, acids, and hydrocarbons [10,11,12,13,14,15,16,17]. Therefore, the selective hydrogenation of CO2 to ethanol has received great attention, and this can be achieved via electrochemical and thermochemical processes [18,19]. However, no appreciable development in the field of photo- and photothermal catalysts has been observed, which offers a sustainable method for CO2 to ethanol conversion.
Currently, palladium (Pd) and platinum (Pt) based catalysts are demonstrated for the selective hydrogenation of CO2 to ethanol [19,20,21]. In such processes, the Pd metal is cheaper than Pt, and Pd-based catalysts have worked well in the C–C coupling reactions as well [22]. Besides, the C–C coupling reaction step is the most important step to control the yield of ethanol [23]. Copper (Cu) is the cheapest and most easily available material, and it has greater affinity for CO2 gas. Moreover, oxides of Cu represent an efficient photocatalyst with excellent redox potential for water splitting (water oxidation) in the photocatalytic hydrogen evaluation reaction [24]. The Cu/Fe2O3 catalyst is a promising catalyst for water-splitting in thermochemical reactions, to produce H2 and capture CO2 at the same time [25]. TiO2 has been proven to be a most promising material in photocatalytic water splitting, pollutant degradation, self-cleaning agents, and many other fields [26,27]. The effect of reaction temperature on photodegradation performance was studied over Pd/TiO2 and Cu/TiO2 photocatalysts. The result revealed that Pd/TiO2 performed well during the photodegradation of methylene blue. [28] Therefore, the combination of two attractive elements is an efficient way to produce ethanol. Previous studies have demonstrated the various combinations of Pd-Cu nanoparticles that are highly selective and active catalysts for the production of ethanol from CO2 by thermochemical reaction [19]. Although selective hydrogenation produces valuable chemicals, the mechanism behind this selectivity is still an uncovered area. Therefore, the mechanism of selective CO2 reduction was probed by operando diffuse reflectance infrared Fourier transform absorption spectroscopy (DRIFTS).
The DRIFT study is used to detect the molecules and ions at their place of origin where the chemicals and substances are being produced in a chemical reaction. DRIFT analysis is a widely used technique for in-situ studies concerning the photodegradation of adsorbed dye and tracing the gas adsorbed reaction mechanism on the catalytic surface. It has also been used in agro-food industries due to its analytical compatibility with physicochemical properties of food commodities [29,30]. Moreover, DRIFT analysis was applicable to test non-transparent and indispersable materials. In our previous work, we have demonstrated photocatalytic efficiency and the reaction mechanism using DRIFT in the oxidation of ethanol [31]. The solid-state photodegradation mechanism was evaluated by DRIFT for the surface adsorbed AB-9 dye on titanium dioxide (TiO2) [32]. This technique was used to investigate the electron transfer and energy conversion processes at the interface, adsorbed species, and dynamics of photo-generated electrons on nitrogen-doped TiO2 in dye-sensitized solar cells (DSSC) [33]. Moreover, the DRIFT analysis was used to identify the reaction mechanism of CO2 hydrogenation [19,20,34].
In this work, we investigate the selective hydrogenation of CO2 to ethanol by Pd2Cu/P25 photocatalyst. A facile one-pot synthesis procedure was used to obtain the Pd2Cu alloys on TiO2 nanoparticles. The physicochemical properties of the as prepared Pd2Cu/P25 catalysts were analyzed using various analytical techniques. The hydrogenation of CO2 was tested using as-prepared Pd2Cu/P25 catalysts following the photothermal method. Moreover, the reactive intermediates and reaction mechanism of hydrogenation of CO2 to ethanol were determined from operando DRIFT analysis.

2. Materials and Methods

2.1. Chemicals

Palladium (II) bis(acetylacetonate) (Pd(acac)2), 99%), Oleylamine (70%), ethyl alcohol (99.9%), and acetone (99.9%) were purchased from Sigma Aldrich. Copper chloride dihydrate (CuCl2·2H2O, 99%), triphenylphosphine (TPP, 99%), and commercial P25 (99.9% grade) were obtained from Nacalai Tesque, Acros Organics, and Degussa, respectively. The gases deployed in this research, such as high purity nitrogen (N2, 99.999%), carbon dioxide (CO2, 99.5%), and helium (He, 99.99%), were procured from Yusheng gas company, Taipei, Taiwan.

2.2. Preparation of Pd2Cu/P25 Photocatalyst

For the synthesis of Pd2Cu/P25 photocatalyst, 10 mg of Pd(acac)2 and 55 mg of CuCl2·2H2O were dissolved in 12 mL of Oleylamine in a three-neck flask. Additionally, 90 mg of triphenylphosphine (TPP) was added. The flask was heated to 80 °C under continuous stirring in an N2 atmosphere for about 20 min until the chemicals were thoroughly dissolved. The temperature was then gradually raised to 170 °C, and the system was kept at this temperature for 4 h. Then, 100 mg of TiO2 (P25) was added to this suspension and kept for 20 min, before being naturally cooled down to room temperature. The precipitate was separated and washed with a mixture of ethyl alcohol and acetone. The obtained precipitate dried overnight at 85 °C, followed by calcination at 450 °C for 4 h.

2.3. DRIFT Analysis

The DRIFT experiments were conducted by Perkin Elmer FT-IR spectroscopy. The custom-designed Harrick DRIFT cell was utilized for this study. Approximately 40 mg of the photocatalyst was placed in a DRIFT cell sample holder. The sample was covered by ZnSe/CaF2 windows for IR radiation and a light source was passed through respectively. The DRIFT chamber was filled with N2 gas for 20 min to remove impurities, before the background spectrum of the sample was acquired. The temperature of the sample was controlled using a thermocouple attached to the DRIFT cell. The wavenumber resolution was fixed at 4 cm−1, and the experiments were carried out in the range of 500–4000 cm−1. For the CO2 reduction experiment, the CO2 was purged into the chamber with the water vapor at 10 cc/s for 120 min. An outlet of the reaction chamber was connected with the mass spectrum during the collection of DRIFT spectra. Meanwhile, the mass spectra wer also collected (QMS 100 Series, capillary; PEEK, and glass-lined plastic inlet type, pressure 10 mbar to 1 bar, Quadrupole type spectrometer, Faraday cup standard electron multiplier optional detector) [31].

2.4. Photothermal CO2 Reduction by the Reactor

The photothermal CO2 reduction was performed in a custom-made reactor. Firstly, 40 mg of Pd2Cu/P25 was dispersed in 50 mL of deionized (DI) water by ultrasonicator, then poured into the reactor. Therein, the magnetic stirrer continuously stirred the dispersion. Subsequently, the CO2 was purged into the reactor and saturated for 30 min, and then the reactor was closed perfectly. The UV-Vis (350 to 750 nm) light source (with an intensity of 380.8 μW) was used to irradiate the reactor. The reaction was kept at 80 °C by the hotplate. During each measurement, 0.1 μL of the reaction mixture was collected from the reaction and analyzed by gas chromatography (GC Model 6890 N/MS Model 5973, Agilent Technologies, Poway, CA, USA).

2.5. Characterizations

The crystal structures of the as-prepared samples were investigated by the X-ray diffractometer (XRD) with Cu-Kα radiation (λ = 1.5406 Å). The Raman spectra and UVDRS spectra of the samples were recorded using confocal Raman spectroscopy (ACRON Technology, Taipei, Taiwan) and UV-Vis-NIR spectroscopy (Agilent Technology, Santa Clara, CA, USA), respectively. The back scattered images of the as-synthesized samples were recorded by scanning electron microscopy (FESEM, JEOL-JSM-7610F, Tokyo, Japan). High-resolution transmission electron microscopy (HRTEM) images were obtained from JEOL TEM (JEM-2010) at 200 kV. The element composition was recorded on the energy dispersive spectrum EDS (HORIBA, X-act, Tokyo, Japan). The JEOL JPS-9030 (Tokyo, Japan) X-ray photoelectron spectrometer (XPS) was deployed to analyze the surface properties of the prepared samples. The dissolved solution of the as prepared photocatalyst was analyzed by the inductively coupled plasma-atomic emission spectrometer (ICP-AES, Shimadzu ICPE-9000, Kyoto, Japan) for the quantitative elemental contents in each sample.

2.6. Dark Adsorption by CO2 with Transmission FTIR

The dark adsorption experiment was carried out in the transmission mode by a spectrometer (JASCO FT-IR 4600, Tokyo, Japan) equipped with a Mercury Cadmium Telluride detector. The sample was pressed into a self-supporting disk (20 mm diameter, ca. 30 mg) and placed in an IR cell attached to a gas-phase closed circulatory system for observation. The sample was pretreated at 723 K for 1 h under vacuum and then probed by a fixed amount of CO2 vapor from 5 to 500 Pa at ambient temperature. After the dark adsorption at 500 Pa, the system was evacuated to vacuum again to remove free CO2 and other free carbonyl species, to investigate the chemically adsorbed CO2 species on the samples.

3. Results and Discussion

3.1. Material Characterizations

Figure 1a shows the X-ray diffraction patterns of Pd2Cu alloy on P25 along with the diffraction pattern of pristine P25. The observed crystalline peaks at 41.11, 48.46, and 71.02° are attributed to the (111), (200), and (220) planes of PdCu alloy that are matched with the ICDD No. 00-048-1551 [35]. The obtained crystalline parameters indicate that the Pd2Cu alloy is crystallized in the body-centered cubic crystal. The crystal structure of Pd2Cu is illustrated in the inset of Figure 1a.
Moreover, the remaining peaks at 25.38, 38.23, 48.47, 54.71, 56.01, 27.58, 36.90, 41.60, and 55.0° are assigned to the (101), (004), (200), (105), (211), (110), (101), (210), and (211) planes of TiO2 (P25) respectively [31]. It can be seen from XRD peaks that the intensities of Pd2Cu peaks are dominated by the P25 crystalline peaks. This is because of the low loading level and high dispersion of Pd2Cu on P25. Figure 1b shows the Raman spectra of Pd2Cu/P25 and P25. The typical peaks appearing at 142, 397, 533, and 651 cm−1 correspond to the Eg, B1g, A1g, and Eg modes of TiO2 (P25), respectively. The additional peaks that appeared at 218 and 279 cm−1 are correlated to the specific interactions between PdCu alloys and P25. The PdCu alloys have no vibrations. However, the corresponding oxides of Pd and Cu revealed significant peaks. The vibrations at 398 and 650 cm−1 are increased for the Pd2Cu/P25 when compared with P25. This might be attributed to the stretching vibrations of Pd-O and Cu-O. Moreover, the light absorption properties of Pd2Cu/P25 were investigated by UV-Vis spectra. As shown in Figure 1c, the UV-visible absorption spectra of Pd2Cu/P25 exhibited stronger absorption in the visible range when compared to the P25. The indirect bandgap of Pd2Cu/P25 is estimated from the plots of (αhv)1/2 vs. hv (energy of light), which shows 3.0 eV for both Pd2Cu/P25 and P25 (Figure 1c, inset). Furthermore, the FESEM backscattered electron (BSE) image gives the topography of the P25 supported Pd2Cu alloys.
Figure 1d shows the BSE image of Pd2Cu/P25, wherein the Pd2Cu alloy shows a stronger backscatter electron signal, which can be seen as dark spots. P25 exhibits a weaker electron signal. This BSE image can help to distinguish the Pd2Cu alloy and P25 in the highly dispersed Pd2Cu/P25 catalyst. Figure 2a displays the TEM images of Pd2Cu alloy on the P25 nanoparticles, wherein the 5 ± 1 nm-sized Pd2Cu nanodots are highly dispersed on 20 ± 5 nm-sized P25 nanoparticles. The high magnification images (Figure 2b,c) clearly show the mono dispersion of Pd2Cu nanodots and their lattice fringes (Figure 2d). Figure 2e shows the SAED patterns of Pd2Cu/P25, which are taken at a random angle to view clear ED patterns. The obtained ring patterns are indexed to the (101), (200), and (111), (311) planes, which are in accordance with the XRD patterns of P25 and Pd2Cu, respectively. The uniform distribution and the ratio of elemental percentage are investigated using the EDX and mapping studies. Figure 2f reveals 82% of P25 and 18% of Pd2Cu alloy in the photocatalyst, respectively. Moreover, elemental mapping of Pd2Cu/P25 was performed and the results are shown in Figure 2g–j. Thus, it confirmed that the Pd2Cu nanodots are uniformly distributed on the P25 nanoparticles, supporting better photothermal CO2 conversion.
The X-ray photoelectron spectra (XPS) were used to explore the surface properties of Pd2Cu/P25. The narrow scan Pd 3d spectrum shows two peaks for spin-orbits 3d5/2 and 3d3/2 at 339 and 345 eV, corresponding to the metallic Pd0 respectively (Figure 3a). Moreover, the deconvolution spectra of Pd 3d show an additional peak for the Pd2+, which is attributed to PdO on the surface [36]. Figure 3b shows the narrow scan spectrum of Cu 2p. The binding energies at 937 and 957 eV correspond to the 2p3/2 and 2p1/2 spin-orbits, respectively. The Cu 2p spectrum was deconvoluted into four peaks. The two main peaks denote Cu in the metallic state, while the minor peaks denote Cu in +2 oxidation states. Similar to Pd, Cu also exhibited CuO on its surface. Moreover, the peak intensity of Cu 2p was lower than that of Pd, which is in accordance with EDX results. Additionally, Figure 3c,d shows the narrow scan peaks of Ti 2p and O 1s, which are indexed to the typical peaks of TiO2 (P25). The Ti 2p spectra shows two peaks at the binding energies of 462.1 and 467.8 eV belonging to the 2p3/2 and 2p1/2 orbitals, respectively. The deconvoluted O 1s peak reveals two binding energies at 533.5 and 535.3 eV, corresponding to the surface hydroxyl groups and lattice oxygen atoms respectively. Overall, these characterizations confirm the formation of Pd2Cu alloy on the P25 support. The amount of metals proportion in the photocatalyst were further confirmed by ICP-AES. For the analysis, the powder of the as-prepared catalyst was soaked in the aqua regia solutions for 6 hours and then mixed with 30wt% HF for overnight. The results were summarized in Table 1.

3.2. Photothermal CO2 Reduction

The photothermal hydrogenation of CO2 over a Pd2Cu/P25 catalyst was evaluated using the photoreactor. The detailed experimental procedure is presented in the experimental section and the schematic diagram of the reactor setup is shown in Figure 4a. The reaction was continued for up to six hours, while the sample was taken from the reaction mixture and subjected to the GCMS analysis. In the first hour, no products were seen in the reaction mixture. Afterwards, the GCMS shows the peaks for ethanol, as shown in Figure 4b. The typical fragmentation of mass 45 and 46 m/z correspond to the [C2H5O]+ and [C2H5OH]+ respectively [37]. The ethanol yield is shown in Table 2, and the yield was compared with other reported catalysts. The photocatalytic CO2 reduction of Pd2Cu/P25 was also compared with pristine P25. The P25 showed no signal for ethanol production. Herein, the P25 (TiO2) oxidized the water at its valence band, and the electrons were utilized by Pd2Cu at the conduction band. Further, the Pd2Cu has reduced the CO2 by capturing the H+ from aqueous media. Therefore, without Pd2Cu, it is difficult to produce ethanol by P25 alone. Figure 4c shows the schematic representation of the overall reaction process. Furthermore, the reaction mechanism of CO2 reduction and the intermediate products were explored by the DRIFT analysis.

3.3. The FTIR Spectra of the Dark Adsorption of CO2 on Photocatalysts

An airtight gas-cell FTIR accessory with 15 cm pathlength for transmission mode measurements was used to analyze the extent of interactions between adsorbates and photocatalysts. Figure 5 illustrates the FTIR spectra of Pd2Cu/P25 probed by CO2 from 5 to 500 Pa at ambient temperature. After the CO2 adsorption at 500 Pa was completed, the system was evacuated to vacuum again to remove the free CO2 and free carbonate species. The evacuated spectrum is symbolized as “Evac.” in Figure 5. This information provides strong evidence of the formation of chemical adsorption bonding between our sample and carbonate species. The peaks are assigned following the reference [43,44,45,46,47]. These spectra feature two major species, namely carbonates and bicarbonates, that were adsorbed on the catalysts by a few different kinds of connections. Starting with bicarbonates, the band at 1215 cm−1 is related to the (OH) vibrational signals of monodentate and bidentate bicarbonates (m-HCO3 and b-HCO3). The band at around 1625 cm−1 can be ascribed to the asymmetric vibrational signals (vas(CO3)) of both monodentate and bidentate bicarbonates (m-HCO3 and b-HCO3) [43,44,45,46]. However, the significant bands at 1425 cm−1, identified as the symmetric vibration signal (vs(CO3)) of the bidentate bicarbonates (b-HCO3), implies that the bicarbonate groups on the Pd2Cu/P25 were dominated by bidentate bicarbonates (b-HCO3).
The bands related to carbonate species were also investigated. The broad bands at 1675 cm−1 and 1325 cm−1 represented the asymmetric vibrational signals (vas(CO3)) and the symmetric vibration signal (vs(CO3)) of the adsorbed bridged carbonates (br-CO32−), respectively [43,44,45,46]. The significant bands at 1525 cm−1 and 1350 cm−1 exhibited the asymmetric vibrational signals (vas (CO3)) and the symmetric vibration signal (vs(CO3)) of the adsorbed bidentate carbonates (b-CO32−), respectively. Besides, a small shoulder at 1500 cm−1 (asymmetric vibrational signals (vas (CO3)) indicated that there was a trace amount of poly-dentate carbonate species (p-CO32−) observed by IR. After the sample reached equilibrium at 500 Pa of CO2, the system was then turned in to vacuum for 30 min. As shown in Figure 5, the remaining signals mostly belonged to carbonates. This result can provide further insights in DRIFT analysis when the reaction was applied.
Thus, the dark adsorption FTIR analysis probed by CO2 was carried out using Pd2Cu/P25. This analysis gave us a closer look at the micro-environment around the active sites and the adsorbent, which helped us rationalize the photocatalytic performance of each sample in the next section.

3.4. In-Situ DRIFT Investigation of Photothermal CO2 Reduction

The mechanisms of CO2 reduction and resultant intermediate products were monitored by in-situ DRIFT analysis. Figure 6a shows the time-dependent DRIFT absorbance spectra of Pd2Cu/P25. The pristine Pd2Cu/P25 shows peaks of around 3500–3000 cm−1 and 1320 cm−1 for the surface hydroxyl group (moisture). These surface hydroxyl groups were initially removed by thermal treatment at 150 °C for 20 m in a nitrogen atmosphere. After removing the moisture, the CO2 and gas-phase H2O were passed into the drift cell at 150 °C. The spectra were collected in time intervals (t) of 20, 40, and 60 min in darkness (absence of light) and in the presence of light. The peaks appearing at 2346 cm−1 and 3700–3500 cm−1 correspond to the ʋ3 and ʋ3 + ʋ1 stretching vibrations of adsorbed CO2, respectively. Additionally, the monodentate carbonate CO32− peak appeared at 1085 cm−1 for the metal carbonate (m-C–O–) [18], which confirmed the CO2 adsorption on Pd2Cu/P25 in dark.
This metal carbonate (m-C–O–) peak and the CO2 peaks (3700–3500 cm−1) were slowly decreased when irradiated with UV-Vis light, and at the same time, the C–H peaks are increased at 2962 cm−1 (Figure 6a). In addition, the *CO peak arises at 1640 cm−1 during the light irradiation, as shown in Figure 6b [19]. The obtained DRIFT spectra were derived to visualize the significant changes. Thereby, the obtained IR spectra were subtracted from the initial absorbance spectra. The T0, T20, T40, and T60 refer to the spectra taken in darkness at t = 20, 40, and 60 min respectively. Similarly, TL0, TL20, TL40, and TL60 refer to the spectra under UV-Vis light irradiation at t = 20, 40, and 60 min respectively. As shown in Figure 6c, the specific peaks at 2937, 2873 cm−1 (asymmetric and symmetric C-H stretching frequency), and 3250 cm−1 (–OH symmetric stretching) are increased with the reaction time [48,49,50]. The strong peaks at 1640 and 1473 cm−1 were also increased, which correspond to the *CO and CH2 (scissoring), respectively. The small peaks at 1266 and 1163 cm−1 refer to the formation of C–C (stretching frequency) and CO2δ−, respectively [19]. The peaks at 1556 and 1425 cm−1 correspond to the CO32− and CHCOH respectively. These two peaks are gradually decreased by further increasing the reaction time. From these observations, we have illustrated the mechanistic pathway of CO2 reduction given below.
Nanomaterials 12 00332 i001
In the above reaction steps, the CO2 gas-phase is adsorbed as a monodentate metal carbonate (m-C–O–) CO32−. On the catalytic (Cu) surface, the CO32− to *CO conversion occurs via π–π interaction by the excited electrons [46]. The local concentration of CO on the catalyst (Pd) undergoes the C–C coupling step, and then the OC–CO reacts with water and produces the intermediates of CHCOH. Furthermore, the addition of water with CHCOH produces the final product of ethanol [18]. Overall, DRIFT studies have demonstrated the reaction pathways wherein Cu was assumed as a CO2 adsorption site. It is very difficult to separate the active sites in DRIFT spectra where the CO2 is adsorbed. However, previous studies have demonstrated that CO2 adsorption is more favorable on the Cu site [25].
On other hand, Pd was engaged in the C–C coupling reaction, as well as the highly selective hydrogenation of CO2 to C2H5OH. The production of ethanol was confirmed by the mass spectra. The CO2 reduction on Pd2Cu/P25 was compared with P25. The corresponding DRIFT results are shown in Figure 6d. Hence, the Pd2Cu/P25 is a more selective and active catalyst that produces ethanol when compared to P25. Moreover, the photothermal hydrogenation of CO2 over the Pd2Cu/P25 catalyst was evaluated at various temperatures under UV-Vis light irradiation (Figure 6e). The DRIFT difference spectra reveal that the C2H5OH formation (identified from C-H peaks arising) increased with an increase in temperature. In contrast, the reaction was continued without light irradiation at 150 °C, exhibiting no C–H peak, and with no ethanol formation in dark (Figure 6f). Therefore, the light provides the driving force for CO2 reduction (0.611 V) and water oxidation (1.23 V). Overall, the reaction begins with the oxidation of water on P25 and then the electrons that evolved during water oxidation are utilized by Pd2Cu. The H2 molecules are generated from the P25 as a result of photocatalytic water oxidation. The Cu provides the active site for CO2 adsorption and Pd oxidizes the generated H2 molecules and involves the C–C bond formation. The key step of the CO2 reduction is the hydrogenation of *CO to *HCO, which is performed by the Pd active sites.

4. Conclusions

This work reports the photothermal reduction of CO2 and the oxidation of H2O over the Pd2Cu/P25 photocatalyst. The formation of ethanol was confirmed from GCMS and online mass spectroscopy. The Pd2Cu/P25 showed an efficient CO2 reduction photothermally at 150 °C with an ethanol production rate of 4.1 mmol g1 h1 under UV-Vis irradiation. Other products were not visible on GCMS spectra due to their low quantity. However, operando DRIFT absorption spectroscopy traces the intermediate products. The mechanism of CO2 reduction on the Pd2Cu/P25 catalyst was monitored by DRIFTS. The reaction begins with the oxidation of water on P25 and then the electrons involved in water oxidation were utilized by Pd2Cu. The Pd active sites utilize those electrons and hydrogenate the *CO to *HCO, which is a key step for the reduction of CO2 to ethanol. Moreover, the Pd2Cu/P25 catalyst demonstrated the highly active and selective hydrogenation of CO2 to C2H5OH.

Author Contributions

Conceptualization, M.E. and T.C.-K.Y.; methodology, M.E. and W.Y.; software, M.E. and W.Y.; validation, W.Y., S.V. and J.-N.C.; formal analysis, S.V. and J.-N.C.; investigation, S.V. and J.-N.C.; resources, T.C.-K.Y. and T.Y.; data curation, M.E., S.V. and J.-N.C.; writing—original draft preparation, M.E., W.Y. and S.V.; writing—review and editing, S.V., T.C.-K.Y. and T.Y.; visualization, M.E., W.Y. and S.V.; supervision, T.C.-K.Y. and T.Y.; project administration, T.C.-K.Y., funding acquisition, T.C.-K.Y. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge partial financial support from the Taiwan Ministry of Science and Technology with project number 110-2221-E-027-006-MY2.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors express gratitude to Precision Analysis and Material Research Center, National Taipei University of Technology, Taipei, Taiwan for providing instrumental facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yoon, Y.; Hall, A.S.; Surendranath, Y. Tuning of silver catalyst mesostructure promotes selective carbon dioxide conversion into fuels. Angew. Chem. Int. Ed. 2016, 55, 15282–15286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Rodemerck, U.; Holeňa, M.; Wagner, E.; Smejkal, Q.; Barkschat, A.; Baerns, M. Catalyst development for CO2 hydrogenation to fuels. ChemCatChem 2013, 5, 1948–1955. [Google Scholar] [CrossRef]
  3. Behrens, M. Promoting the Synthesis of Methanol: Understanding the requirements for an industrial catalyst for the conversion of CO2. Angew. Chem. Int. Ed. 2016, 55, 14906–14908. [Google Scholar] [CrossRef]
  4. Behrens, M. Heterogeneous catalysis of CO2 conversion to methanol on copper surfaces, Angew. Chem. Int. Ed. 2014, 53, 12022–12024. [Google Scholar] [CrossRef] [PubMed]
  5. Robatjazi, H.; Zhao, H.; Swearer, D.F.; Hogan, N.J.; Zhou, L.; Alabastri, A.; McClain, M.J.; Nordlander, P.; Halas, N.J. Plasmon-induced selective carbon dioxide conversion on earth-abundant aluminum–cuprous oxide antenna-reactor nanoparticles. Nat. Commun. 2017, 8, 27. [Google Scholar] [CrossRef]
  6. Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 4, 3703–3727. [Google Scholar] [CrossRef] [Green Version]
  7. Klankermayer, J.; Wesselbaum, S.; Beydoun, K.; Leitner, W. Selective catalytic synthesis using the combination of carbon dioxide and hydrogen: Catalytic chess at the interface of energy and chemistry. Angew. Chem. Int. Ed. 2016, 55, 7296–7343. [Google Scholar] [CrossRef] [PubMed]
  8. Lu, Q.; Rosen, J.; Zhou, Y.; Hutchings, G.S.; Kimmel, Y.C.; Chen, J.G.; Jiao, F. A selective and efficient electrocatalyst for carbon dioxide reduction. Nat. Commun. 2014, 5, 3242. [Google Scholar] [CrossRef]
  9. Kwak, J.H.; Kovarik, L.; Szanyi, J. Photocatalytic reduction of CO2 on TiO2 and other semiconductors, CO2 reduction on supported Ru/Al2O3 catalysts: Cluster size dependence of product selectivity. ACS Catal. 2013, 3, 2449–2455. [Google Scholar] [CrossRef]
  10. Porosoff, M.D.; Yang, X.; Boscoboinik, J.A.; Chen, J.G. Molybdenum carbide as alternative catalysts to precious metals for highly selective reduction of CO2 to CO. Angew. Chem. Int. Ed. 2014, 53, 6705–6709. [Google Scholar] [CrossRef]
  11. Martin, O.; Martín, A.J.; Mondelli, C.; Mitchell, S.; Segawa, T.F.; Hauert, R.; Drouilly, C.; Curulla-Ferré, D.; Pérez-Ramírez, J. Indium oxide as a superior catalyst for methanol synthesis by CO2 hydrogenation. Angew. Chem. Int. Ed. 2016, 55, 6261–6265. [Google Scholar] [CrossRef]
  12. Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjær, C.F.; Hummelshøj, J.S.; Dah, S.; Chorkendorff, I.; Nørskov, J.K. Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nat. Chem. 2014, 6, 320–324. [Google Scholar] [CrossRef] [PubMed]
  13. Graciani, J.; Mudiyanselage, K.; Xu, F.; Baber, A.E.; Evans, J.; Senanayake, S.D.; Stacchiola, D.J.; Liu, P.; Hrbek, J.; Sanz, J.F.; et al. Highly active copper-ceria and copper-ceria-titania catalysts for methanol synthesis from CO2. Science 2014, 345, 546–550. [Google Scholar] [CrossRef] [PubMed]
  14. Moret, S.; Dyson, P.J.; Laurenczy, G. Direct synthesis of formic acid from carbon dioxide by hydrogenation in acidic media. Nat. Commun. 2014, 5, 4017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Zhu, Y.; Zhang, S.; Ye, Y.; Zhang, X.; Wang, L.; Zhu, W.; Cheng, F.; Tao, F. Catalytic conversion of carbon dioxide to methane on Ruthenium-Cobalt bimetallic nanocatalysts and correlation between surface chemistry of catalysts under reaction conditions and catalytic performances. ACS Catal. 2012, 2, 2403–2408. [Google Scholar] [CrossRef]
  16. Mistry, H.; Varela, A.S.; Bonifacio, C.S.; Zegkinoglou, I.; Sinev, I.; Choi, Y.-W.; Kisslinger, K.; Stach, E.A.; Yang, J.C.; Strasser, P.; et al. Highly selective plasma-activated copper catalysts for carbon dioxide reduction to ethylene. Nat. Commun. 2016, 7, 12123. [Google Scholar] [CrossRef]
  17. Choi, Y.H.; Jang, Y.J.; Park, H.; Kimb, W.Y.; Lee, Y.H.; Choi, S.H.; Lee, J.S. Carbon dioxide Fischer–Tropsch synthesis: A new path to carbon-neutral fuels. Appl. Catal. B Environ. 2017, 202, 605–610. [Google Scholar] [CrossRef]
  18. Li, F.; Li, Y.C.; Wang, Z.; Li, J.; Nam, D.-H.; Lum, Y.; Luo, M.; Wang, X.; Ozden, A.; Hung, S.-F.; et al. Cooperative CO2-to-ethanol conversion via enriched intermediates at molecule–metal catalyst interfaces. Nat. Catal. 2020, 3, 75–82. [Google Scholar] [CrossRef]
  19. Bai, S.; Shao, Q.; Wang, P.; Dai, Q.; Wang, X.; Huang, X. Highly Active and Selective Hydrogenation of CO2 to Ethanol by Ordered Pd−Cu Nanoparticles. J. Am. Chem. Soc. 2017, 139, 6827–6830. [Google Scholar] [CrossRef]
  20. Qian, C.; Sun, W.; Hung, D.L.H.; Qiu, C.; Makarem, M.; Kumar, S.G.H.; Wan, L.; Ghoussoub, M.; Wood, T.E.; Xia, M.; et al. Catalytic CO2 reduction by palladium-decorated silicon–hydride nanosheets. Nat. Catal. 2019, 2, 46–54. [Google Scholar] [CrossRef]
  21. Fujitani, T.; Nakamura, I. Methanol Synthesis from CO and CO2 Hydrogenations over Supported Palladium Catalysts. Bull. Chem. Soc. Jpn. 2002, 75, 1393–1398. [Google Scholar] [CrossRef]
  22. Crudden, C.M.; Sateesh, M.; Lewis, R. Mercaptopropyl-modified mesoporous silica: Remarkable support for the preparation of a reusable, heterogeneous palladium catalyst for coupling reactions. J. Am. Chem. Soc. 2005, 127, 10045–10050. [Google Scholar] [CrossRef]
  23. Pang, S.H.; Schoenbaum, C.A.; Schwartz, D.K.; Medlin, J.W. Directing reaction pathways by catalyst active-site selection using self-assembled monolayers. Nat. Commun. 2013, 4, 2448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Lee, G.-J.; Anandan, S.; Masten, S.J.; Wu, J.J. Photocatalytic hydrogen evolution from water splitting using Cu doped ZnS microspheres under visible light irradiation. Renew. Energy 2016, 89, 2619. [Google Scholar] [CrossRef]
  25. Imtiaz, Q.; Yüzbasi, N.S.; Abdala, P.M.; Kierzkowska, A.M.; Beek, W.V.; Brodaa, M.; Müller, C.R. Development of MgAl2O4-stabilized, Cu-doped, Fe2O3-based oxygen carriers for thermochemical water-splitting. J. Mater. Chem. A. 2016, 4, 113–123. [Google Scholar] [CrossRef]
  26. Hunge, Y.M.; Yadav, A.A.; Mathe, V.L. Photocatalytic hydrogen production using TiO2 nanogranules prepared by hydrothermal route. Chem. Phys. Lett. 2019, 731, 136582. [Google Scholar] [CrossRef]
  27. Gosavi, S.; Tabei, R.; Roy, N.; Latthe, S.S.; Hunge, Y.M.; Suzuki, N.; Kondo, T.; Yuasa, M.; Teshima, K.; Fujishima, A.; et al. Low Temperature Deposition of TiO2 Thin Films through Atmospheric Pressure Plasma Jet Processing. Catalysts 2021, 11, 91. [Google Scholar] [CrossRef]
  28. Chen, Y.-W.; Hsu, Y.-H. Effects of reaction temperature on the photocatalytic activity of TiO2 with Pd and Cu cocatalysts. Catalysts 2021, 11, 966. [Google Scholar] [CrossRef]
  29. Lohumi, S.; Lee, S.; Lee, H.; Cho, B.K. A review of vibrational spectroscopic techniques for the detection of food authenticity and adulteration. Trends Food Sci. Technol. 2015, 46, 85–98. [Google Scholar] [CrossRef]
  30. Manju, G.; Archana, J.; Verma, K.K. Dispersive liquid–liquid microextraction and diffuse reflectance-Fourier transform infrared spectroscopy for iodate determination in food grade salt and food samples. Food Chem. 2022, 368, 130810. [Google Scholar]
  31. Elavarasan, M.; Uma, K.; Yang, T.C.-K. Photocatalytic oxidation of ethanol using ultrasonic modified TiO2; an in-situ diffuse reflectance infrared spectroscopy study. Results Phys. 2019, 13, 1022372. [Google Scholar] [CrossRef]
  32. Yang, T.C.-K.; Wang, S.-F.; Tsai, S.H.-Y.; Lin, S.-Y. Intrinsic photocatalytic oxidation of the dye adsorbed on TiO2 photocatalysts by diffuse reflectance infrared Fourier transform spectroscopy. Appl. Catal. B Environ. 2001, 30, 293–301. [Google Scholar] [CrossRef]
  33. Liu, J.; Winwarid, P.; Yang, T.C.-K.; Chuang, S.S.-C. In situ infrared study of photo-generated electrons and adsorbed species on nitrogen-doped TiO2 in dye-sensitized solar cells. Phys. Chem. Chem. Phys. 2018, 20, 19572–19580. [Google Scholar] [CrossRef]
  34. Xu, F.; Meng, K.; Cheng, B.; Yu, J.; Ho, W. Enhanced Photocatalytic Activity and Selectivity for CO2 Reduction over a TiO2 Nanofiber Mat Using Ag and MgO as Bi-Cocatalyst. ChemCatChem. 2019, 11, 465–472. [Google Scholar] [CrossRef]
  35. Yang, Q.; Le, X.; Li, Y.; Ian, T.M.; Guo, F.; Tao, M.; Yang, R.; Qi, L.; Lin, Z.; Gu, S.; et al. BCC-Phased PdCu Alloy as a Highly Active Electrocatalyst for Hydrogen Oxidation in Alkaline Electrolytes. J. Am. Chem. Soc. 2018, 140, 16580–16588. [Google Scholar] [CrossRef]
  36. Gu, Z.; Xiong, Z.; Ren, F.; Li, S.; Xu, H.; Yan, B.; Du, Y. Flower-like PdCu catalyst with high electrocatalytic properties for ethylene glycol oxidation. J. Taiwan Inst. Chem. E 2018, 83, 32–39. [Google Scholar] [CrossRef]
  37. Tiscione, N.B.; Alford, I.; Yeatman, D.T.; Shan, X. Ethanol Analysis by Headspace Gas Chromatography with Simultaneous Flame-Ionization and Mass Spectrometry Detection. J. Anal. Toxicol. 2011, 35, 501–511. [Google Scholar] [CrossRef] [Green Version]
  38. He, Z.; Qian, Q.; Ma, J.; Meng, Q.; Zhou, H.; Song, J.; Liu, Z.; Han, B. Water-enhanced synthesis of higher alcohols from CO2 hydrogenation over a Pt/Co3O4 catalyst under Milder conditions, Angew. Chem. Int. Ed. 2016, 55, 737–741. [Google Scholar] [CrossRef]
  39. Pan, X.; Fan, Z.; Chen, W.; Ding, Y.; Luo, H.; Bao, X. Enhanced ethanol production inside carbon-nanotube reactors containing catalytic particles. Nat. Mater. 2007, 6, 507–511. [Google Scholar] [CrossRef] [PubMed]
  40. Yang, N.; Medford, A.J.; Liu, X.; Studt, F.; Bligaard, T.; Bent, S.F.; Nørskov, J.K. Intrinsic selectivity and structure sensitivity of rhodium catalysts for C2+ oxygenate production. J. Am. Chem. Soc. 2016, 138, 3705–3714. [Google Scholar] [CrossRef]
  41. Wang, D.; Bi, Q.; Yin, G.; Zhao, W.; Huang, F.; Xie, X.; Jiang, M. Direct synthesis of ethanol via CO2 hydrogenation using supported gold catalysts. Chem. Commun. 2016, 52, 14226–14229. [Google Scholar] [CrossRef] [PubMed]
  42. Jiang, X.; Koizumi, N.; Guo, X.; Song, C. Bimetallic Pd–Cu catalysts for selective CO2 hydrogenation to methanol. Appl. Catal. B Environ. 2015, 170, 173–185. [Google Scholar] [CrossRef]
  43. Baltrusaitis, J.; Jensen, J.H.; Grassian, V.H. FTIR spectroscopy combined with isotope labeling and quantum chemical calculations to investigate adsorbed bicarbonate formation following reaction of carbon dioxide with surface hydroxyl groups on Fe2O3 and Al2O3. J. Phys. Chem. B 2006, 110, 12005–12016. [Google Scholar] [CrossRef]
  44. Turek, A.M.; Wachs, I.E.; DeCanio, E. Acidic properties of alumina-supported metal oxide catalysts: An infrared spectroscopy study. J. Phys. Chem. B 1992, 96, 5000–5007. [Google Scholar] [CrossRef]
  45. Collins, S.E.; Baltanás, M.A.; Bonivardi, A.L. Infrared spectroscopic study of the carbon dioxide adsorption on the surface of Ga2O3 polymorphs. J. Phys. Chem. B 2006, 110, 5498–5507. [Google Scholar] [CrossRef]
  46. Collins, S.E.; Baltanas, M.A.; Bonivardi, A.L. An infrared study of the intermediates of methanol synthesis from carbon dioxide over Pd/β-Ga2O3. J. Catal. 2004, 226, 410–421. [Google Scholar] [CrossRef]
  47. Köck, E.M.; Kogler, M.; Bielz, T.; Klötzer, B.; Penner, S. In situ FT-IR spectroscopic study of CO2 and CO adsorption on Y2O3, ZrO2, and yttria-stabilized ZrO2. J. Phys. Chem. C 2013, 117, 17666–17673. [Google Scholar]
  48. Yu, Z.; Chuang, S.S.C. In situ IR study of adsorbed species and photo-generated electrons during photocatalytic oxidation of ethanol on TiO2. J. Catal. 2007, 246, 118–126. [Google Scholar] [CrossRef]
  49. Elavarasan, M.; Uma, K.; Yang, T.C.-K. Nanocubes phase adaptation of In2O3/TiO2 heterojunction photocatalysts for the dye degradation and tracing of adsorbed species during photo-oxidation of ethanol. J. Taiwan Inst. Chem. E 2021, 120, 1–9. [Google Scholar] [CrossRef]
  50. Guzman, F.; Chuang, S.S.C. Tracing the Reaction Steps Involving Oxygen and IR Observable Species in Ethanol Photocatalytic Oxidation on TiO2. J. Am. Chem. Soc. 2010, 132, 1502–1503. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD pattern, (b) Raman spectra (c) UV−Vis (DRS), inset is showing Tauc plots of P25 and Pd2Cu/P25, (d) FESEM Backscattered−Electron (BSE) Image of Pd2Cu/P25.
Figure 1. (a) XRD pattern, (b) Raman spectra (c) UV−Vis (DRS), inset is showing Tauc plots of P25 and Pd2Cu/P25, (d) FESEM Backscattered−Electron (BSE) Image of Pd2Cu/P25.
Nanomaterials 12 00332 g001
Figure 2. (ac) low and high magnified TEM images, (d) lattice fringes, (e) SAED patterns, (f) EDX spectra, (gj) mapping of Pd2Cu/P25 (Elements: Pd (g); Cu (h); Ti (i); O (j)).
Figure 2. (ac) low and high magnified TEM images, (d) lattice fringes, (e) SAED patterns, (f) EDX spectra, (gj) mapping of Pd2Cu/P25 (Elements: Pd (g); Cu (h); Ti (i); O (j)).
Nanomaterials 12 00332 g002
Figure 3. (a) Pd 3d, (b) Cu 2p, (c) Ti 2p, and (d) O 1s XPS spectra of Pd2Cu/P25.
Figure 3. (a) Pd 3d, (b) Cu 2p, (c) Ti 2p, and (d) O 1s XPS spectra of Pd2Cu/P25.
Nanomaterials 12 00332 g003
Figure 4. (a) Schematic diagram of photothermal CO2 reduction using photoreactor, (b) GCMS spectrum of ethanol after 1h UV-Vis light irradiation at 80 °C. (c) The overall process of CO2 reduction and H2O oxidation on Pd2Cu/P25 (CB: conduction band, VB: valence band, h+: holes, and e: electrons).
Figure 4. (a) Schematic diagram of photothermal CO2 reduction using photoreactor, (b) GCMS spectrum of ethanol after 1h UV-Vis light irradiation at 80 °C. (c) The overall process of CO2 reduction and H2O oxidation on Pd2Cu/P25 (CB: conduction band, VB: valence band, h+: holes, and e: electrons).
Nanomaterials 12 00332 g004
Figure 5. (a) The FTIR spectra of the dark adsorption on the Pd2Cu/P25 photocatalysts and (b) peak assignments at various pressure gradients from 5 Pa to 500 Pa of CO2 vapor and finally in vacuum condition.
Figure 5. (a) The FTIR spectra of the dark adsorption on the Pd2Cu/P25 photocatalysts and (b) peak assignments at various pressure gradients from 5 Pa to 500 Pa of CO2 vapor and finally in vacuum condition.
Nanomaterials 12 00332 g005
Figure 6. (a,b) Time-dependent DRIFT spectra of Pd2Cu/P25 and (c) IR difference spectra of CO2 reduction on Pd2Cu/P25. (d) IR difference spectra of P25 and Pd2Cu/P25 after 1 h light irradiation. DRIFT spectra CO2 reduction on Pd2Cu/P25 at various temperatures (e) with and (f) without light irradiation.
Figure 6. (a,b) Time-dependent DRIFT spectra of Pd2Cu/P25 and (c) IR difference spectra of CO2 reduction on Pd2Cu/P25. (d) IR difference spectra of P25 and Pd2Cu/P25 after 1 h light irradiation. DRIFT spectra CO2 reduction on Pd2Cu/P25 at various temperatures (e) with and (f) without light irradiation.
Nanomaterials 12 00332 g006
Table 1. ICP-AES results of Pd2Cu/P25 photocatalyst.
Table 1. ICP-AES results of Pd2Cu/P25 photocatalyst.
ICP AnalysisPdCuTi
mg L−10.04160.01290.232
Molecular Weight106.4263.5447.867
mMolL−13.9 × 10−42.0 × 10−44.8 × 10−3
Molar ratios~2124
Table 2. The comparison of ethanol yield on Pd2Cu/P25 with various catalysts.
Table 2. The comparison of ethanol yield on Pd2Cu/P25 with various catalysts.
CatalystsGas (COx)Temperature (°C)C2H5OH Yields
(mmol g−1 h−1)
Reference
1% Pt/Co3O4CO22000.07[38]
1.2% RhMnLiFeCO3203.5[39]
Rh/SiO2CO2500.37[40]
Au/TiO2CO22002.82[41]
8.7% Pd-10% Cu/SiO2CO22501.12[42]
Pd2Cu/P25CO21504.1This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Elavarasan, M.; Yang, W.; Velmurugan, S.; Chen, J.-N.; Yang, T.C.-K.; Yokoi, T. Highly Efficient Photothermal Reduction of CO2 on Pd2Cu Dispersed TiO2 Photocatalyst and Operando DRIFT Spectroscopic Analysis of Reactive Intermediates. Nanomaterials 2022, 12, 332. https://doi.org/10.3390/nano12030332

AMA Style

Elavarasan M, Yang W, Velmurugan S, Chen J-N, Yang TC-K, Yokoi T. Highly Efficient Photothermal Reduction of CO2 on Pd2Cu Dispersed TiO2 Photocatalyst and Operando DRIFT Spectroscopic Analysis of Reactive Intermediates. Nanomaterials. 2022; 12(3):332. https://doi.org/10.3390/nano12030332

Chicago/Turabian Style

Elavarasan, Munirathinam, Willie Yang, Sethupathi Velmurugan, Jyy-Ning Chen, Thomas C.-K. Yang, and Toshiyuki Yokoi. 2022. "Highly Efficient Photothermal Reduction of CO2 on Pd2Cu Dispersed TiO2 Photocatalyst and Operando DRIFT Spectroscopic Analysis of Reactive Intermediates" Nanomaterials 12, no. 3: 332. https://doi.org/10.3390/nano12030332

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