Napat Lertthanaphola,
Natthanicha Prawiseta,
Pornpinun Soontornapaluka,
Nutkamol Kitjanukita,
Wannisa Neamsunga,
Natpichan Pienutsaa,
Kittapas Chusria,
Thirawit Sornsuchata,
Prowpatchara Chantharaa,
Poomiwat Phadungbuta,
Panpailin Seeharajb,
Pattaraporn Kim-Lohsoontornc and
Sira Srinives
*a
aNanocomposite Engineering Laboratory (NanoCEN), Department of Chemical Engineering, Faculty of Engineering, Mahidol University, Nakorn Pathom, 73170, Thailand. E-mail: Sira.sri@mahidol.edu
bAdvanced Materials Research Unit, Department of Chemistry, Faculty of Science, King Mongkut's Institute of Technology Ladkrabang, Bangkok, 10520, Thailand
cCenter of Excellence on Catalysis and Catalytic Reaction Engineering, Department of Chemical Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok, 10330, Thailand
First published on 26th August 2022
Photoreduction of CO2 to a high-value product is an interesting approach that not only captures CO2 but also converts it into other products that can be sold or used in industry. The mechanism for the CO2 conversion relies strongly on photo-generated electrons that further couple with CO2 and form active radicals for the reaction. In this research, we synthesized a heterostructure of copper-doped sodium dititanate nanosheets and graphene oxide (CTGN) following a one-step hydrothermal process with assistance from a sodium hydroxide soft template. The role of the template here is to facilitate the formation of the nanosheets, creating the nanosheet–graphene 2D–2D heterostructure. The heterostructure yields excellent charge mobility and a low charge recombination rate, while the nanosheet–graphene interfaces house active radicals and stabilize intermediates. The CTGN exhibits an outstanding photoactivity in the photoreduction of CO2, producing liquid fuels, including acetone, methanol, ethanol and i-propanol.
Different techniques were introduced and demonstrated for CO2 conversion, including CO2 fixation and conversion by microalgae, CO2 hydrogenation by metal oxide catalyst,8 and CO2 splitting using a metal oxide electrocatalyst.9 Photoreduction of CO2 to liquid fuels is an attractive alternative that relies on photocatalysts such as zinc oxide (ZnO),3,10 cadmium sulfide (CdS),2 and titanium dioxide (TiO2).5,6,10–19 TiO2 is a popular photocatalyst that has been used in the decomposition of organic pollutants in wastewater. It has also featured heavily as a potent photocatalyst for CO2 conversion. An issue regarding TiO2 concerns the wide bandgap energy, which limits the number of photo-generated electrons resulting in a fast pairing rate of electrons and holes. TiO2 also responds only to UV light, preventing it from utilizing the full intensity of natural sunlight.
A photocatalytic heterostructure between two 2-dimensional nanostructures, defined as a 2D–2D heterostructure, can be an ideal photocatalytic platform that provides excellent charge mobility and charge separation.13,15,20–22 The heterostructure contains interfaces between two semiconductors with unequal bandgap values, which induces a local electric field that directs the flow of charge carriers. The team of J. Sun22 synthesized a TiO2 nanosheets/graphene 2D–2D heterostructure by introducing hydrofluoric acid (HF) to a titanate–graphene oxide (GO) mixture in a solvothermal process. The TiO2 nanosheets grew on and were in good contact with the GO. Zhao et al.13 and Keerthana et al.15 demonstrated the use of alkali solutions such as sodium hydroxide (NaOH) and potassium hydroxide (KOH) as a soft template in the formation of sodium dititanate (Na2Ti2O5) nanosheets. The mechanism involved hydrolysis of a titanate precursor, followed by a formation of the dititanate interlayers. The layers were further intercalated by the alkali ions, stabilized, and became nanosheets.4,13,15,23
Graphene is a superior choice for one-half of the 2D–2D heterostructure since it has a good charge transfer ability, chemical stability, and outstanding light absorption properties.2,24 It can be synthesized following a chemical exfoliation approach, yielding GO, which is a few layers of graphene sheet with carbon–hydrogen–oxygen functional groups. The functional groups serve as defects in the nanostructure and provide sites for the precipitation and immobilization of metal/metal oxide nanostructures. The sodium dititanate nanosheets can be synthesized and immobilized on GO via a hydrothermal process in an alkali solution. The 2D–2D photocatalytic heterostructure can be of great use to the photoreduction of CO2 to liquid fuels.
In this work, we synthesized the 2D–2D photocatalytic heterostructure of copper-doped sodium dititanate nanosheets/GO (CTGN). The heterostructure was synthesized using a one-step hydrothermal process with the addition of a NaOH soft template. Some chemical, physical and crystallographic properties of the solid samples were studied using analytical instruments, including X-ray diffraction (XRD), Raman microscope, Fourier-transform infrared spectroscopy (FTIR), UV-Visible spectroscopy (UV-Vis), high-resolution transmission electron microscope (HR-TEM), field effect scanning electron microscope (FE-SEM) and electron dispersive spectroscopy (EDS). The photocatalytic property was characterized using photoluminescence spectroscopy (PL). Liquid samples from the photoreduction of CO2 were analyzed using gas chromatography (GC) to obtain composition of the liquid fuels.
(αhν)1/n = α0(hν − Eg), | (1) |
The Ti sample (Fig. 1(a) and (b)) appears to be in the form of nanoparticles with a diameter of 5.9 ± 1.5 nm while the light diffraction pattern (Fig. 1(b), inset) reveals a combination of single-crystalline and polycrystalline structures. The HR-TEM lattice fringe has a space of 0.34 nm, which correlates with the lattice pattern and the (101) planar of anatase TiO2.10 The TiN (Fig. 1(c) and (d)) has the shape of nanosheets of 77.9 ± 33.0 × 90.2 ± 37.9 nm2 in size.31 The diffraction pattern (Fig. 1(d), inset) exhibits mixed crystallography of polycrystalline and amorphous structures. Effects of the NaOH soft template on the formation of titanium nanostructure were well in agreement with the report from Zhao and his team.13
![]() | ||
Fig. 1 The HR-TEM and zoom-in HR-TEM images of Ti (a and b) and TiN (c and d): the light diffraction patterns of Ti (b, inset) and TiN (d, inset). |
The HR-TEM image of CT (Fig. 2(a)) shows nanoparticles with an average size of 7.2 ± 3.7 nm. TG has 5.8 ± 1.1 nm nanoparticles immobilized on GO (Fig. 2(b)). CTG shows nanoparticles with an average size of 5.8 ± 1.1 nm (Fig. 2(c)), decorated on GO. It is clear that with no assistance from the NaOH, TiO2 takes the form of nanoparticles with diameter sizes ranging from 3 to 11 nm.7,24,32 CTN (Fig. 2(d)) appears as a combination of nanosheets and nanoparticles, in which the nanoparticle has an average size of 6.4 ± 1.8 nm and nanosheets provide an average size of 51.7 ± 18.7 × 60.1 ± 26.5 nm2. TGN (Fig. 2(e)) is observed to be nanoparticles and nanosheets with wrinkles, in which the nanoparticles are 5.7 ± 1.8 nm in size and nanosheets are 35.5 ± 26.0 × 83.7 ± 34.1 nm2. CTGN (Fig. 2(f)) shows nanoparticles and nanosheets with sizes of 6.4 ± 1.4 nm and 33.45 ± 10.1 × 86.6 ± 35 nm2. The atomic composition of the CTGN was analyzed using electron dispersive X-ray spectroscopy (EDS) attached to the FE-SEM (Fig. S2 and S3†). The components of copper (Cu), sodium (Na), oxygen (O), carbon (C) and titanium (Ti) were all identified.
Crystal structure and phase composition of solid samples were analyzed using XRD. The XRD spectra of Ti, CT, TG and CTG (Fig. 3(a)) reveal diffraction peaks at 25.3°, 38.0°, and 48.0° (2θ), which can be interpreted as (101), (004), and (200) lattice planes of the anatase TiO2.10,18,19,23,31 The minor peaks at 54.4°, 63.1°, 69.4°, and 75.5° can also be ascribed to anatase TiO2.7,18 A small peak at 30.6° corresponds to a weak signal of brookite TiO2.10 Fig. 3(b) shows XRD spectra of TiN, CTN, TGN and CTGN with the major diffraction peak at 9.04°. The peak was indexed for the (200) plane of the dititanates.10,23 Other minor peaks located at 47.8° and 62.5° can be identified as anatase TiO2.7,10,23 XRD spectra from TiN, CTN, TGN and CTGN can be related to those of the sodium dititanate15 (Na2Ti2O5) and anatase TiO2. It is worth mentioning that no copper or copper oxide peaks were located.
![]() | ||
Fig. 3 (a) XRD spectra of Ti, CT, TG, and CTG; (b) XRD spectra of TiN, CTN, TGN, and CTGN; (c) Raman spectra of Ti, CT, TG, and CTG; (d) Raman spectra of TiN, CTN, TGN, and CTGN. |
The XRD results were cross-analyzed using Raman spectroscopy. For Ti and CT (Fig. 3(c)), the spectra reveal four characteristic Raman active modes for anatase TiO2 with Eg (144.6 and 636.5 cm−1), B1g (395.6 cm−1) and A1g (515.4 cm−1).18,33,34 For TG and CTG, the spectra peaks at 395.6, 515.4, and 636.5 cm−1 correspond to the B1g, A1g, and Eg of the anatase TiO2 Raman active modes. Relative intensities of the D and G band (ID/IG) for TG and CTG are determined to be 1.00 and 0.99. For TiN and CTN (Fig. 3(d)), the spectra display different modes of crystal structures in which the band signals from O–Ti–O (115.8 cm−1), Na–O–Ti (274.2 cm−1) and Ti–O (429.6, 702.4, and 893.9 cm−1) are indicated.20 Small bands observed at 377.1, 564.1, and 642.5 cm−1 are ascribed to the Eg, B1g and A1g modes of the anatase TiO2. Raman active bands are in good agreement with results from the XRD analysis, revealing crystallographic structures of Ti and CTN to be a combination of dititanate (Na2Ti2O5) and anatase TiO2. For the TGN and CTGN (Fig. 3(d)), the presence of dititanate structures is noticed as the bands for O–Ti–O (112.7 cm−1), Na–O–Ti (274.2 cm−1) and Ti–O (432.7, 702.4, and 893.9 cm−1) are verified. The weak Raman bands for anatase TiO2 are spotted at 377.1, 567.1 and 642.5, which correlate to the Eg, B1g and A1g of the anatase. The ID/IG for TGN and CTGN are determined to be 0.99 and 0.99, respectively, revealing an equivalent degree of the disordered and graphitic carbon.35,36
The chemical composition of the CTG and CTGN was analyzed using XPS (Fig. 4 and 5). A survey scan of CTG (Fig. 4(a)) displays elemental peaks at binding energies of 284.0, 458.0, 529.0, and 931.1 eV, which are ascribed to C 1s, Ti 2p, O 1s, and Cu 2p. The minor peaks at 36.2, 564.1 and 1073.1 eV are interpreted as Ti 3p, Ti 2s and Ti LMM. A narrow scan on the CTG Ti 2p (Fig. 4(b)) confirms the presence of Ti 2p3/2 and Ti 2p1/2 at the binding energies of 458.0 and 463.9 eV, which can be assigned to Ti4+. Fig. 4(c) displays a narrow scan of O 1s of the CTG sample, in which the peaks at 529.0, 529.7, and 531.1 eV can be assigned to Ti–O–Ti, Ti–O–Ti, and Ti–OH/H–O–C. A narrow scan of C 1s (Fig. S4(a)†) indicates the presence of O–CO (288.7 eV), C–O–C (287.4 eV), C
O (285.4 eV), and C–C/C
C/C–H (284.5 eV). The C–Ti bond (283.4 eV) indicates physical/chemical interactions between TiO2 and GO. Oxidative states of the Cu component are analyzed in a narrow scan of Cu 2p (Fig. 4(d)), in which major peaks of Cu 2p1/2 (952.8 eV) and Cu 2p3/2 (932.4 eV) are indexed. The Cu 2p peaks verify the presence of the Cu2+, in which the satellite peaks at 957.6 and 939.2 eV confirm the presence of the Cu2+. Fig. 5 exhibits the XPS spectra of the CTGN. A wide scan of the CTGN (Fig. 5(a)) presents characteristic peaks at binding energies of 284.8, 456.8, 529.8, and 931.7 eV, which are interpreted as C 1s, Ti 2p, O 1s, and Cu 2p components. The elemental peaks of Na 2s (61.9 eV), Na KLL (494.9 eV)37 and Na 1s (1074.1 eV)20 are also observed, indicating the presence of Na and the formation of dititanate (Na2Ti2O5). The narrow scan of CTGN (Fig. 5(b)) shows the binding energies of 456.8 (Ti 2p3/2) and 462.7 (Ti 2p1/2) eV, which reveal the oxidative state of +4 for the Ti (Ti4+).20 A narrow scan on O 1s of the CTGN (Fig. 5(c)) displays peaks at the binding energies of 528.1, 529.1, 530.5, 531.8, and 533.1 eV, which correspond to Ti–O–Ti, Ti–O–Ti, Ti–O–Na/Ti–O–Ti,20 Ti–OH/H–O–C, and C
O. Carbon components of CTGN are observed in a narrow scan of C 1s (Fig. S4(b)†). Binding energy peaks of 288.2, 286.5, 284.9 and 283.4 eV are observed and can be ascribed to O–C
O, C–O–C, C–C/C
C/C–H and C–Ti components. The C–Ti interaction indicates good adhesion between the dititanate nanosheets and GO, which is the key to the formation of the 2D–2D heterostructure. Fig. 5(d) shows the Cu 2p spectrum of the CTGN, which includes Cu 2p1/2 (958.0 eV), Cu 2p3/2 (931.3 eV) and the Cu2+ satellite. The peaks can be analyzed as the Cu2+ oxidative state for the Cu on the CTGN.
![]() | ||
Fig. 4 A survey scan on CTG (a); a narrow scan on CTG Ti 2p (b); a narrow scan on CTG O 1s (c); and a narrow scan on CTG Cu 2p (d). |
![]() | ||
Fig. 5 A survey scan on CTGN (a); a narrow scan on CTGN Ti 2p (b); a narrow scan on CTGN O 1s (c); a narrow scan on CTGN Cu 2p (d). |
Optical properties and energy bandgaps were characterized using UV-Vis spectroscopy (Fig. 6). The bandgap values were determined from UV-Vis spectra (Fig. 6, left) following Tauc correlation (eqn (1)) (Fig. 6, right). The Ti sample (Fig. 6(a), left) exhibits light absorption in the UV region (250–350 nm) and poor absorption in the visible range (350–700 nm). This confirms the disadvantage of normal TiO2 which can only absorb and be illuminated by UV light.7,10,15 The energy bandgap of Ti is determined to be 3.18 eV (Fig. 6(a), right), which agrees well with the reported value of 3.2 eV.6,19 The CT (Fig. 6(b), left) shows an enhanced light absorption ability since Cu reduces the electron–hole pairing rate and provides a transition state for photoelectrons to rest on. The bandgap for the CT sample is calculated as 2.91 eV, which is slightly lower than that of the Ti. The TG sample (Fig. 6(c), left) exhibits better light absorption ability compared to that of the Ti, absorbing UV light and part of the visible light, with the bandgap value of 2.53 eV. CTG (Fig. 6(d), left) shows light absorbance that reveals an energy bandgap of 2.28 eV as a combined effect of Cu doping and compositing of TiO2 with GO. For the soft template-induced samples, including TiN, CTN, TGN and CTGN, the UV-Vis spectra were also obtained (Fig. 6(e)–(h)). The TiN (Fig. 6(e)) presents a light absorption ability similar to that of the Ti, in which the TiN absorbs light effectively in the UV region but badly in the visible region. The bandgap is 3.14 eV, which is on the same scale as the reported value for the dititanate.15 The CTN (Fig. 6(f)) absorbs UV light and part of the visible light, showing an improvement in light absorption due to Cu doping. The energy bandgap of 3.09 eV is determined. The TGN (Fig. 6(g)) is the TiN immobilized on the GO sheet. It shows light absorption in both UV and visible regions with the bandgap value of 3.08 eV. The CTGN (Fig. 6(h)) reveals an outstanding light absorption ability from the boosting of Cu doping and GO support. Among the soft template-induced samples, the CTGN expresses the best optical properties in absorbing UV and visible light with the energy bandgap of 3.07 eV.
![]() | ||
Fig. 6 UV-Vis spectra (left) and Tauc plots (right) of Ti (a), CT (b), TG (c), CTG (d), and TiN (e), CTN (f), TGN (g) and CTGN (h). |
The chemical functionality of the samples was analyzed using FTIR (Fig. 7). The IR transmittance of the Ti, CT, TG and CTG (Fig. 7, left) peaks at 1416, 1633, and 3402 cm−1, which correspond to the carboxyl (COOH), adsorbed water on Ti (Ti–OH),38 and hydroxyl (C–OH) groups.15,27,30 A broad peak from 500 to 800 cm−1 can be ascribed to the Ti–O–Ti bond.15 For the TiN, CTN, TGN and CTGN (Fig. 7, right), an IR transmittance peak at 3421 cm−1 indicates the presence of the hydroxyl group (–OH), while the peak at 1455 cm−1 correlates to stretching of –COO− and Na+.39 A broad band between 500 to 800 cm−1 is related to a combined signal from Ti–O–Ti and Ti–O.15
CO2 conversion (%) = (CO2 solubility − total carbon consumption)/(CO2 solubility) × 100 | (2) |
The conversions are 99.6, 99.6, 99.2, 99.2, 99.7, 99.6, 98.0 and 97.2% for Ti, CT, TG, CTG, TiN, TGN and CTGN, presenting superior photoactivity of the CTGN over other photocatalysts.
To explain the photoactivity, selected samples, including Ti, CTG, TiN, TGN and CTGN were analyzed for their photocurrents. The photocatalyst-coated glassy carbon working electrode was employed as a working electrode (WE)16,26 while held at a constant potential of 0 V (vs. Ag/AgCl RE). Photocurrents were monitored while the WE was illuminated by the UV lamp for 50 s and unlit for another 50 s to complete a cycle (Fig. 8(b)). The samples respond quickly to the incident light as the photocurrent rises and drops sharply during the illumination and the darkness. The photocurrent reveals the ability of photocatalysts to generate and transfer photoelectrons to active radicals in the medium. The intensity of the photocurrent relies on two parameters: illumination and thermal diffusion.40 The illumination-induced photocurrent occurs quickly in a millisecond while the thermal-induced current emits slowly within seconds. The average photocurrents for the Ti, CTG, TiN, TGN and CTGN are monitored as 1.80 × 10−1 ± 1.45 × 10−2, 6.22 × 10−1 ± 0.89 × 10−2, 2.94 × 10−1 ± 3.51 × 10−2, 4.28 × 10−1 ± 4.68 × 10−2 and 19.67 × 10−1 ± 5.21 × 10−2 μA cm−2. The samples took 10 to 40 seconds to reach their saturated photocurrents, indicating a combination of photo-illumination and thermal induction effects on the photocurrents. The CTGN exhibits great photoelectroactivity, providing a photocurrent 3.2 times higher than that of the CTG and comparable to that of the reported 1D nanostructure.26
The Ti, CTG, TiN, TGN and CTGN were analyzed further for their photoluminescence properties using photoluminescence spectroscopy (PL). The samples were excited at a wavelength of 345 nm while photons emitted from the samples in a relaxation state were collected in the 300 to 700 nm range (Fig. 8(c)). PL spectra from Ti, CTG, TiN, TGN and CTGN samples show a characteristic peak at 475 nm, indicating the main luminescence wavelength for the radiated photons. The photon emissions, in this case, correlate to the recombination effect between electrons and holes in which the higher the PL peak intensity, the higher the recombination rate.19 To quantify the differences in PL spectra, the quenching factor was determined17 by dividing the integrated area under the PL spectra over the 350 to 650 nm range of a photocatalyst with that of the TiN. The quenching factors for Ti, CTG, TiN, TGN and CTGN were determined to be 0.92, 0.53, 1.00, 0.38 and 0.42, respectively. Ti and TiN provide nearly the same value of quenching factors, indicating that the TiO2 (Ti) and dititanate (TiN) on their own have equivalent photoactivity. The quenching factor for CTG is significantly lower than that of Ti as a result of the copper dopant and GO support. Both the TGN and CTGN exhibit excellent characteristics for photoluminescence and photon radiation,17,22 showing low quenching factors. The PL supports our CO2 photoreduction results that the NaOH soft template-induced dititanate/graphene, TGN and CTGN, yield outstanding photoactivity.
CO2 + 6H+ + 6e− → CH3OH + H2O | (3) |
2CO2 + 12H+ + 12e− → C2H5OH + 3H2O | (4) |
3CO2 + 18H+ + 18e− → C3H7OH + 5H2O | (5) |
3CO2 + 16H+ + 16e− → (CH3)2CO + 5H2O | (6) |
The key to an excellent photocatalyst is to have photoelectron and proton generating sites close to one another and to inhibit electron–hole recombination (Fig. 9). The 2D–2D heterostructure tends to have outstanding properties in charge transfer and charge separation. They allow photoelectrons to travel along the shortest path through the structures and combine with other active radicals in the solution phase. The photo-generated holes can interact with and be stabilized by charges and radicals in the dititanate/GO interfaces. The interfaces can also serve as n–p heterojunctions that present resting sites for adsorption of CO2 and active radicals, providing sufficient CO2 feed and stabilizing intermediates.
The excellent photocatalytic performance of the CTGN was realized by implementing the three approaches: synthesis of the 2D dititanate nanosheets, doping of copper on the dititanate, and immobilization of the nanosheet on GO. The CTGN yields the energy bandgap of 3.07 eV, which is relatively lower than that of the Ti (3.18 eV), TiN (3.14 eV), CTN (3.09 eV) and TGN (3.08 eV), but is higher than that of the CT (2.91 eV), TG (2.53 eV) and CTG (2.28 eV). The bandgap results agree well with the literature since the dititanate provides a higher optical bandgap value when compared to that of the anatase, rutile or brookite TiO2. The photocatalytic performance of the CTGN was observed to be the best at CO2 photoreduction, producing liquid fuels at rates of 113, 157, 265 and 171 μmol gcat−1 h−1 for acetone, methanol, ethanol, and i-propanol. CTGN photoactivity was well supported by photoelectrochemical studies, in which the CTGN revealed a photocurrent of 19.67 × 10−1 ± 5.21 × 10−2 μA cm−2, which is significantly more intense than that of TiN (2.94 × 10−1 ± 3.51 × 10−2 μA cm−2), TGN (4.28 × 10−1 ± 4.68 × 10−2 μA cm−2) and CTG (6.22 × 10−1 ± 0.89 × 10−2 μA cm−2). The PL studies also confirm the strong photoactivity of the CTGN by revealing its low quenching factor, which can be interpreted as a low electron–hole recombination rate.
CTGN performance in photoreducing CO2 to liquid fuels was benchmarked with other research works (Table 1). The group of L. I. Ibarra-Rodríguez4 synthesized Na2Ti6O13/CuO/Cu2O heterostructure via a solid-state and impregnation technique. Their photocatalyst yield CO2 photoreduction products of formaldehyde and ethanol at a production rate of 25 and 4.6 μmol gcat−1 h−1, respectively. N. Lertthanaphol and his team6 (our previous work) utilized the one-step hydrothermal technique in synthesizing the Cu–TiO2/GO composite. The composite exhibited good photoactivity in reducing CO2 to ethanol at a production rate of 233 μmol gcat−1 h−1. P. Seeharaj and her team7 used CeO2/CuO/TiO2 heterostructure photocatalyst for CO2 conversion. They obtained ethanol as the only product at a production rate of 30.5 μmol gcat−1 h−1. H. Hsu and his team29 demonstrated photoactivity of the GO in photoreducing gas-phase CO2. The CO2 was continually fed in the chamber with a GO-coated disk and converted to methanol at 0.172 μmol gcat−1 h−1 of production rate.
Photocatalysts | Experimental details | Bandgap (eV) | Production rate | Ref. |
---|---|---|---|---|
CTGN | Mercury lamp: 160 W; visible + UV | 3.07 | Acetone: 113 μmol gcat−1 h−1 | This study |
CO2 in 20 mL DI water | Methanol: 157 μmol gcat−1 h−1 | |||
Catalyst: 2 mg | Ethanol: 265 μmol gcat−1 h−1 | |||
Quartz reactor: 25 mL | i-propanol: 171 μmol gcat−1 h−1 | |||
Na2Ti6O13–5% CuO/Cu2O | UV-Vis lamp: 4400 μW cm−2; 254 nm | 3.58 | Formaldehyde: 25 μmol gcat−1 h−1 | 4 |
2-psi pressurized CO2 in 200 mL DI water | Methanol: 4.6 μmol gcat−1 h−1 | |||
Catalyst: 100 mg | ||||
Borosilicate reactor: 250 mL | ||||
Cu–TiO2/GO | Mercury lamp: 160 W; visible + UV | 2.11 | Ethanol: 233 μmol gcat−1 h−1 | 6 (previous study) |
CO2 in 25 mL DI water | ||||
Catalyst: 2.5 mg | ||||
Borosilicate reactor: 30 mL | ||||
1% CeO2/3% CuO/TiO2 | Mercury lamp: 15 W; UV | 2.88 | Ethanol: 30.5 μmol gcat−1 h−1 | 7 |
CO2 in 150 mL distilled water | ||||
Catalyst: 150 mg | ||||
Borosilicate reactor with a quartz window | ||||
Graphene oxide | Halogen lamp: 300 W | 3.2–4.4 | Methanol: 0.172 μmol gcat−1 h−1 | 29 |
Continuous gas-flow reactor | ||||
Catalyst: 200 mg | ||||
Quartz reactor: 300 mL |
Footnote |
† Electronic supplementary information (ESI) available. See https://doi.org/10.1039/d2ra04283e |
This journal is © The Royal Society of Chemistry 2022 |