Author Notes
Increasing emissions of carbon dioxide and the depletion of finite fossil fuels have led to many concerns about global warming and energy crises. Consequently, a sustainable and alternative method, photocatalytic CO2 reduction to chemical fuels has received considerable attention. This perspective highlights recent knowledge and the main challenges in CO2 photoreduction primarily from the theoretical field. The fundamental understanding of CO2 adsorption and reaction mechanism at an atomic level is fully addressed, and the relevant effects dominating the process of CO2 photoreduction are also elucidated. Moreover, recent development of photocatalysts including structural modification is presented, which greatly affects the efficiency and selectivity of CO2 conversion. Finally, the possibilities and challenges in this field are discussed.
I. INTRODUCTION
To date, the depletion of finite fossil fuels possesses numerous challenges to the survival of humans, and the pursuit for new energy sources is urgent. Simultaneously, the global warming caused by increasing emissions of the greenhouse gas “CO2,” mainly from the combustion of fossil fuels, induces serious environmental crises on the earth such as the polar ice caps melting and fast rising sea levels.1–3 Therefore, as a sustainable alternative to the depletion of fossil fuels as well as a means to reduce CO2 emissions, reduction of CO2 to useful fuels by water is receiving extensive attention.4–23 Due to the high stability of CO2 (ΔGf0 = −394.4 kJ·mol−1) caused by the highly oxidized state of carbon and stable C=O bonds, the transformation of CO2 to reduced products requires a significant energy input (ΔG > 0).24 Inspired by photosynthesis of plants, photocatalytic reduction of CO2 is gaining increasing interest and considered as one of the most promising strategies to produce renewable fuels as it is driven by photo-excited charge carriers. The first work of CO2 photoreduction, reported by Inoue et al., showed that the photocatalytic reduction of CO2 in an aqueous suspension of a semiconductor can form formaldehyde, formic acid, methanol, and methane.25 Following this breakthrough, numerous researchers have focused on the photocatalytic mechanism of CO2 conversion over a wide variety of semiconductors.8,9,17,22,26–34 However, the reaction efficiency and selectivity for the production of hydrocarbon fuels from CO2 and H2O are still unsatisfactory. A comprehensive understanding of the photoreduction mechanism, including the adsorption and activation of CO2 molecules, and subsequent processes of reduction remains elusive.35,36
There are numerous informative perspectives and reviews introducing current research and recent advances of CO2 photoreduction. These perspectives also highlight design strategies of photocatalysts as well as the primary challenges in the field.3,7,24,35,37–41 Different from those perspectives and reviews, this work mainly focuses on understanding the adsorption and reaction mechanism of CO2 as well as the modification of photocatalysts from theoretical perspectives. The adsorption of CO2 is an important step during the reduction process. The high antibonding orbital of gas-phased CO2 determines its need to adsorb on the surface first in order to be activated before accomplishing photocatalytic reduction.42 The adsorption and activation of CO2 are somewhat complex because it involves photogenerated electrons.10,36 It is therefore important to know how the photogenerated electrons participate in the adsorption and activation of CO2 molecules. In this perspective, the role of the photogenerated electrons in the adsorption and activation of CO2 molecules is elucidated. Additionally, the dominating factors affecting the adsorption and reaction pathways, such as the solvation effect and surface defects, have been addressed. The solvation effect is a significant factor in CO2 reduction as most of the experiments occur in aqueous solution.8,43–45 The water environment can change the adsorption configuration and reaction energies of CO2 photoreduction and is therefore important when discussing the reaction.46 Surface defects such as oxygen vacancies are common during the preparation of photocatalysts and create distinctive electronic structures. This unique defective system can greatly alter the adsorption behavior of CO2 and induce different reaction pathways.47 Upon chemical adsorption, CO2 is reduced by surface charge carriers. The reaction involves multiple proton-coupled electron transfers and can lead to the formation of many different products depending on the chosen reaction pathways, which makes the process rather complex.7,35 Therefore, the reaction mechanism and pathways of CO2 reduction along with their dependent factors will be discussed in detail. The modification of photocatalysts is a substantial and important issue within the literature. It is acknowledged that incorporation of co-catalysts can improve the performance of semiconductor photocatalysts.24 It is well known that the use of co-catalysts provides additional active sites; however, many other structural and electronic changes are evident which enhance the activity of photocatalysts. Thus, the beneficial impact of co-catalyst loading on the photocatalytic system will be fully elucidated. The discussions in this perspective will improve the existing and relevant understanding of CO2 photoreduction and should help future efforts to further improve the overall efficiency and selectivity of the reaction. To conclude this work, a brief perspective on the challenges and possibilities within this field is presented.
II. MECHANISM OF CO2 PHOTOREDUCTION
A. Fundamentals of CO2 photoreduction
B. Adsorption of CO2 molecule
CO2 adsorption has been widely studied on various photocatalysts, in particular, on titanium dioxide, because it is highly stable under photocatalytic conditions, inexpensive to use, and able to perform a variety of photocatalytic reactions such as water splitting.49 Figure 1 illustrates the adsorption of CO2 on a TiO2 surface with different adsorption configurations in the presence of one photogenerated electron. The configuration of A is a linear physical adsorption of CO2, with one oxygen of CO2 weakly bonding with the Ti site. The CO2 molecule can also have chemical bonding with the surface and act as an electron donor and acceptor simultaneously to form configuration B. The lone pair of electrons on the oxygen of CO2 can promote the adsorption of CO2 on the Ti sites and form the adsorption structure C. The carbon atom can also gain electrons from oxygen on the TiO2 surface, forming a carbonate CO3− species (D). The most stable configuration of CO2 adsorption is the linear physical adsorption configuration (A) with O weakly interacting with the adsorption site.10 However, it is proposed that multiple reaction pathways of CO2 photoreduction start with the formation of a CO2⋅− radical, which was detected by infrared spectroscopy on the surface of TiO2 under UV illumination. This was also verified on a MgO surface by electron paramagnetic resonance.33,50–52 Theoretically, CO2⋅− was also found to be activated by CO2 in the presence of a photogenerated electron on the TiO2 surface.53 Moreover, the electronic structure of the activated CO2⋅− is greatly changed, and the lowest unoccupied molecular orbital of CO2⋅− can be even lower than the conduction band (CB) minimum of TiO2, which greatly facilitates CO2 reduction. For the activated CO2⋅− seen in configuration C, we can notice that CO2 adsorbs with a bent configuration and at an OCO angle of 136.9°. The electron spin density largely accumulates around the oxygen of the adsorbed CO2 [see Fig. 1(b)], which indicates the photogenerated electron transfers from the TiO2 surface to CO2 forming CO2⋅−. The adsorption and activation of CO2 can be influenced by several factors including surface defects and the solvation effect, which will be discussed in Secs. II B 1 and II B 2.
(a) Adsorption configurations of CO2 on an anatase (101) photocatalyst.10 The electronic structure (b) illustrates the spin density of configuration C. The iso-surface value was set to 0.05 e/Å3. Reproduced with permission from He et al., Energy Environ. Sci. 5(3), 6196–6205 (2012). Copyright 2012 The Royal Society of Chemistry.
(a) Adsorption configurations of CO2 on an anatase (101) photocatalyst.10 The electronic structure (b) illustrates the spin density of configuration C. The iso-surface value was set to 0.05 e/Å3. Reproduced with permission from He et al., Energy Environ. Sci. 5(3), 6196–6205 (2012). Copyright 2012 The Royal Society of Chemistry.
1. Surface defects for CO2 adsorption
Surface defects commonly occur in the preparation of photocatalytic materials and occasionally act as the active reaction sites since they change the geometrical and electronic structures of photocatalysts. For example, the oxygen vacancy (VO) in a TiO2 system has been found to participate in adsorption and activation of CO2 molecules during photoreduction. A brief schematic diagram of CO2 adsorption at the VO site on the TiO2 (110) surface is shown in Fig. 2(a).18 One of the oxygen atoms of CO2 is located at the VO site and the CO2 molecule is tilted away from the surface by 57° along [110] direction. This particularly linear adsorption configuration was found to have an adsorption energy of 0.44 eV. Using scanning tunneling microscopy (STM), Lee et al. investigated the adsorption of CO2 on the reduced rutile (110) surface at 55 K. As illustrated in Fig. 2(b), they found that CO2 molecules (bright spots) occupy all the VO sites rather than OHb (dull spots). The inset STM images illustrate two CO2 molecules before and after thermal diffusion from their VO sites (dotted ellipses), which depict the CO2 molecules adsorption upon the VO sites. Their STM results are consistent with other work that showed CO2 can adsorb at the VO site on the reduced TiO2 (110) surface with good stability and is also supported further with the results of thermal desorption studies.54,55
(a) Schematic showing an oxygen vacancy defect (VO, black square), a bridging hydroxyl (OHb, black circle), and a CO2 molecule adsorbed at a VO site on the reduced (1 × 1) TiO2 (110) surface. Five-coordinated Ti(5f) atoms and bridging oxygens (Obr) are indicated in red and blue, respectively. The molecular axis of the adsorbed CO2 is perpendicular to the direction of the bridging-oxygen row and is tilted away from the surface normal by 57°, as shown in the inset. (b) STM image (1.5 V, 5 pA, 15 nm × 15 nm) of the TiO2 (110) surface after adsorption of CO2 at 55 K. All the VO sites are occupied by CO2. Three CO2 and two OHb features are marked with diamonds and circles, respectively. The inset shows two STM images (5.1 nm × 2.6 nm) of the same area on the surface. Two CO2 molecules (in the dotted ellipse in the upper inset) diffused away from their VO sites, leaving two intact VO sites visible (lower inset image). Reproduced with permission from Lee et al., J. Am. Chem. Soc. 133(26), 10066–10069 (2011). Copyright 2011 American Chemical Society.
(a) Schematic showing an oxygen vacancy defect (VO, black square), a bridging hydroxyl (OHb, black circle), and a CO2 molecule adsorbed at a VO site on the reduced (1 × 1) TiO2 (110) surface. Five-coordinated Ti(5f) atoms and bridging oxygens (Obr) are indicated in red and blue, respectively. The molecular axis of the adsorbed CO2 is perpendicular to the direction of the bridging-oxygen row and is tilted away from the surface normal by 57°, as shown in the inset. (b) STM image (1.5 V, 5 pA, 15 nm × 15 nm) of the TiO2 (110) surface after adsorption of CO2 at 55 K. All the VO sites are occupied by CO2. Three CO2 and two OHb features are marked with diamonds and circles, respectively. The inset shows two STM images (5.1 nm × 2.6 nm) of the same area on the surface. Two CO2 molecules (in the dotted ellipse in the upper inset) diffused away from their VO sites, leaving two intact VO sites visible (lower inset image). Reproduced with permission from Lee et al., J. Am. Chem. Soc. 133(26), 10066–10069 (2011). Copyright 2011 American Chemical Society.
Huygh et al. comprehensively studied the influence of oxygen vacancies on CO2 adsorption and activation upon an anatase (001) surface using density functional theory (DFT) calculations.56 Four different oxygen vacancies were identified and named as VO1, VO2, VO3, and VO4 as shown in Fig. 3(a). The corresponding formation energies were 4.42, 4.95, 4.76, and 5.71 eV, respectively. However, VO4 was discovered to be highly unstable and was not considered. Although direct formation of the oxygen vacancy is very difficult, it is much easier to form oxygen vacancies from the oxidization of H2O to O2 in the presence of holes. The activation barrier is as low as 0.07 eV.57 Therefore, the oxygen vacancies are possible to be generated. With respect to CO2 adsorption, both bent and linear adsorption configurations were identified. The linear physisorbed configurations were found to easily convert to the bent chemically adsorbed configurations with small activation barriers, which are consistent with the work conducted by He et al.10 Considering CO2 adsorption on the different oxygen vacancies, six stable bent adsorption configurations of CO2 were found, as shown in Fig. 3(b). The relevant results for each configuration are summarized in Table I. In terms of VO1, three configurations of CO2 adsorption were identified. VO1(I) shows a monodentate carbonate configuration for CO2 adsorption with C binding to the formed Lewis base center (Ti3+). VO1(II) has a bent adsorption configuration in which one oxygen of CO2 occupies the oxygen vacancy. In contrast, VO1(III) shows CO2 located near the surface oxygen vacancy. The adsorption structure of VO1(I) is the most stable with an adsorption energy of −2.11 eV resulting from the largest charge transfer from the surface to the CO2 molecule. The adsorption of CO2 upon VO2 was also explored. One adsorption configuration [VO2(I)] was found with CO2 bonding with two Ti3+ ions located in the same [100] row. The adsorption energy for this configuration was calculated to be −1.10 eV. Finally, two monodentate carbonate configurations for CO2 adsorption on VO3 were identified, and in both configurations, the C atom bonds with the Ti site near the oxygen vacancy, however with different orientations of CO2. The adsorption energies are −1.04 and 1.01 eV, respectively. Therefore, the most stable adsorption configuration of CO2 on the defective surface is CO2 adsorbing on the VO1 site with a bent monodentate adsorption configuration. It is also more stable than CO2 adsorbing on the perfect anatase surface (0.68 eV). In conclusion, the oxygen vacancy can be seen to highly promote the adsorption of CO2 molecules.
(a) Illustration showing the different oxygen vacancies of an anatase (001) surface. Gray, red, and yellow are Ti atoms, O atoms, and oxygen vacancies, respectively. (b) The new adsorption configurations of CO2 on a reduced anatase (001) surface with different oxygen vacancies. These are shown from both the side and top views as presented in the upper and lower panels, respectively. The colors of Ti, , , and C atoms are blue, red, red and yellow plus sign, and brown, respectively. Reproduced with permission from Huygh et al., J. Phys. Chem. C 120(38), 21659–21669 (2016). Copyright 2016 American Chemical Society.
(a) Illustration showing the different oxygen vacancies of an anatase (001) surface. Gray, red, and yellow are Ti atoms, O atoms, and oxygen vacancies, respectively. (b) The new adsorption configurations of CO2 on a reduced anatase (001) surface with different oxygen vacancies. These are shown from both the side and top views as presented in the upper and lower panels, respectively. The colors of Ti, , , and C atoms are blue, red, red and yellow plus sign, and brown, respectively. Reproduced with permission from Huygh et al., J. Phys. Chem. C 120(38), 21659–21669 (2016). Copyright 2016 American Chemical Society.
Adsorption geometries and energies of CO2 on the reduced anatase (001) surface with different oxygen vacancies. Reproduced with permission from Huygh et al., J. Phys. Chem. C 120(38), 21659–21669 (2016). Copyright 2016 American Chemical Society.
. | dTi–O(CO2) . | dC–O(TiO2) . | dC–O(CO2) . | (OCO) . | Eads . |
---|---|---|---|---|---|
Configuration . | (Å) . | (Å) . | (Å) . | (deg) . | (eV) . |
CO2 (g) | 1.18, 1.18 | 180 | |||
VO1(I) | 2.06, 1.97 | 1.27, 1.29 | 125.1 | −2.11 | |
VO1(II) | 1.97, 2.64, 1.88 | 1.29, 1.35 | 108.9 | −1.10 | |
VO1(III) | 2.02 | 1.47 | 1.20, 1.30 | 134.7 | −1.64 |
VO2(I) | 2.01, 1.99 | 1.27, 1.27 | 126.1 | −1.10 | |
VO3(I) | 1.93 | 1.37 | 1.22, 1.34 | 128.6 | −1.04 |
VO3(II) | 1.91 | 1.37 | 1.21, 1.35 | 126.9 | −1.01 |
. | dTi–O(CO2) . | dC–O(TiO2) . | dC–O(CO2) . | (OCO) . | Eads . |
---|---|---|---|---|---|
Configuration . | (Å) . | (Å) . | (Å) . | (deg) . | (eV) . |
CO2 (g) | 1.18, 1.18 | 180 | |||
VO1(I) | 2.06, 1.97 | 1.27, 1.29 | 125.1 | −2.11 | |
VO1(II) | 1.97, 2.64, 1.88 | 1.29, 1.35 | 108.9 | −1.10 | |
VO1(III) | 2.02 | 1.47 | 1.20, 1.30 | 134.7 | −1.64 |
VO2(I) | 2.01, 1.99 | 1.27, 1.27 | 126.1 | −1.10 | |
VO3(I) | 1.93 | 1.37 | 1.22, 1.34 | 128.6 | −1.04 |
VO3(II) | 1.91 | 1.37 | 1.21, 1.35 | 126.9 | −1.01 |
The adsorption of CO2 in the presence of oxygen vacancies was also investigated on a Cu2O (111) surface using DFT by Wu and co-workers.58 The adsorption of CO2 can be promoted when the CO2 molecule occupies the oxygen vacancy and forms the reactive CO2⋅− radical. When CO2 adsorbs on the Cu and O sites near the formed oxygen vacancy, the adsorption energies are even smaller than those corresponding with adsorption on the perfect surface. Therefore, the oxygen vacancy only delivers enhanced activity for CO2 adsorption and if CO2 adsorbs anywhere else, only limited activity will be observed.
In general, the oxygen vacancies play an important role in the photoreduction of CO2 on the surface of photocatalysts. The vacancies greatly improve the adsorption and activation of CO2 when CO2 molecules adsorb upon them, which is beneficial in the conversion of CO2 to useful products.
2. Solvation effect for CO2 adsorption
In numerous experiments reported within the literature, the photocatalytic reduction of CO2 is proceeded with water because of its hydrogen resource. Using H2O as a reducing agent over (Au, Cu)/TiO2 materials, Neatu et al. reported a high production rate of 2.2 ± 0.3 mmol g−1 h−1 for CO2 photoreduction to CH4.5,59 Wang et al. also found that the photocatalytic reduction of CO2 can be significantly enhanced in the presence of H2O using a Pt@Cu2O co-catalyst loaded on TiO2.60 Accordingly, the effect of water should be elucidated. The interactions of CO2 and water over titania nanoparticles were thoroughly investigated using electron paramagnetic resonance.43 The detection of redox products including H atoms, OH radicals, and CH3 radicals indicates water participating in CO2 photoreduction.
Aside from acting as a H donator, the solvation effect of H2O is also an important issue and has been extensively studied.11,61,62 Yin et al. simulated the adsorption of CO2 on rutile TiO2(110) situated in a water environment using the periodic continuum solvation model (PCSM).46 In total, four typical configurations of CO2 adsorption were examined on the rutile (110) surface as shown in Fig. 4(a). C-1 is a tilted monodentate configuration with one oxygen of CO2 adsorbing at a Ti5c site; C-2 is a horizontal bidentate configuration adsorbing along the surface Ti5c row; C-3 shows CO2 adsorbing between two adjacent bridge oxygen atoms; similarly, C-4 is the adsorption of CO2 on the top of one bridge oxygen atom. From Table II, C-2 is the most stable adsorption configuration in the gas phase with an adsorption energy of −0.34 eV. When the solvation effect is considered, the adsorption energy of CO2 can be increased by about 0.3 eV relative to its adsorption in the gas phase. Interestingly, they also found that the most stable configuration of CO2 adsorption changes from C-2 to C-1 with solvation effect, which is consistent with the results from low temperature scanning tunneling microscopy (STM).63,64
Front (uppermost row) and top (bottom row) views of CO2 (a) adsorbed and (b) co-adsorbed with H2O on a rutile (110) surface. The O atom is red and the Ti atom is light-blue for the TiO2 system, while the O atom is orange and the C atom is green for the CO2 molecule. Reproduced with permission from Yin et al., J. Phys. Chem. Lett. 6(13), 2538–2545 (2015). Copyright 2015 American Chemical Society.
Front (uppermost row) and top (bottom row) views of CO2 (a) adsorbed and (b) co-adsorbed with H2O on a rutile (110) surface. The O atom is red and the Ti atom is light-blue for the TiO2 system, while the O atom is orange and the C atom is green for the CO2 molecule. Reproduced with permission from Yin et al., J. Phys. Chem. Lett. 6(13), 2538–2545 (2015). Copyright 2015 American Chemical Society.
Binding energies and representative geometrical parameters for the adsorption (defined as C-i, i = 1–4) and co-adsorption of CO2 and H2O molecules (defined as CW-i, i = 1–6) on the rutile (110) surface. The binding energies of CO2 with solvation effect are highlighted in bold. Reproduced with permission from Yin et al., J. Phys. Chem. Lett. 6(13), 2538–2545 (2015). Copyright 2015 American Chemical Society.
. | Eads (eV) . | O–Ti (Å) . | C–Ob (Å) . | ∠OCO (deg) . |
---|---|---|---|---|
C-1 | 0.33 | 2.69 | 3.05 | 178.16 |
−0.69 | 2.69 | 3.05 | 178.52 | |
C-2 | 0.34 | 3.14 | 3.91 | 179.20 |
−0.67 | 3.15 | 3.92 | 178.94 | |
C-3 | 0.22 | 3.93 | 3.07 | 179.12 |
−0.57 | 4.00 | 3.10 | 179.09 | |
C-4 | 0.19 | 4.44 | 2.97 | 178.86 |
−0.54 | 4.44 | 2.97 | 178.92 | |
CW-1 | −0.40 | 2.94 | 3.22 | 178.76 |
−0.75 | 2.86 | 3.22 | 179.07 | |
CW-2 | 0.42 | 2.60 | 3.11 | 178.15 |
−0.81 | 2.51 | 3.10 | 178.24 | |
CW-3 | 0.43 | 3.03 | 3.88 | 179.73 |
−0.75 | 2.94 | 3.89 | 179.17 | |
CW-4 | 0.30 | 4.73 | 2.82 | 177.38 |
−0.65 | 4.71 | 2.83 | 178.14 | |
CW-5 | 0.46 | 4.68 | 2.84 | 177.60 |
−0.83 | 4.70 | 2.85 | 178.04 | |
CW-6 | 0.39 | 4.38 | 3.05 | 178.15 |
−0.75 | 4.49 | 2.91 | 178.85 |
. | Eads (eV) . | O–Ti (Å) . | C–Ob (Å) . | ∠OCO (deg) . |
---|---|---|---|---|
C-1 | 0.33 | 2.69 | 3.05 | 178.16 |
−0.69 | 2.69 | 3.05 | 178.52 | |
C-2 | 0.34 | 3.14 | 3.91 | 179.20 |
−0.67 | 3.15 | 3.92 | 178.94 | |
C-3 | 0.22 | 3.93 | 3.07 | 179.12 |
−0.57 | 4.00 | 3.10 | 179.09 | |
C-4 | 0.19 | 4.44 | 2.97 | 178.86 |
−0.54 | 4.44 | 2.97 | 178.92 | |
CW-1 | −0.40 | 2.94 | 3.22 | 178.76 |
−0.75 | 2.86 | 3.22 | 179.07 | |
CW-2 | 0.42 | 2.60 | 3.11 | 178.15 |
−0.81 | 2.51 | 3.10 | 178.24 | |
CW-3 | 0.43 | 3.03 | 3.88 | 179.73 |
−0.75 | 2.94 | 3.89 | 179.17 | |
CW-4 | 0.30 | 4.73 | 2.82 | 177.38 |
−0.65 | 4.71 | 2.83 | 178.14 | |
CW-5 | 0.46 | 4.68 | 2.84 | 177.60 |
−0.83 | 4.70 | 2.85 | 178.04 | |
CW-6 | 0.39 | 4.38 | 3.05 | 178.15 |
−0.75 | 4.49 | 2.91 | 178.85 |
To further explore the role of water, they also investigated the co-adsorption of CO2 and H2O on the rutile (110) surface. From Fig. 4(b), six configurations were considered and denoted as Cw-i, i = 1–6. As shown in Table II, the co-adsorption effect was also found to greatly affect the adsorption of CO2 on rutile (110). Relative to single CO2 adsorption, the co-adsorption energy of CO2 and H2O increases from −0.06 to −0.12 eV. It was suggested that the increased co-adsorption energy may result from hydrogen bonding between CO2 and H2O. When the solvation effect is also considered, the adsorption energies change by about 0.3 eV and the most stable adsorption configuration changes from C-2 to Cw-2. Furthermore, the interactions of the surface with adsorbate are also changed with parameters shown in Table II. Compared to single CO2 adsorption in the gas phase, the distances of the O–Ti bond in Cw-1 and Cw-3 are significantly decreased by about 0.11 Å. When solvation is included, the corresponding distances are also decreased by about 0.1 Å. The O–C–O angle is also affected by the solvation and co-adsorption effect. The co-adsorption effect enlarges the O–C–O angle by 0.5° in the gas phase, while the solvation effect changes the O–C–O angle by 0.1°–0.7°. A similar trend was discovered by Sorescu et al. from co-adsorption with H2O on rutile (110) at 1/8 monolayer (ML) and 1 ML coverages.11 They found that co-adsorbed H2O molecules generally lead to a small increase in the stability of CO2 on the surface through the formation of hydrogen bonds. The adsorption of CO2 is greatly influenced by H2O and the solvation effect can even change the most stable adsorption configuration of CO2. In conclusion, the solvation effect and co-adsorption of H2O significantly affect the adsorption of CO2.
C. Thermodynamics and kinetics of CO2 reduction
Following the adsorption of CO2, it is worth stressing that the subsequent mechanism of photocatalytic CO2 reduction, including electron and proton transfers, is also vital. The chosen pathway affects the efficiency of these types of reactions and even determines the final products. As shown in Eqs. (2)–(6), the conversion of CO2 to useful chemicals consists of steps involving C=O bond breaking and C–H bond formation. The conversion reaction can contain as many as eight electron and four proton transfers for CH4 formation. Depending on the different adsorption states of the CO2 molecules and the number of electron and proton transfers, the reaction can lead to the formation of many different products including HCOOH, HCHO, CH3OH, and CH4. Therefore, the photocatalytic mechanism of CO2 is considerably complicated.
To date, many researchers have been dedicated to explicate the reaction mechanism of CO2 reduction.4,6,60,65–67 Based on different CO2 adsorption models, He et al. investigated the photoreduction of CO2 to HCOOH on a perfect anatase (101) surface. They discovered two competing reaction mechanisms.10 The first mechanism involves a single-electron (1e−) reduction process (CO2 + e− → CO2⋅−; CO2⋅− + H+ + e− → HCOO−; HCOO− + H+ → HCOOH), in which the first step is the chemical adsorption of CO2 to CO2⋅− by one electron followed by the conversion of CO2⋅− to HCOO− by a proton-coupled electron mechanism [see Fig. 5(a)]. The other mechanism is a concerted two-electron process from a linear physisorption configuration of CO2 (CO2⋅− + H+ + 2e− → HCOO−; HCOO− + H+ → HCOOH) as shown in Fig. 5(b). For the first reduction mechanism, CO2 is initially activated to the bent CO2⋅− bidentate by one electron from a physisorbed CO2 configuration. The activation barrier is 0.87 eV. This step is followed by a proton-coupled electron mechanism to form HCOO− with a barrier of 0.46 eV. Then HCOO− receives a proton from the breaking of one of the Ti–O bonds to form HCOOH. The barrier for this step is 0.46 eV. Thus, the rate determining step is the activation of CO2 to CO2⋅−. The second mechanism is a concerted proton-coupled electron process of CO2 photoreduction, in which CO2 is activated by two electrons and one proton simultaneously. The structure of the transition state (TS) is illustrated in Fig. 5(b), in which CO2 is bent (∠OCO = 140°) with carbon bonded toward an adsorbed proton by a hydrogen bond of 2.1 Å. The energy barrier along this pathway (0.82 eV) is slightly lower than that of pathway I (0.87 eV) and is therefore more favorable. Interestingly, this mechanism leads to a linear HCOO− adsorbed configuration which has a low proton hydrogenation barrier of 0.01 eV. Thus, it is thermodynamically favorable to CO2 to reduce to HCOOH. However, there is a low possibility that two electrons will participate in the reaction simultaneously, and therefore this step may be difficult to occur. Accordingly, the relevant micro-kinetic studies including both the kinetics and electron concentration are worth being investigated in the future.
Illustration of reaction pathways. (a) Pathway I: via the B1 configuration to form HCOOH; (b) pathway II: via the A1 configuration to form HCOOH (concerted mechanism). The sum of the energies for CO2 and H2 molecules is the zero reference for energy. The colors of Ti, O, C, and H atoms are green, red, silver, and white, respectively. The symbols in brackets label the states: g is the gas phase; a represents adsorbed species on the anatase (101) surface; A1 and B1 are the two adsorption configurations of A and C shown in Fig. 1(a) and correspond to CO2/CO2⋅− on the surface; A1H and B1H are the adsorption configurations of HCOO on the surface; A1H2 and B1H2 are the adsorption configurations of HCOOH on the surface. The sign of “…” indicates two species in proximity. Reproduced with permission from He et al., Energy Environ. Sci. 5(3), 6196–6205 (2012). Copyright 2012 Royal Society of Chemistry.
Illustration of reaction pathways. (a) Pathway I: via the B1 configuration to form HCOOH; (b) pathway II: via the A1 configuration to form HCOOH (concerted mechanism). The sum of the energies for CO2 and H2 molecules is the zero reference for energy. The colors of Ti, O, C, and H atoms are green, red, silver, and white, respectively. The symbols in brackets label the states: g is the gas phase; a represents adsorbed species on the anatase (101) surface; A1 and B1 are the two adsorption configurations of A and C shown in Fig. 1(a) and correspond to CO2/CO2⋅− on the surface; A1H and B1H are the adsorption configurations of HCOO on the surface; A1H2 and B1H2 are the adsorption configurations of HCOOH on the surface. The sign of “…” indicates two species in proximity. Reproduced with permission from He et al., Energy Environ. Sci. 5(3), 6196–6205 (2012). Copyright 2012 Royal Society of Chemistry.
The mechanism of photoreduction for CO2 conversion to CH4 was also investigated theoretically on the anatase TiO2(101) surface by Ji and Luo.36 Similarly to the photoreduction of CO2 to HCOOH, two reaction pathways to form CH4 were identified: pathway I proceeds via an intermediate of HCOOH (CO2 → HCOOH → H2CO → CH3OH → CH4), while pathway II gives rise to the formation of CH4 by the intermediate of CO (CO2 → CO → H2CO → CH3OH → CH4). Both pathways prefer the two-electron mechanism where two photogenerated electrons concertedly participate in one elemental step. For pathway I, the initial step is the photoreduction of the CO2 molecule to form HCOOH as shown in Eq. (2), and this was also reported in the work of He et al.10 The rate-determining step is the activation of CO2, in which the linear CO2 converts to a bent HCOO with a proton and an electron. The barrier is 1.49 eV and it is higher than that in the work of He et al. The different barriers for this step might be due to the different sizes of supercell used in their studies. In the second step, the hydrogenation of HCOO has virtually no barrier, meaning that the effective barrier of HCOOH formation from CO2 is 1.49 eV (see Fig. 6). Comparatively, the deoxygenation of CO2 to CO in pathway II appears to be a little easier with an energy barrier of 1.41 eV. Although the energy barrier for the formation of HCOO from CO is higher, 0.88 eV, the effective barrier for this path is lower than pathway I by 0.08 eV. This indicates that HCOOH is more favorable to form from pathway II than CO from pathway I. The subsequent H2CO formation also has two pathways. The first is the formation of H2CO from the deoxygenation of HCOOH and another is from the hydrogenation of CO. In the photoreduction of HCOOH to H2CO, it is more inclined to have a concerted mechanism to form H2CO directly with two electrons participating simultaneously (HCOOH + 2e− + 2H+ → H2CO + H2O). The energy barrier was calculated to be 1.84 eV. For the hydrogenation of CO, a one-electron photoreduction is observed, in which a HCO⋅ intermediate forms first and then it couples with a proton as well as an electron to form H2CO. However, it should be noted that the HCO intermediate is difficult to generate with an energy barrier as high as 2.08 eV. This step is also thermodynamically unfavorable. For the hydrogenation of CO to H2CO (CO + 2e− + 2H+ → H2CO), the two-electron path is also more favorable than the one-electron mechanism. In the two-electron reduction process, HCO− will form after one proton transfers concertedly with two electrons to CO and then couples with a proton to form H2CO. The effective reaction barrier is 1.07 eV and is a little lower than that of the one-electron mechanism by 0.12 eV. Considering the HCOOH path, the hydrogenation of CO is highly favorable. Subsequent reactions to form CH4 proceed with relative ease as shown in Fig. 6. The energy barriers of H2CO to CH3OH and CH3OH to CH4 are 0.60 and 0.95 eV, respectively. From their study, the rate-determining step for pathway I is the photoreduction of HCOOH to H2CO, while that of pathway II is the photoreduction of CO2 to CO. The lower barrier of pathway II compared with that of pathway I indicates that it is more favorable to CO2 to be reduced to CH4 via the CO intermediate. The rate-determining step of CO2 to CO also suggests that a surface oxygen vacancy could be an active site for photocatalytic CO2 reduction and reduce the energy required for the deoxygenation of CO2. This possibility will be discussed later.
Energy profile for the photocatalytic reduction of CO2 to CH4. Adapted with permission from Y. Ji and Y. Luo, ACS Catal. 6(3), 2018–2025 (2016). Copyright 2016 American Chemical Society.
Energy profile for the photocatalytic reduction of CO2 to CH4. Adapted with permission from Y. Ji and Y. Luo, ACS Catal. 6(3), 2018–2025 (2016). Copyright 2016 American Chemical Society.
1. Effect of H2O on the mechanism of photocatalytic CO2 reduction
In aqueous photocatalytic systems, water can act as donors of both protons [Eqs. (2)–(6)] and electrons [Eqs. (7)–(9)]. Taking the TiO2 system as an example, H2O dissociates on the surface of TiO2 through its oxygen atom adsorbing at a surface Ti site, forming terminal and bridging OH groups.11,46,68–70 The dissociated OH and H on the surface can interact with the external molecular H2O network via hydrogen bonding.71 Possible interactions with the large water environment can have a considerable impact on the mechanism of CO2 photoreduction.
Dimitrijevic and co-workers addressed the role of water in the overall mechanism of CO2 reduction on the TiO2 surface from both experimental work and theoretical calculations.43 They used a 1:1 molar ratio of CO2/H2O in the system, and upon illumination, H atoms, ⋅OCH3 radicals, and ⋅CH3 radicals were detected. This suggests competitive transfer of electrons to adsorbed protons and adsorbed CO2 on the surface of TiO2. Thus, water acts as an electron acceptor for the H atoms formed from the reaction of photogenerated electrons with protons on the surface. In their study, the initial step of CO2 photoreduction was found to be the photocatalytic reduction of CO2 to HCOO− and was also verified by the first principles calculation (see Fig. 7). In the presence of two photogenerated electrons and one adsorbed proton next to a CO2 molecule, CO2 can be activated and converted to a formate on the TiO2 surface (CO2 + 2e− + H+ → HCOO−). This involves a concerted two-electron, one-proton mechanism. In the transition state (TS, see Fig. 7), the distance between carbon and the adsorbed proton is 2.1 Å. The configuration of CO2 is bent (∠OCO = 140°) and bonds strongly to the surface five-fold Ti site, with a Ti–O bond length of 2.05 Å. The effective barrier for the formation of formate was calculated to be 0.82 eV. The relatively high activation barrier of 0.82 eV also explains the relatively low efficiency for CO2 photoreduction on the pure surface of TiO2 at room temperature.
Energy profile showing the reaction pathway from reactants CO2 and hydrogen to product formate on the anatase (101) surface. Important bond lengths (Å) and angles (°) are marked in all structures with atom colors showing Ti in gray, O in red, C in blue, and H in white, respectively. The species along the reaction pathway are labeled at the bottom, where “g” denotes gas phase and “a” denotes adsorbed species on the anatase (101) surface; A1 is a linear vertical adsorption configuration of CO2 on the surface. The “+” sign indicates non-interacting species, while “…” indicates two species in close proximity. Adapted with permission from Dimitrijevic et al., J. Am. Chem. Soc. 133(11), 3964–3971 (2011). Copyright 2011 American Chemical Society.
Energy profile showing the reaction pathway from reactants CO2 and hydrogen to product formate on the anatase (101) surface. Important bond lengths (Å) and angles (°) are marked in all structures with atom colors showing Ti in gray, O in red, C in blue, and H in white, respectively. The species along the reaction pathway are labeled at the bottom, where “g” denotes gas phase and “a” denotes adsorbed species on the anatase (101) surface; A1 is a linear vertical adsorption configuration of CO2 on the surface. The “+” sign indicates non-interacting species, while “…” indicates two species in close proximity. Adapted with permission from Dimitrijevic et al., J. Am. Chem. Soc. 133(11), 3964–3971 (2011). Copyright 2011 American Chemical Society.
Furthermore, in their study, they also identified that water induces a better charge separation in this system. It was found that the number of spins increases from 0.8 × 1015 spins/cm3 for the dry TiO2 system to 1.0 × 1015 spins/cm3 for the aqueous TiO2 system. From Fig. 8(a), the intensities of signals associated with trapped electrons in the aqueous TiO2 system are evidently higher than those in the dry TiO2 system. To give a better understanding, the signals of oxidation products normalized for different spectra are presented [see Fig. 8(b)]. As the EPR spectra were recorded under illumination, the intensities of the signals correspond to the steady-state concentrations of products. This is ruled by the steady recombination of electrons and holes and the formation of products from subsequent reactions. In the presence of H2O, the charges are stabilized due to the dipolar interaction with water and the band bending at the interface of TiO2/H2O. This suppresses the recombination of charges.
(a) EPR spectra of dry TiO2 (black line) and aqueous TiO2 (red line) under illumination at 4.5 K. (b) EPR spectra of 10 mg of TiO2 under illumination at 4.5 K in the presence of (i) 5 mmol of CO2 and 2 mmol of H2O (blue line), (ii) 2 mmol of H2O (red line), and (iii) 0.5M aqueous Na2CO3 (black line). The simulated spectrum is presented as a gray line. Power, 0.2 mW; modulation amplitude, 0.5 mT. The light source was 355 nm photons from a Nd:YAG laser (power, 9 mJ). The inset shows part of the EPR spectra that correspond to photogenerated electrons of the same samples illuminated and measured at 50 K. Reproduced with permission from Dimitrijevic et al., J. Am. Chem. Soc. 133(11), 3964–3971 (2011). Copyright 2011 American Chemical Society.
(a) EPR spectra of dry TiO2 (black line) and aqueous TiO2 (red line) under illumination at 4.5 K. (b) EPR spectra of 10 mg of TiO2 under illumination at 4.5 K in the presence of (i) 5 mmol of CO2 and 2 mmol of H2O (blue line), (ii) 2 mmol of H2O (red line), and (iii) 0.5M aqueous Na2CO3 (black line). The simulated spectrum is presented as a gray line. Power, 0.2 mW; modulation amplitude, 0.5 mT. The light source was 355 nm photons from a Nd:YAG laser (power, 9 mJ). The inset shows part of the EPR spectra that correspond to photogenerated electrons of the same samples illuminated and measured at 50 K. Reproduced with permission from Dimitrijevic et al., J. Am. Chem. Soc. 133(11), 3964–3971 (2011). Copyright 2011 American Chemical Society.
2. Photoreduction of CO2 in the defective system
The mechanism of CO2 photoreduction at oxygen vacancies on the TiO2 surface was investigated by Ji and Luo.47 As shown in Fig. 9(a), they found that CO2 is more favorable to dissociate at VO (0.73 eV) than on the perfect surface (1.19 eV). This is consistent with the previous work conducted by Sorescu et al.11 The subsequent hydrogenation of CO to CH3OH and CH4 has been thoroughly studied. The hydrogenation of CO to CH3OH follows four steps: CO → HCO → CH2O → CH3O → CH3OH [see Fig. 9(c)]. The rate-determining step was found to be the hydrogenation of CH2O to CH3O at the VO site. The energy barrier was calculated to be 1.26 eV. It is lower than the rate-determining step on the perfect surface by 0.15 eV,36 indicating that the VO site is more active than the Ti site. In addition, the hydrogenation step of CH3O to CH3OH at the VO site is endothermic by 0.65 eV with a barrier of 0.71 eV. The adsorption energy of the formed CH3OH at the VO vacancy is 1.33 eV. It indicates that CH3OH prefers to dissociate to CH3O by overcoming a small barrier of 0.06 eV. For CH4 formation [see Figs. 9(b) and 9(c)], the path is as follows: CO → HCO → CH2O → CH3O → CH3⋅→ CH4. The rate-determining step is also the hydrogenation of CH2O to CH3O at the VO site. However, the subsequent path from CH3O to CH4 is exothermic with an effective barrier of 0.71 eV. Therefore, CH4 is favorable to form at the VO site. From their study, the oxygen vacancy on the defective surface is more active than the surface Ti site. This can greatly lower the energy barrier of the deoxygenation processes so that the photoreduction of CO2 can be achieved more easily. The regeneration of oxygen vacancies was also investigated. Experimentally, Fujishima et al. found that OV can be generated under UV illumination during the generation of O2.72 Theoretically, Li and Selloni reported that OV can be generated with a low barrier in the presence of holes.57 Therefore, the oxygen atom is possible to be removed to regenerate the O vacancy in the TiO2 system.
(a) Adsorption and direct dissociation of CO2 at the VO site. (b) Formation of CH3⋅ and CH4. Step M-I, formation of CH3⋅ via the breaking of the C–O bond; step M-II, an electron transfers to CH3⋅, forming a CH3− anion; and step M-III, the CH3− anion recombines with a Hb to form CH4. (c) Energy profile for the hydrogenation of CO to CH3OH. Step H-I, hydrogenation of CO to HCO; step H-II, HCO to CH2O; step H-III, CH2O to CH3O; and step H-IV, CH3O to CH3OH. Reproduced with permission from Y. Ji and Y. Luo, J. Am. Chem. Soc. 138(49), 15896–15902 (2016). Copyright 2016 American Chemical Society.
(a) Adsorption and direct dissociation of CO2 at the VO site. (b) Formation of CH3⋅ and CH4. Step M-I, formation of CH3⋅ via the breaking of the C–O bond; step M-II, an electron transfers to CH3⋅, forming a CH3− anion; and step M-III, the CH3− anion recombines with a Hb to form CH4. (c) Energy profile for the hydrogenation of CO to CH3OH. Step H-I, hydrogenation of CO to HCO; step H-II, HCO to CH2O; step H-III, CH2O to CH3O; and step H-IV, CH3O to CH3OH. Reproduced with permission from Y. Ji and Y. Luo, J. Am. Chem. Soc. 138(49), 15896–15902 (2016). Copyright 2016 American Chemical Society.
III. CO2 REDUCTION ON SEMICONDUCTOR PHOTOCATALYSTS
A. Overview of semiconductor photocatalysts
Since the photoreduction of CO2 was first reported on TiO2 in 1979,25 numerous studies have been focused on the photocatalytic mechanism, activity, and selectivity of CO2 reduction. However, the efficiency of CO2 conversion is also largely controlled by the performance of the photocatalysts.13,73–80 The ability to enhance the photocatalytic efficiency by screening the proper semiconducting photocatalysts is also an important issue.24,81–86 Semiconductor photocatalysts play two main roles in photocatalytic reactions. The first role is producing the photogenerated electrons and holes to participate in the reduction and oxidation reactions. Another is providing the active sites for reactions to occur. The use of semiconductors as photocatalysts is attractive because of their appropriate band structure. The band structure consists of a conduction band (CB) and a valence band (VB) with a bandgap inside (see Fig. 10).7,24,87 The VB lies below the gap and is occupied by electrons at the ground state, while the CB lies above the gap and is unoccupied. Upon illumination, an electron is excited from the VB into the CB, thereby leaving an empty state in the VB (this constitutes a hole). Once the charge carriers are separated, they may migrate to the surface of the photocatalyst and the adsorbed molecules, thereby initiating the redox reactions. For the photocatalytic reduction of CO2, the photogenerated electrons transfer from the photocatalyst to the adsorbed CO2 to reduce CO2. Oppositely, the holes participate in the oxidizing reactions, for example, the oxidation of water. This requires the semiconductor to have a higher CB energy level than the CO2 reduction potential to drive the photogenerated electrons from the CB to the surface CO2 species. Therefore, an ideal photocatalyst for CO2 reduction should satisfy the following fundamental requirements: (1) high light absorption efficiency; (2) excellent ability for spatial charge carrier separation; (3) small electron/hole effective masses; (4) narrow bandgap with suitable CB and VB energy levels; and (5) active sites for CO2 adsorption on the surface. A single semiconductor does not usually satisfy all these requirements. Moreover, because of the fast recombination of photogenerated electrons and holes plus the lack of appropriate reaction sites, single semiconductor photocatalysts usually do not show high efficiency in photocatalytic reactions. This is verified by a higher activity observed on the metal-loaded TiO2 photocatalysts compared with pure TiO2 during the photocatalytic reduction of CO2.5 Although the pure TiO2 shows low activity, it is worth noting that heterojunction structures were identified in the TiO2 system recently, linking anatase with rutile.88 This kind of structure allows the migration of holes and hinders the electron transfer from anatase to rutile. This would improve the separation of charge carriers and enhance the photocatalytic efficiency. Thus, it also deserves to investigate CO2 photoreduction on this kind of heterojunction structure in the future.
Schematic diagram of CO2 photoreduction in the presence of H2O on photocatalysts.
Schematic diagram of CO2 photoreduction in the presence of H2O on photocatalysts.
B. Modification of photocatalysts on photocatalytic CO2 reduction
1. Role of co-catalysts
To improve the efficiency of photocatalytic reactions, composite photocatalysts were widely developed by loading proper co-catalysts such as metal or metal oxides on semiconductor photocatalysts. In composite photocatalysts, the roles of co-catalysts include87,89 (1) providing trapping sites and promoting the separation of the photogenerated charges to enhance the efficiency; (2) modifying the electronic structures of photocatalysts to promote the adsorption of reactants; and (3) catalyzing the reactions by lowering the activation energy. Numerous kinds of co-catalysts have been incorporated into the photocatalysts to improve the activity of photocatalytic CO2 reduction,90–94 including metal and metal oxide co-catalysts.
2. Metal and metal oxide co-catalysts
Loading noble metals such as Pt, Au, Ag, and Cu on photocatalysts have been reported and have shown an enhanced activity of CO2 photoreduction.17,26,59,95–97 Umezawa et al. investigated the effect of Pt clusters on the photoreduction of CO2 to fuel with water on a rutile TiO2(110) surface based on the reaction CO2 + H2O + 2e− → HCOO− + OH−.96 In their study, the co-adsorption of CO2/H2O and HCOO/OH species on the TiO2(110) surface in the absence and presence of Pt clusters was fully explored. Relative to the physical adsorption of CO2 and H2O, the chemical adsorption energy of HCOO and OH is highly promoted when these species adsorb on the TiO2 surface with Pt clusters in the vicinity. This is most apparent when they absorb on a Pt cluster [see Figs. 11(a) and 11(b)]. It is consistent with the experimental observation which revealed that Pt nanoparticles loaded on the TiO2 surface can improve the efficiency of CO2 photoreduction to CH4.13 Figure 11(c) illustrates three models of HCOO and OH adsorption. Model A shows a structure of HCOO and OH adsorbing on a pure TiO2 surface. HCOO adopts a bidentate adsorption configuration, whilst OH takes up a monodentate configuration. Model B also presents the same adsorption configuration with a Pt cluster neighboring. Finally, Model C illustrates the adsorption of HCOO and OH on the Pt sites. From analyzing the density of states (DOS’s), they could rationalize the change of adsorption of HCOO and OH in the presence of the Pt cluster. They reported that two electrons were transferred from the HOMO of the Pt cluster to the empty states at the top of the valence band of adsorbates. This two-electron transfer helps facilitate the ionization of HCOO− and OH−. The strong chemisorption of HCOO− and OH− on a Pt cluster further stabilizes the system. Therefore, in this system, the introduction of a Pt co-catalyst provides active sites for the adsorption of various reactants and intermediates whilst simultaneously acting as a reservoir of electrons to promote the adsorption of adsorbates. Furthermore, Yang et al. investigated the effect of different sizes of Pt clusters from 2D to 3D morphology toward CO2 photoreduction on the anatase TiO2 (101) surface.42 Compared to the 2D Pt cluster, they found that the 3D Pt cluster is more beneficial to the formation of the bent CO2− configuration, with electronic charge accumulation at the carbon atom. The differences can be attributed to the following reasons: (1) the 3D Pt cluster provides more accessible Pt–TiO2 interfacial active sites which are beneficial to the activation of CO2 and charge transfer; (2) the 3D cluster possesses higher structural relaxation than the 2D cluster, which highly promotes the binding strength of CO2.
(a) Energy diagram displaying the chemisorption energies of HCOO and OH with respect to physisorption of CO2 and H2O on TiO2(110). Blue, red, black, and pink particles are Ti, O, C, and H atoms, respectively. (b) Energy diagram for chemisorption energies of HCOO and OH on various adsorption sites on Pt10/TiO2(110) with respect to the physisorption of CO2 and H2O on a trough of TiO2(110) in Pt10/TiO2(110). Gray particles are Pt atoms. (c) Upper panel shows the adsorption structures of HCOO and OH on the Ti sites of a clean TiO2(110) surface (Model A), the Ti sites of Pt10/TiO2(110) surface (Model B), and the Pt sites of Pt10/TiO2(110) (Model C). Adsorption energies for each model with respect to CO2 and H2O co-adsorption on the same support are also shown below each structure. Lower panel illustrates the total and local density of states (DOS’s) for Models A, B, and C. The width of smearing was set at 0.1 eV. The zero energy is at the conduction-band minimum of a clean surface of TiO2(110) and is corrected for the surfaces containing Pt10 or adsorbates such that the averaged electrostatic potential of a Ti atom located at the back side of the slab is aligned with that of the clean TiO2(110). The vertical broken line denotes the highest occupied state of each system. Reproduced with permission from Umezawa et al., J. Phys. Chem. C 120(17), 9160–9164 (2016). Copyright 2016 American Chemical Society.
(a) Energy diagram displaying the chemisorption energies of HCOO and OH with respect to physisorption of CO2 and H2O on TiO2(110). Blue, red, black, and pink particles are Ti, O, C, and H atoms, respectively. (b) Energy diagram for chemisorption energies of HCOO and OH on various adsorption sites on Pt10/TiO2(110) with respect to the physisorption of CO2 and H2O on a trough of TiO2(110) in Pt10/TiO2(110). Gray particles are Pt atoms. (c) Upper panel shows the adsorption structures of HCOO and OH on the Ti sites of a clean TiO2(110) surface (Model A), the Ti sites of Pt10/TiO2(110) surface (Model B), and the Pt sites of Pt10/TiO2(110) (Model C). Adsorption energies for each model with respect to CO2 and H2O co-adsorption on the same support are also shown below each structure. Lower panel illustrates the total and local density of states (DOS’s) for Models A, B, and C. The width of smearing was set at 0.1 eV. The zero energy is at the conduction-band minimum of a clean surface of TiO2(110) and is corrected for the surfaces containing Pt10 or adsorbates such that the averaged electrostatic potential of a Ti atom located at the back side of the slab is aligned with that of the clean TiO2(110). The vertical broken line denotes the highest occupied state of each system. Reproduced with permission from Umezawa et al., J. Phys. Chem. C 120(17), 9160–9164 (2016). Copyright 2016 American Chemical Society.
In the presence of a Pt co-catalyst, the addition of MgO oxide onto the TiO2 surface was also found to highly enhance the efficiency of photoreduction of CO2 to CH4 with H2O [see Fig. 12(a)].45 A positive correlation between the activity of CH4 formation and the amount of adsorbed CO2 has been observed in the ternary nanocomposites containing TiO2, Pt, and a basic metal oxide [see Fig. 12(b)]. The interface between TiO2, Pt, and MgO plays a crucial role in the photocatalytic reaction. The addition of MgO was found to largely enhance the amount of CO2 adsorption on the catalyst surface. The adsorbed CO2 molecules are subsequently reduced to CH4 on the adjacent Pt nanoparticles by the photogenerated electrons enriched from TiO2.
(a) Proposed functioning mechanisms of the MgO layer and Pt nanoparticles over TiO2 for the photocatalytic reduction of CO2 in the presence of H2O. (b) Correlation between the amount of CH4 formed and the amount of CO2 chemisorbed on the Pt–TiO2 photocatalysts modified with various basic metal oxides. Reaction conditions: catalyst, 0.020 g; CO2 pressure, 2 MPa; H2O, 1.0 ml; and irradiation time, 10 h. Reproduced with permission from Xie et al., Chem. Commun. 49(24), 2451–2453 (2013). Copyright 2013 Royal Society of Chemistry.
(a) Proposed functioning mechanisms of the MgO layer and Pt nanoparticles over TiO2 for the photocatalytic reduction of CO2 in the presence of H2O. (b) Correlation between the amount of CH4 formed and the amount of CO2 chemisorbed on the Pt–TiO2 photocatalysts modified with various basic metal oxides. Reaction conditions: catalyst, 0.020 g; CO2 pressure, 2 MPa; H2O, 1.0 ml; and irradiation time, 10 h. Reproduced with permission from Xie et al., Chem. Commun. 49(24), 2451–2453 (2013). Copyright 2013 Royal Society of Chemistry.
One of the main problems associated with photocatalytic CO2 reduction in aqueous solution is the photoreduction of protons to hydrogens (2H+ + 2e− → H2). This can take place simultaneously and consequently competes with CO2 reduction.5,98 However, the reduction of CO2 is much more difficult than the reduction of protons. The reasons are shown in the following: (1) breaking the O=C=O double bonds is very difficult due to the high antibonding orbital of the CO2 molecule; (2) hydrogen formation from protons is a two-electron transfer process, whereas photocatalytic CO2 reduction to methanol and methane is six-electron and eight-electron transfer processes, respectively. In aqueous solution, competition between photoreduction of CO2 and protons exists. Designing and synthesizing photocatalysts with high selectivity for CO2 photoreduction and low selectivity for proton reduction are crucial. Neatu et al. co-deposited an Au and Cu alloy on TiO2 films and found that the photocatalytic efficiency and selectivity for CO2 photoreduction to methane by H2O increased dramatically,5 especially when the ratio of Au and Cu is 1:2. At this ratio, the formation rate of methane is 2.2 ± 0.3 mmol·g−1·h−1. It is substantially larger than that of H2 formation which was only 0.29 mmol·g−1·h−1. The selectivity of CH4 formation can be as high as 97%. This is a major increase when comparing to the formation rates of CH4 and H2 on the pure TiO2 surface, which were only 0.05 and 0.048 mmol g−1 h−1, respectively. The results demonstrate that an (Au, Cu) alloy is an extremely efficient material for CO2 photoreduction to CH4 with H2O by effectively inhibiting the competitive reaction of H2 formation. The high selectivity of photocatalytic production to CH4 is due to the presence of Cu bonding to CO on the photocatalyst, while the visible light photo-response would be introduced by not only the TiO2 photocatalyst but also the surface plasmon band of Au.
3. Z-scheme photocatalysts
Co-catalysts can improve the efficiency and selectivity of CO2 photoreduction. However, their limited ability to harvest visible-light and the rapid charge recombination still constitutes serious obstacles for these photocatalysts to efficiently generate solar fuels. Therefore, a solution to expand the light absorption of harvesters and enhance the separation and transfer of photogenerated charge carriers is very important. Instead of focusing on co-catalysts, some researchers have turned their attention to prepare photocatalysts consisting of different semiconductors, such as TiO2/ZnO,99 CdS/TiO2,100 13 g-C3N4/ZnO,101 and Ag3PO4/g-C3N4.102 The formed Z-scheme heterojunctions can considerably diminish the recombination of electron-hole pairs and can also utilize visible light more efficiently. As shown in the Z-scheme photocatalytic system (see Fig. 13), the reaction involves two steps of photoexcitation. First, the electrons in the VB of photocatalyst I (PC I) are excited into its CB under sunlight. Similarly, the electrons in the VB of PC II are excited into its CB. The photogenerated electrons in the CB of PC I are then transferred to the VB of photocatalyst II and recombined with holes (PC I). As a result, the photogenerated electrons of PC II are located in its CB and the holes of PC I are left in its VB. Finally, the photogenerated electrons in the CB of PC II proceed with the reduction of CO2, while the photogenerated holes in the VB of PC I are used to oxidize H2O. Thus, this kind of Z-scheme photocatalytic system illustrates a strong reducibility through PC II and a strong oxidizability by PC I. The Z-scheme system can be composed by two photocatalysts with narrow bandgaps and this is beneficial to enhance the absorption efficiency of visible light.
He et al. designed an Ag3PO4/g-C3N4 composite photocatalyst (see Fig. 14) and proposed that the addition of Ag3PO4 increased the photo-absorption performance of g-C3N4 by using diffuse reflectance spectroscopy.102 They also found that this kind of Z-scheme structure highly promotes the separation of electron-hole pairs and enhances the efficiency of photocatalytic CO2 reduction. Under simulated sunlight irradiation, the optimal Ag3PO4/g-C3N4 photocatalyst can generate a CO2 conversion rate of 57.5 μmol h−1 gcat−1, which is 6.1 and 10.4 times higher than those of g-C3N4 and P25, respectively. They also found the formation of Ag nanoparticles in the composite and proposed that they act as a charge transmission bridge in this system. The photogenerated electrons in the CB of Ag3PO4 migrate to metallic Ag, where they combine with the holes transferred from the VB of g-C3N4. This is because the conduction band minimum of Ag3PO4 is more negative than the Fermi level of Ag nanoparticles. This kind of charge transmission effectively improves the separation of electron-hole pairs. This enables the trapping of electrons in the CB of g-C3N4 and creation of holes in the VB of Ag3PO4. Furthermore, the negative ECB of g-C3N4 also results in strong reducibility and facilitates the reduction of CO2 to hydrocarbons. Based on the g-C3N4/ZnO system, Yu et al. also found enhanced photocatalytic CO2 reduction when the system follows a direct-scheme mechanism.101 This kind of photocatalytic system exhibited a 2.3-fold increase in the photocatalytic activity for CO2 reduction when compared to the pure g-C3N4 system, and the original selectivity of pure g-C3N4 to convert CO2 directly into CH3OH was maintained. Therefore, the Z-scheme shows an excellent activity for CO2 photoreduction and highly improves the efficiency of reactions.
Photocatalytic mechanism scheme of the Ag3PO4/g-C3N4 composite. Adapted with permission from He et al., Environ. Sci. Technol. 49(1), 649–656 (2015). Copyright 2015 American Chemical Society.
Photocatalytic mechanism scheme of the Ag3PO4/g-C3N4 composite. Adapted with permission from He et al., Environ. Sci. Technol. 49(1), 649–656 (2015). Copyright 2015 American Chemical Society.
IV. CONCLUSION AND OUTLOOK
This work has summarized the present status of the photocatalytic reduction of CO2 to fuels in heterogeneous catalysis from both experimental and theoretical perspectives. Although significant achievements have been made toward the conversion of CO2, research is still far from achieving artificial photosynthesis. This is due to the complex photocatalytic mechanism and low efficiency and selectivity of products. A more comprehensive understanding of the reaction mechanism and modification of photocatalysts is required. Because of the inert character of the CO2 molecule, the adsorption of CO2 on the surface of catalysts is a key step to activate CO2 and determines the subsequent photoreduction mechanism. The advancements in DFT have enabled the adsorption and activation of the CO2 molecule to be widely investigated at the atomic level. The activation of CO2 is from a linear physisorbed configuration to a bent chemical adsorbed CO2⋅− configuration by coupling with an electron. Surface defects such as oxygen vacancies formed during the preparation and fabrication of photocatalysts can effectively promote the adsorption and activation of CO2. Photocatalytic CO2 reduction generally proceeds with aqueous dispersion. Results prove that the solvation effect has considerable influence on CO2 photoreduction. It even changes the most stable adsorption configuration of CO2. Therefore, the influence of surface defects and solvation effect is pivotal to reveal and improve the understanding of CO2 photoreduction.
The photoreduction of CO2 with H2O involves as many as eight electron and six proton transfers resulting in the formation of different products such as CO, HCOOH, HCHO, CH3OH, and CH4. This makes the mechanism of CO2 photoreduction extremely complicated. In the gas phase, a concerted two-electron process for CO2 photoreduction with adsorbed protons was found on the pure TiO2 surface and the formation of CH4 was achieved via a CO intermediate. However, in the actual aqueous environment, protons widely dissolve in the solution. It is still unknown how the protons participate in the hydrogenation reaction. Will they adsorb on the surface first and then react with CO2, or directly attack CO2? Therefore, an in-depth understanding elucidating the role of protons in aqueous CO2 conversion is much desired.
Furthermore, designing of proper photocatalysts is another important issue as it determines the activity and selectivity of CO2 photoreduction. The early research in this field employed suspensions of simple semiconductor materials such as pure TiO2, which exhibit low efficiency. Except the nanostructuring techniques leading the increasing surface area and active reaction sites, recent developments have also focused on various composite photocatalysts by introducing metal and metal oxide co-catalysts, as well as Z-scheme systems. The use of composite photocatalysts greatly improves the charge separation since the photogenerated electrons and holes can be trapped efficiently and effectively. These kinds of composite photocatalysts also provide active sites for activation and reaction of the adsorbates. For example, in Au-Cu alloys loaded on TiO2 photocatalysts, the role of Cu is to promote CO2 adsorption and activation, while Au inclines to promote the hydrogenation reaction and act as a light harvester due to a surface plasmon band in the visible light region. Additionally, Z-scheme photocatalytic systems have shown strong reducibility and oxidizability, whilst their narrow bandgap is extremely beneficial to enhance the adsorption efficiency of visible light. Although various photocatalysts have shown good performance, the photocatalytic reduction of protons to hydrogens, a competitive reaction in aqueous CO2 photoreduction, remains a major obstacle for the selectivity of the reaction. Therefore, the design of ideal photocatalysts to enhance the photocatalytic reduction of CO2 and suppress H2 formation need to be explored further.
ACKNOWLEDGMENTS
This project was supported by the National Key Basic Research Program of China (Grant No. 2013CB933201), the National Natural Science Foundation of China (Grant Nos. 21333003, 21303052, and 21622305), Shanghai Rising-Star Program (No. 14QA1401100), “Chen Guang” Project (No. 13CG24), Young Elite Scientist Sponsorship Program by CAST, and Fundamental Research Funds for the Central Universities.