DOI:
10.1039/D5RA03826J
(Paper)
RSC Adv., 2025,
15, 23448-23460
Determining the structure and properties of CO2 reduction photocatalysts: single atom cobalt atoms supported on various carbon nitrides†
Received
30th May 2025
, Accepted 26th June 2025
First published on 7th July 2025
Abstract
Carbon nitride materials paired with Co atoms have been shown to be effective for CO2 photoreduction. However, the structures of Co/carbon nitrides are uncertain, especially when the degree of polymerization of the carbon nitrides is unknown. Literature has focused predominantly on the fully condensed carbon nitride, g-C3N4, despite evidence of other carbon nitrides being present in carbon nitride catalysts. We therefore used density functional theory (DFT) to model Co binding with molecular, partially condensed polymeric, and fully condensed polymeric carbon nitrides. We found strong binding of Co to the carbon nitrides, and larger coordination tended to lead to stronger binding of Co, as well as more cationic Co atoms. Co had a significant effect on photocatalytic processes. Co lowered band gaps significantly, enabling greater photoexcitation yields. Activation of CO2 into a bent, anionic state is an important initial step during CO2 reduction. Our calculations show that all Co/carbon nitride systems, except two-layer carbon nitrides, activated CO2. For bent CO2, strong interactions occurred between Co/carbon nitride systems, as evidenced by density of states plots and calculated interaction energies. On the other hand, weak interactions occurred with adsorbed linear CO2. Activation energies of CO2 spanned a wide range of values (−2.33 to −0.38 eV), and intermediate values over partially condensed carbon nitrides may be more likely to enable CO2 reduction. Our work provides insights and understanding of Co/carbon nitride catalysts, and motivates study of carbon nitrides beyond g-C3N4 as photocatalysts.
1 Introduction
Utilizing sunlight with efficient, low-cost photocatalysts is a vital step towards a clean, sustainable future. Carbon nitride materials have attracted much interest as photocatalysts,1–9 in part because these materials consist of cheap, abundant elements. They also have moderate band gaps for photoexcitation. These catalysts have been studied for a number of reactions, including water splitting,10–14 water purification,15–17 and CO2 reduction.18–23 Reducing atmospheric CO2 is an essential goal for mitigating climate change.
Carbon nitrides can have various compositions and structures.24–28 One of the most studied carbon nitrides is graphitic C3N4, or simply g-C3N4, which has a two-dimensional graphene-like structure with interconnected heptazine units. g-C3N4 may be considered a fully condensed two-dimensional material whereby melem molecules (C6N10H6) combine together to release NH3.28–30 Both experimental and computational work have shown that various polymeric carbon nitrides form due to incomplete condensation under different reaction conditions.13,28,31–36 Density functional theory (DFT) work suggests that g-C3N4 may be hard to synthesize under typical synthesis conditions (e.g., atmospheric pressure), and that partially condensed polymeric carbon nitride structures (e.g., melon – [C6N9H3]n) may be more prevalent than many researchers have considered.31–33,35,37 Furthermore, most experimental carbon nitrides contain a significant amount of hydrogen25,28,35 and also deviate from the ideal g-C3N4 C
:
N ratio of 0.75, 24,34 again indicating the likely presence of partially condensed polymers in many studies. Indeed recent DFT studies have attempted to more fully understand polymeric carbon nitrides such as melon.31,38–40 Other DFT work33 suggests that polymeric carbon nitrides may have many different structural domains (both partially and fully condensed) rather than a single structure. Thus, there is a need to better understand and characterize polymeric carbon nitrides beyond just two-dimensional g-C3N4, particularly partially condensed carbon nitrides. In this work we concentrate on characterizing polymeric carbon nitride materials as photocatalysts for CO2 reduction, especially when paired with single metal atoms.
One promising way to increase photocatalytic activity of carbon nitrides is to pair them with single metal atoms.41–48 Single atom catalysts (SACs) have attracted much interest because they can have high catalytic activity per atom, and can also have novel chemical or catalytic properties compared to bulk materials. The carbon nitride acts as a support and photoabsorber for these SACs. Synergy between the metal atoms and carbon nitride can decrease band gaps/increase photoabsorption,49–55 and also increase charge separation/lower charge recombination (necessary for long-lived charge carriers).42,46,55–59 Such carbon nitride/single metal atom catalysts may be active for reactions such as the H2 evolution reaction52,58,60,61 or CO2 reduction.62–66 Of particular interest is Co, which has demonstrated activity for CO2 photoreduction.67–74 Using Co/carbon nitride catalysts for CO2 photoreduction is a promising way to deal with the CO2 problem. Co/carbon nitride catalysts have demonstrated better adsorption of CO2, higher light utilization, and higher production of CO compared to pure carbon nitrides.72
Identifying the geometry of Co/carbon nitrides is an important step towards connecting their structure to catalytic activity, as well as a precursor for simulating these materials using atomistic modeling methods. Without details on their structure, detailed reaction mechanisms and structure–activity relationships are unclear. Indeed, there is ambiguity about the structure of Co/carbon nitrides. For example, the Co–N coordination number of Co/carbon nitride catalysts has been reported to be in the range of two to four depending on the material.61,68,71–76 We have shown through theory and experiment a potential four-coordinated structure.74 In contrast, Co may be four- or six-coordinated in bulk cobalt oxides. Other work involving doped carbon nitrides reported four-coordinated Co–P structures.77 Besides details on the geometry of Co atoms, identifying the oxidation state, effect on photoabsorption, and potential reaction sites are important for clarifying structure–activity relationships. The modeling literature has predominantly focused on how Co SACs bind to g-C3N4.78–81 A few DFT papers have been published on Co SACs supported by melon, but these papers assumed a dehydrogenated form of melon,82,83 while another paper modeled SACs to one-layer melon.84 Accordingly, it is critical to understand the nature of Co interacting with a full range of carbon nitrides, including molecular, partially condensed polymers, and fully condensed polymers, as well as multi-layer carbon nitrides, as all these types of carbon nitrides may be present in real-world catalysts.
In this work, we used DFT to model Co/carbon nitride catalysts in order to clarify their structures and determine potential photocatalytic activity, especially for CO2 photoreduction. We modeled molecular carbon nitrides, partially condensed carbon nitrides, and fully condensed two-dimensional carbon nitrides. We assessed how different carbon nitride supports interact with Co atoms, leading to different complexes and Co oxidation states. We also elucidated how Co atoms may affect band gaps of these materials, which is important for photocatalysis. Finally we modeled the activation of CO2 over these materials, a necessary first step towards CO2 photoreduction. Our work provides fundamental details on these catalysts, and ties together how different carbon nitride supports affect SAC stability, structure, and photocatalytic activity.
2 Methodology
2.1 Simulation parameters
We used the Vienna Ab initio Simulation Package (VASP)85–88 with the Perdew Burke Ernzerhof exchange-correlation functional89 for all the DFT simulations. Valence electrons were represented by a plane wave basis set, while core electrons were represented by projector augmented wave (PAW) potentials.90,91 The number of valence electrons for each atom were: Co (9), C (4), N (5), H (1), and O (6). A cutoff energy of 450 eV was used for the plane waves. Criterion for convergence of electronic and ionic relaxations were 1 × 10−5 eV and 0.02 eV Å−1. The Grimme D3 correction92 with Becke–Johnson damping93 was used to incorporate van der Waals interactions. We also used Gaussian smearing with a σ value of 0.1 eV. Charges of atoms were determined using Bader charge analysis.94–96 We used the HSE06 hybrid functional,97–99 combined with Gaussian smearing and a σ value of 0.01 eV, for calculating higher resolution band gaps and the density of states. The VASPKIT code was used for post-processing of our simulations.100 Images of structures were generated using VESTA.101
Binding energies (ΔEbinding) of Co atoms were calculated by eqn (1):
|
ΔEbinding = ECo/Carbon Nitride − ECarbon Nitride − ECo
| (1) |
Here
ECo/Carbon Nitride is the energy of the Co/carbon nitride complex,
ECarbon Nitride is the energy of the carbon nitride material, and
ECo is the energy of a lone Co atom. Adsorption energies (Δ
Eads) of CO
2 were calculated by
eqn (2):
|
ΔEads = ECO2/Co/Carbon Nitride − ECo/Carbon Nitride − ECO2
| (2) |
Here
ECO2/Co/Carbon Nitride is the energy of the CO
2/Co/carbon nitride complex and
ECO2 is the energy of a CO
2 molecule.
2.2 Structural models
In this work we modeled several carbon nitrides, including polymeric building blocks (melem and melem dimer), partially condensed polymers (melon), and fully condensed carbon nitrides (g-C3N4). Fig. 1 shows the carbon nitrides we modeled. These carbon nitrides have been previously modeled.31,32,40 Interestingly, recent simulations suggest that experimentally relevant carbon nitrides consist of both partially and fully condensed polymer structures.33 Melem is a building block of polymeric carbon nitrides.24,25,102 Melem dimers may also form during the polymerization process. Melem and melem dimers were modeled in large supercells of size 30 Å × 30 Å × 30 Å. We used a single k-point for these calculations. Condensation of melem can form melon.24,33,102 Our melon model was based on Fina et al.37 and is further discussed in the ESI Section 1.† As Fig. 1 shows, the stable melon structure consisted of melon strands arranged together within the simulation cell. The single-layer melon structure had lattice lengths of 12.8 Å and 16.8 Å (based on lattice optimization) with a vacuum spacing of ∼30 Å, and included 72 atoms. Our two-layer melon structure was also based on Fina et al.’s melon structure.37,40 This simulation cell had lattice vectors of 12.8 Å and 16.7 Å after lattice optimization. The z lattice vector was set to ∼30 Å, resulting in a vacuum spacing of ∼27 Å.
 |
| Fig. 1 Carbon nitrides molecules, melem (a) and melem dimer (b), modeled in this work. Periodic carbon nitride polymers , melon (c and d) and graphitic C3N4 (e and f), modeled in this work. Also shown are the simulation cells for the periodic structures. Brown spheres represent carbon atoms, blue spheres represent nitrogen atoms, and pink spheres represent hydrogen atoms. | |
Single-layer g-C3N4 (similar to Melissen et al.31) was modeled using a (2 × 2) orthorhombic supercell containing 56 atoms and with optimized lattice lengths of 11.8 Å and 13.1 Å. The vacuum space was ∼30 Å. Corrugated g-C3N4 is more stable than planar g-C3N4, as found in literature31 and our own calculations, so we modeled all g-C3N4 as corrugated. We also modeled two-layer g-C3N4, as layers of g-C3N4 may be attracted by van der Waal forces. Two-layer g-C3N4 (taken from Botari et al.32) was modeled with an orthorhombic cell containing 112 atoms with lattice parameters being 11.9 Å and 13.3 Å. The vacuum space was ∼26 Å. Test calculations (see Table S1†) indicated that a (2 × 2 × 1) k-point mesh was sufficient for all the periodic polymer calculations (i.e., melon and g-C3N4), and we used such a k-point mesh in our work for all the periodic carbon nitrides. Analysis of our carbon nitride models (see the ESI Section 2†) indicates agreement with literature.
A challenge is quickly identifying whether Co–C or Co–N bonds occur in a structure, or the coordination number of Co. We used a simple procedure of designating a bond if the Co–X (X = C, N) bond distance was less than some threshold cutoff value, similar to previous work.103,104 Co–X bond distances in Co/carbon–nitrogen materials (such g-C3N4 and nitrogen doped graphene) were found to range from 2.02 to 2.44 Å (Co–C) and 2.04 to 2.37 Å (Co–N) (from both experimental61,75,105–108 and computational work61,75,105). Therefore, we set the cutoff value for determining Co–C and Co–N bonds to 2.5 Å.
3 Results and discussion
3.1 Binding of Co to carbon nitride materials
3.1.1 Binding of Co to molecular carbon nitrides. We considered how single Co atoms bind to the different carbon nitrides, first examining Co bound to carbon nitride molecules. In order to identify the most stable structures, we modeled several possible Co/molecular complexes. For example, we modeled Co binding to two adjacent N atoms within a single carbon nitride molecule, as well as combinations in which N atoms from two carbon nitride molecules interacted with the Co atom. Furthermore, the Co atom was placed in different orientations, such as the same plane with the carbon nitrides or sandwiched between two molecules. We present those structures which are most stable after geometry optimization, or lowest in energy, in Fig. 2. Table 1 also summarizes details on these structures. We note that several structures had quite different geometries, but similar binding energies for Co. Thus, we present several different structures for each molecule, and label them based on the number of carbon nitride molecules and their configuration. A simplified representation of each structure showing only the local coordination around the Co atom is provided in Fig. S3.† For example, Co/melem-1 refers to the first structure with Co binding with one melem, while Co/melem-2 refers to the second structure with Co binding with one melem. Co/2-melem-1 refers to the first configuration of Co binding with two melem molecules. Co/melem-dimer-1 refers to the first configuration with Co binding to a melem dimer, while Co/2-melem-dimer-1 refers to first configuration with Co binding to two melem dimers. In addition to the structures presented, we also identified other less stable structures, as well as structures that were similar to those in the main text. These alternative structures are included in Fig. S1.†
 |
| Fig. 2 Stable structures where Co was bound to carbon nitride molecules. Brown spheres represent carbon atoms, blue spheres represent nitrogen atoms, pink spheres represent hydrogen atoms, and green spheres represent cobalt atoms. The notation system is decribed in the main text. When two molecules are present, these are designated by the prefix ’2-’. Each unique configuration is identified by the final number (e.g., ’-1′, or ’-2′). A simplified representation of each structure showing only the local coordination around the Co atom is provided in Fig. S3.† | |
Table 1 Results for Co bound to molecular carbon nitrides. Shown are calculated Co Bader charges, Co binding energies, Co coordination numbers (Co–N coordination number, followed by Co–C coordination number), and distances between Co and closest atoms (typically N). In some cases Co was bound to C atoms, as indicated in the table. Structures in the table correspond to the same structures shown in Fig. 2. Simplified images of the local structures are found in Fig. S3
Models |
Co charge |e−| |
ΔEbinding−Co (eV) |
Co CN |
Co–M distance (Å) |
Co/melem-1 |
0.10 |
−1.11 |
1 |
2.07 |
Co/melem-2 |
0.15 |
−1.12 |
2 |
2.16 2.16 |
Co/2-melem-1 |
0.34 |
−3.14 |
2 |
1.86 1.86 |
Co/2-melem-2 |
0.68 |
−3.36 |
4/2 |
2.02 2.17 2.15 2.02 1.90(C) 1.90(C) |
Co/2-melem-3 |
0.66 |
−3.31 |
4/2 |
2.03 2.10 2.00 2.10 2.45(C) 1.88(C) |
Co/2-melem-4 |
0.70 |
−3.15 |
4/2 |
2.20 2.02 2.04 2.16 1.90(C) 1.90(C) |
Co/melem-dimer |
0.55 |
−2.78 |
2 |
1.86 1.86 |
Co/2-melem-dimer-1 |
0.96 |
−5.23 |
4 |
1.96 2.01 1.96 2.01 |
Co/2-melem-dimer-2 |
0.77 |
−5.28 |
4 |
1.91 1.91 1.91 1.91 |
Binding energies for all molecular structures were found to be exothermic, ranging from −5.28 to −1.11 eV. The weakest binding occurred with just single melem molecules (e.g., −1.11 eV for Co/melem-1), where the Co interacted with just one or two N atoms. Notably, binding energies were more exothermic when Co interacted with two molecules (e.g., −3.36 eV for Co/2-melem-2), such as with two melem molecules or with two melem dimers. As noted above, Co binding with single melem molecules resulted in either one or two Co–N coordination. Co binding with other molecular carbon nitrides (such as multiple melems or dimers) typically resulted in larger coordination numbers (up to 6), as well as more exothermic binding energies (ranging from −5.28 to −2.78 eV). Notably, Co binding with two melems tended to form Co–C moieties, whereas no other carbon nitride molecules had such bonds.
Fig. S2† shows a comparison between binding energies and Co coordination number. R2 values for a regression fit are around 0.5, indicating reasonably moderate correlation. Our results show that Co with higher coordination was generably most stable, as seen for example by Co/2 melems or Co/2 melem dimers, where the most stable structures have Co with four Co–N coordination (Co/2-melem-2 and Co/2-melem-dimer-2). Coordination values beyond four tend to have a small affect on binding energies (see Fig. S2†). We also observed a variety of geometries for the Co/melem complexes. Simplified representations indicating the local geometries and coordination coordination around the Co atom are provided in Fig. S3.† The local geometries are also listed in Table S4.† Notably, the bent and linear geometries had the weakest binding energies (−1.12 eV for Co/melem-2, −2.78 eV for Co/melem-dimer, and −3.14 eV for Co/2-melem-1), while the strained trigonal prismatic, tetrahedral, and square planar geometries had the strongest binding energies, ranging from −5.28 to −3.15 eV, likely due to the increased number of Co–X bonds in these later structures.
3.1.2 Binding of Co to partially condensed polymeric carbon nitrides. Fig. 3 shows structures for Co bound to a partially condensed polymeric carbon nitride, melon. Details of these calculations are summarized in Table 2. We present several different structures for Co binding to single-layer melon (denoted as Co/melon) and two-layer melon (denoted as Co/2-melon), with each configuration designated by a different number (e.g., 1, 2, etc.). The most stable structure with Co binding to single-layer melon was Co/melon-1 (Ebinding = −3.00 eV), where Co was coordinated to two N atoms. An alternative structure (Co/melon-2) had a binding energy of −2.84 eV, and Co was coordinated to three N atoms. Simplified images of Co/melon structures illustrating the local geometries are found in Fig. S4.† In Co/melon-1, Co was bound to nitrogen atoms from two melon strands, forming a linear structure. Co/melon-2 formed a trigonal planar structure interacting with two strands of melon.
 |
| Fig. 3 Stable structures where Co was bound to melon strands. Brown spheres represent carbon atoms, blue spheres represent nitrogen atoms, pink spheres represent hydrogen atoms, and green spheres represent cobalt atoms. When two melons are present, these are designated by the prefix ’2-’. Each unique configuration is identified by the final number (e.g., ’-1′, ’-2′, etc.). Top and side views are shown. Simplified images of the Co/melon structures are found in Fig. S4.† | |
Table 2 Results for Co bound to carbon nitride polymers. Shown are calculated Co Bader charges, Co binding energies, Co coordination numbers (Co–N coordination number, followed by Co–C coordination number), and distances between Co and closest atoms (typically N). In some cases Co was bound to C atoms, as indicated in the table. Structures in the table correspond to the same structures shown in Fig. 3 and 4. Simplified images of the local structures are found in Fig. S4
Models |
Co charge |e−| |
ΔEbinding−Co (eV) |
Co CN |
Co–M distance (Å) |
Co/melon-1 |
0.49 |
−3.00 |
2 |
1.83 1.83 |
Co/melon-2 |
0.72 |
−2.84 |
3 |
2.06 1.90 1.92 |
Co/2-melon-1 |
0.78 |
−3.36 |
4/1 |
1.96 2.30 1.94 2.01 1.99(C) |
Co/2-melon-2 |
0.77 |
−3.61 |
3/1 |
1.99 1.97 1.93 1.94(C) |
Co/2-melon-3 |
0.78 |
−3.46 |
4/1 |
1.94 2.02 2.36 1.98 1.96(C) |
Co/2-melon-4 |
0.86 |
−3.35 |
4 |
2.02 2.16 2.01 2.06 |
Co/C3N4-1 |
0.71 |
−2.88 |
2/1 |
1.87 1.89 1.93(C) |
Co/C3N4-2 |
0.75 |
−3.06 |
3/1 |
1.95 2.06 1.95 1.96(C) |
Co/2-C3N4-1 |
0.80 |
−3.84 |
4/2 |
2.08 1.96 2.05 2.02 1.95(C) 1.95(C) |
Co/2-C3N4-2 |
0.77 |
−3.92 |
3/3 |
1.94 1.91 1.98 2.10(C) 2.06(C) 2.48(C) |
Co/2-C3N4-3 |
0.77 |
−3.87 |
3/2 |
1.96 1.95 1.90 1.96(C) 2.09(C) |
Co/2-C3N4-4 |
0.82 |
−3.66 |
3/1 |
2.04 1.92 1.97 1.97(C) |
The binding energies of Co interacting with two melon layers ranged from −3.61 to −3.35 eV. Co is therefore more stable interacting with two melon layers compared to just one melon layer, although the binding energies were not drastically more exothermic (−3.61 eV for for Co/2-melon-2 and -3.00 eV for Co/melon-1). Co binding with two-layer melon formed either three or four Co–N bonds with both the layers, and up to two Co–C bonds. The most stable configuration was Co/2-melon-2 (Ebinding = −3.61 eV), where the Co was coordinated to three N atoms and one C atom, having a square planar structure. The local geometries are listed in Table S4† and shown in Fig. S4,† indicating a variety of Co complexes, including linear, trigonal planar, square pyramidal, square planar, and tetrahedral. We also observed strained square pyramidal (Co/2-melon-1 and Co/2-melon-3, both with four Co–N and one Co–C bonds), and tetrahedral (Co/2-melon-4 with four Co–N bonds) geometries. Our calculations show that Co binding to two-layer melon tended to form Co–C bonds, whereas Co binding to single layer melon did not form any Co–C bonds.
3.1.3 Binding of Co to fully condensed polymeric carbon nitrides. Following the modeling of Co/melon structures, we explored Co binding with fully polymerized carbon nitride g-C3N4. Fig. 4 shows results for Co bound to polymeric (periodic) g-C3N4, and details of these calculations are summarized in Table 2. We found two stable Co/C3N4 structures. Co/C3N4-1 had a binding energy of −2.88 eV. Co/C3N4-2 was the most stable g-C3N4 structure identified (Ebinding = −3.06 eV). We also modeled Co binding with two-layer g-C3N4, and the binding energies were between −3.92 and −3.66 eV. The coordination numbers were between three and four for single-layer C3N4, and between four and six for two-layer C3N4. Notably, all the stable geometries with Co bound to C3N4 had Co–C bonds (one to three Co–C bonds). It is also worth mentioning that Co binding with 2 molecular carbon nitrides is about two times more stable than Co binding to just one molecule; however, Co binding with 2-layer C3N4 is only ∼ 30% more stable than Co binding with 1-layer C3N4. For example, there is not a drastic change in coordination for Co/C3N4-2 (four coordination with 3 Co–N bonds) compared to Co/2-C3N4-2 (six coordination with 3 Co–N bonds). Coordination geometries of all these structures are listed in Table S4, and shown in Fig. S4.†
 |
| Fig. 4 Stable structures where Co was bound to g-C3N4. Brown spheres represent carbon atoms, blue spheres represent nitrogen atoms, pink spheres represent hydrogen atoms, and green spheres represent cobalt atoms. When two g-C3N4 sheets are present, these are designated by the prefix ’2-’. Each unique configuration is identified by the final number (e.g., ’-1′, ’-2′, etc.). Top and side views are shown. Simplified images of Co/C3N4 structures are found in Fig. S4.† | |
3.1.4 Connection to experimental work. Some experimental work has suggested four-coordinated bonding with N (Co–N4) when Co interacts with polymeric carbon nitrides.61,107,109 We observed four-coordinated Co in Co/C3N4-2, where 3 Co–N and 1 Co–C bonds formed. Fig. S1† shows that Co/2-melem-6 forms a Co–N4 arrangement, although this structure was less stable than other melem structures. In our previous work,74 we identified a similar possible Co–N4 geometry, although this involved Co binding with two heptazine molecules (chosen to mimic carbon nitride edges), rather than melem, melon, or g-C3N4. It should be noted that experimental work, such as Extended X-ray Absorption Fine Structure (EXAFS), may have trouble distinguishing between Co–N and Co–C bonds.110–113 Thus experimental work from literature may be observing both Co–N and Co–C bonds (such as in Co/C3N4-2, or Co/2-melon-2) or Co interacting with carbon nitride edges (such as in our previous work74). There may also be fitting errors in methods such as EXAFS due to low metal concentration.113 There is some literature that suggests the formation of M–C bonds. Fao et al.'s computation work114 reported Co coordinated to three N atoms and one C atom with a single-layer corrugated g-C3N4. Zheng et al.'s computation work75 found Co binding with corrugated g-C3N4 to form three Co–N bonds and two Co–C bonds, although no bond lengths were reported. Wang et al.'s work115 combined experimental results, EXAFS and X-ray photoelectron spectroscopy (XPS), with DFT results to show Cu binding to g-C3N4 forming one Cu–C and two Cu–N bonds.Identification of the exact carbon nitride structure from experiment is not trivial, and it is possible that some literature work has misidentified their carbon nitride structure. See the introduction for discussion on partially condensed carbon nitrides and the difficulty of synthesizing g-C3N4. For example, temperature plays a vital role in what carbon nitride structure is synthesized and observed.40 It is possible that experimental papers finding four-coordinated Co atoms could be Co binding to partially condensed carbon nitrides. For example, Huang et al.'s experimental work116 argued that Co formed four identical Co–N bonds when incorporating Co into a carbon nitride. Their experimental work identified a significant number of NH2 groups in the carbon nitride, which suggests a structure other than fully condensed g-C3N4. This configuration resembles Co/2-melem-dimer-2 found in our study, as our structure also had four identical Co–N bonds. Ding et al.'s work109 presented a symmetric Co–N4 structure with 1.98 Å, similar to Co/2-melem-dimer-1, which has four Co–N bonds all similar in length. However, in Xiong et al.'s work,117 their experimental results indicted a Co coordination of 2.3, with lengths of 1.95 Å, similar to several of our modeled carbon nitrides: Co/melem-2, Co/2-melem-1, or Co/melem-dimer.
3.2 Electronic properties of Co-carbon nitride materials
3.2.1 Bader charge analysis. We further examined the electronic properties of the Co/carbon nitride structures. Tables 1, 2, and S4† provide Co Bader charges of different structures, which indicate the oxidation state of the Co atoms. Co became positively charged in all cases once it interacted with the carbon nitride supports. This agrees with experimental61,67,118 and computational75,119 observations on the cationic nature of Co. Larger Co Bader charge indicated more oxidation of Co, and presumably more chemical interactions between the Co and carbon nitride substrate. The Co charges in Co/molecular carbon nitrides ranged from +0.1 to +0.96 |e−|, with an average value of +0.55 |e−|. With polymeric carbon nitrides, Co charges were similar for single-layer and two-layer structures, ranging from +0.72 to +0.86 |e−|, except for Co/melon-1, which had a Co charge of + 0.49 |e−|. The average Co charge with the polymeric carbon nitrides was +0.75 |e−|. Thus, the Co atoms were all cationic when bound to carbon nitride molecules, and Co atoms bound to polymeric carbon nitrides were generally more cationic than Co bound to molecular carbon nitrides.We next determined the relationships between the Co charges and structure of the complexes. Fig. 5 illustrates how the charges correlate to Co coordination number, binding energies, and structure of the carbon nitride support. The plots (a) and (b) show general trends that the Co charge increases (i.e., becomes more oxidized) with both increasing Co–N coordination number and total Co coordination number. This connects with our observation that the Co charge was larger when Co was bound to more molecules (e.g., two melem molecules compared to one molecule), indicating that more chemical interactions (as evidenced through increased coordination number) lead to stronger charge transfer between Co and neighboring atoms. Indeed, Fig. S2† demonstrates how coordination number correlates to binding energy. Plot (c) shows that the Co charge increases with stronger binding (i.e., more negative binding energy). These results again indicate that stronger interactions between the Co atom and nearby bonding atoms lead to more charge transfer. Finally, we show in plot (d) how the size of the carbon nitride substrate correlates to Co charge. Larger carbon nitrides have more triazine rings, and generally bind stronger to Co (see Table 2), leading to more oxidation of the Co atom. Our results connecting the Co charge to structure are especially important in deciphering experimental data from SACs supported on carbon nitrides. For example, these trends could help interpret coordination numbers and/or oxidation states obtained through methods, such as from XAS.120 Indeed, the oxidation state of Co has been linked to its reactivity.67
 |
| Fig. 5 Correlations between the Bader charges of Co atoms and (a) Co–N coordination numbers, (b) total Co coordination numbers (both C and N), (c) Co binding energies, and (d) size of the carbon nitride (i.e., number of triazine rings). The red dashed lines are second-order polynomial fits. | |
3.2.2 Band gaps. A key property for photocatalysis is the band gap, as a lower band gaps enables a broader spectrum of light to initiate photoexcitation. Co atoms may create favorable reaction sites (as discussed in Section 3.3), but also facilitate increased photoexcitation in Co/carbon nitride photocatalysts. Experimental work shows Co supported on carbon nitrides leads to a reduction in band gaps.,118,121–123 although the exact structures of the carbon nitrides in several of these papers still need clarification. The reduction of the band gap caused by binding of a Co atom to carbon nitrides has also been demonstrated in computational work involving g-C3N4.78,79 On the other hand, computational work on the effect of Co paired with other carbon nitrides (such as melem or melon) is missing. Therefore, we calculated band gaps of Co interacting with a variety of carbon nitrides.We show the band gaps of pristine carbon nitrides and the most stable Co/carbon nitrides calculated using the HSE06 hybrid functional in Fig. 6. Our results demonstrate that the band gap decreases with increased polymerization of the carbon nitrides (see left panel of Fig. 6), as already suggested in literature.36 Notably, the incorporation of a Co atom significantly reduces the band gaps (see right panel of Fig. 6). For example, the band gap of Co/2-melem-2 is 2.61 eV, which is 2 eV lower than a pristine melem molecule. Other Co/carbon nitride band gaps are much smaller, being in the range of 0.15 to 1.49 eV. There does not appear to be a clear relationship between carbon nitride size and Co/carbon nitride band gap. All Co/carbon nitrides have band gaps < 1.2 eV, with the exception of Co/2-melem-2 (2.61 eV) and Co/2-C3N4-2 (1.49 eV). However, our results do show that Co does indeed cause band gap reduction for Co supported on a variety of carbon nitrides.
 |
| Fig. 6 Band gaps of pristine carbon nitrides and the most stable Co/carbon nitrides calculated using the HSE06 functional. The left graph shows the band gaps for pristine carbon nitrides, while the right graph shows the band gaps for the most stable Co/carbon nitrides. The arrow shows the trends in band gaps for carbon nitrides, which decrease with increasing polymerization. | |
3.2.3 Electronic states. In the following section we briefly analyze the electronic states to better understand how Co leads to diminished band gaps and better photocatalysis. We show the density of states (DOS) using the HSE06 functional of carbon nitride molecules in Fig. S6 and polymers in S7.† The valence bands of the bare carbon nitrides consist mainly of states from nitrogen, and the conduction bands consist of states from carbon and nitrogen. Upon addition of Co, however, states arise in the carbon nitride band gaps consisting primarily of Co electronic states. These states effectively lower the band gaps as lower energies are needed to excite electrons from these gap states to the conduction bands. Previous literature indicated that similar gap states could arise over g-C3N4,124 and a melon-like structures.82 Our work shows that these gap states arise, regardless of the degree of polymerization (i.e., molecular, partially condensed, and fully condensed carbon nitrides). Experiment also indicates that Co lowers electron–hole recombination when added to carbon nitrides.123 This lowered recombination could be facilitated by the midgap states created by Co in all carbon nitrides.
3.3 CO2 activation over Co–carbon nitride materials
CO2 photoreduction involves many different steps, and a key step is the activation of CO2. The formation of bent, anionic CO2 is important, as this enables the further reactivity of this otherwise stable linear molecule.125 If formation of bent, adsorbed CO2 is unfavorable, then CO2 reduction will be slow or hindered. We note that CO2 adsorption is weak on pure carbon nitrides, as the process has been reported for single-layer g-C3N4 to be slightly exothermic (−0.42 eV)126 or endothermic (0.02 eV,78 and 0.24 eV).72 Thus, improving the adsorption and reactivity of carbon nitrides is needed, such as pairing them with SACs. Experimental work on Co/carbon nitride catalysts shows these to be potential catalysts for CO2 reduction.67,68,71–74 In this section we examine the formation of activated CO2 over the various Co/carbon nitride structures.
3.3.1 Formation of bent CO2. We modeled both linear and bent CO2 adsorbed on our most stable Co/CN structures. Fig. 7 presents only the most stable configurations (i.e., lowest CO2 adsorption energies) for adsorbed CO2, while Fig. S10† shows all the CO2 structures we identified. Table 3 also provides adsorption energies and Bader charges for linear and bent CO2 on the different Co/CN structures. All the Co/carbon nitrides activated CO2, except for two-layer melon (Co/2-melon-2) and two-layer C3N4 (Co/2-C3N4-2). Several adsorption energies of bent CO2 were quite exothermic (from −2.33 to −1.12 eV), including Co/melem-2, Co/2-melem-2, and Co/melem-dimer. In all cases of bent CO2, the CO2 molecules exhibited O–C–O angles between 128.5° to 145.0°, and the C–O bonds were elongated to 1.20–1.30 Å (compared to the gas-phase CO2 bond length of 1.17 Å), indicating activated states, similar to previous work on other materials.127–129 The Bader charge analysis also shows that bent CO2 molecules were negatively charged from −0.74 |e−| to −0.47 |e−|, while the linear CO2 molecules were close to neutral (−0.10 |e−| to 0.00 |e−|). This again, confirms the anionic, activated state of bent CO2.
 |
| Fig. 7 Most stable structures after CO2 adsorption to Co–carbon nitride materials. Shown are (a) bent CO2 + Co/melem-2, (b) bent CO2 + Co/2-melem-2, (c) bent CO2 + Co/melem-dimer, (d) bent CO2 + Co/2-melem-dimer-2, (e) linear CO2 + Co/2-melem-dimer-2, (f) bent CO2 + Co/melon-1, (g) linear CO2 + Co/2-melon-2, (h) bent CO2 + Co/C3N4-2, and (i) linear CO2 + Co/2-C3N4-2. Brown spheres represent carbon atoms, blue spheres represent nitrogen atoms, pink spheres represent hydrogen atoms, red spheres represent oxygen atoms, and green spheres represent cobalt atoms. | |
Table 3 Adsorption energies and deformation energies of linear and bent CO2 on different Co/carbon nitride systems. The most stable CO2 for each Co/carbon nitride is marked in gray. Eads refers to the adsorption energy, ECo/CN deform is the deformation energy of the Co/carbon nitride ECO2deform, represents the deformation energy of CO2, and Eint is the interaction energy between the adsorbent and CO2
Co/CN structures |
CO2 configuration |
O–C–O angle (°) |
CO2 charge |e−| |
Eads (eV) |
ECo/CN deform (eV) |
ECO2deform (eV) |
Eint (eV) |
Co/melem-2 |
Bent |
141.0 |
−0.67 |
−2.33 |
−0.03 |
1.55 |
−3.85 |
|
Linear |
177.7 |
−0.03 |
−0.16 |
0.01 |
0.01 |
−0.18 |
Co/2-melem-2 |
Bent |
139.0 |
−0.74 |
−1.12 |
1.29 |
1.57 |
−3.98 |
|
Linear |
176.9 |
−0.02 |
−0.23 |
0.03 |
0.01 |
−0.27 |
Co/melem-dimer |
Bent |
141.6 |
−0.60 |
−1.29 |
0.22 |
1.38 |
−2.89 |
|
Linear |
177.7 |
−0.10 |
−0.44 |
0.03 |
0.01 |
−0.48 |
Co/2-melem-dimer-2 |
Bent |
128.5 |
−0.63 |
−0.38 |
1.22 |
2.25 |
−3.86 |
|
Linear |
175.7 |
−0.04 |
−0.38 |
−0.01 |
0.01 |
−0.38 |
Co/melon-1 |
Bent |
141.1 |
−0.60 |
−0.81 |
0.66 |
1.44 |
−2.91 |
|
Linear |
180.0 |
0.00 |
−0.05 |
0.00 |
0.00 |
−0.05 |
Co/2-melon-2 |
Bent |
141.7 |
−0.60 |
0.02 |
0.74 |
1.21 |
−1.93 |
|
Linear |
179.1 |
0.00 |
−0.21 |
0.00 |
0.00 |
−0.20 |
Co/C3N4-2 |
Bent |
145.0 |
−0.47 |
−0.79 |
0.34 |
1.17 |
−2.30 |
|
Linear |
175.8 |
−0.01 |
−0.31 |
0.11 |
0.02 |
−0.45 |
Co/2-C3N4-2 |
Bent |
141.5 |
−0.57 |
1.13 |
2.00 |
1.28 |
−2.16 |
|
Linear |
179.0 |
0.00 |
−0.31 |
0.00 |
0.00 |
−0.31 |
These results demonstrated that Co could significantly improve CO2 adsorption and activation on carbon nitride materials. The Co atoms can facilitate charge transfer and serve as active centers to adsorb and activate CO2 molecules. As we mentioned previously, adsorption of CO2 on pristine g-C3N4,72,78,126,130 and melon131 was reported to form linear, weakly bound CO2. Co/molecular carbon nitrides and Co/single-layer polymeric carbon nitrides could however activate CO2 into bent, exothermic structures. The adsorption energies of linear and bent CO2 over Co/2-melem-dimer-2 were equivalent to each other. We modeled both linear and bent CO2 adsorbed to the Co atom in several different orientations. CO2, however, was only weakly absorbed to Co in the bent structure (Fig. 7d), likely due to steric hindrance as the CO2 situated between two carbon nitride molecules. In the linear structure, the Co atom preferred to interact only with the carbon nitride (Fig. 7e), despite the initial geometry with CO2 placed interacting with the Co atom. Again, steric hindrance prevents linear CO2 from interacting with the Co atom. In addition, we found that Co/two-layer polymeric carbon nitrides (melon and g-C3N4) only weakly absorbed linear CO2, and activation of CO2 into the bent form was endothermic. The Co atoms are sandwiched between the two carbon nitride layers for both carbon nitrides. A linear CO2 molecule placed between the two carbon nitride layers, to interact with the Co atom, was sterically hindered, and the CO2 molecules moved to the locations shown in Fig. 7g and i. Our efforts to form bent CO2 in these two-layered structures led to displacement of Co, and weakening of Co-carbon nitride interactions, as illustrated in Fig. S10.† As shown, in Co/2-melon-2 the Co atom migrates away from one melon layer to bond with a bent CO2. The Co atom after CO2 adsorption only interacts with one carbon nitride layer, and is therefore weakened compared to its pre-adsorbed state (Co interacting with two carbon nitride layers). In the case of Co/2-C3N4-2, an adsorbed, bent CO2 interacts with the Co atom as it sits between the two carbon nitride layers. The CO2 molecule is thus confined between the two layers and experiences steric repulsion, making activation of CO2 unfavorable.
In order to further understand the role of the carbon nitride support, we plotted the stable bent CO2 adsorption energies versus carbon nitride size in Fig. 8. These results show that smaller carbon nitride supports have the strongest adsorption of bent CO2, while larger carbon nitrides have the weakest CO2 adsorption energies. The Sabatier principle prescribes that reactant adsorption energies should neither be too strong or too weak. In this case, intermediate adsorption energies occur for Co/melon-1 (−0.81 eV), Co/2-melem-2 (−1.12 eV), and Co/melem-dimer (−1.29 eV). Thus, based on the Sabatier principle, some molecular and partially condensed complexes may give the best CO2 reduction activity, although further clarification of the reaction mechanism is needed. We should also note that molecular carbon nitrides or melon strands may be representative of exposed carbon nitride edges, which may be a very reactive region.132
 |
| Fig. 8 Correlation between the adsorption energies of CO2 over Co/carbon nitrides and the size of the carbon nitride (i.e., number of triazine rings). The red dashed line shows a linear fit. | |
3.3.2 The nature of linear and activated CO2. Generally bent, activated CO2 was more stable on the Co/carbon nitrides compared to linear CO2, with the exception of two-layer structures, Co/2-melon-2 and Co/2-C3N4-2. We next analyze the different structures to further explain the nature of linear versus bent CO2. For bent CO2 there were significant changes in Co geometry after CO2 adsorption. For structures that activated CO2 (all but two-layer carbon nitrides), we observed elongation of the original Co–N bonds and breaking of the Co–C bonds. But, there were also newly formed Co–C and Co–O bonds with the CO2 molecule. For example, in Co/2-melem-2, the Co atom had four Co–N and two Co–C bonds. After CO2 absorbed, the number of Co–N bonds reduced to two, and the two Co–C bonds broke, while new Co–C and Co–O bonds formed with CO2. For structures that favorably bound linear CO2 (Fig. 7g and i), the geometries and Co coordination environment remained essentially the same as the pre-adsorption state. We also found that Co became more cationic upon interacting with CO2. For instance, the average charge of Co over carbon nitrides that activated bent CO2 increased by +0.22|e−|. On the other hand, the average Co charge for those structures that preferred linear CO2 did not change, again indicating weak interactions.In order to better understand why certain Co/carbon nitrides preferred linear over bent CO2, we examined the deformation and interaction energies of the complexes, which are listed in Table 3. The adsorption energy reflects a balance between the deformation of both the adsorbate and substrate upon adsorption, and their chemical interactions. The deformation energies embody the geometric changes of the adsorbate and substrate from their pre-adsorbed to post-adsorbed states. The interaction energy represents the bond formation energy between the adsorbate and substrate.133–136 The adsorption energy can be written as: ΔEads = Eint + ECo/CNdeform + ECO2deform. Thus, whether adsorption is exothermic or not depends on the nature of structural changes upon adsorption and the nature of favorable (or unfavorable) electronic interactions. Deformation energies (Edeform) of Co/carbon nitrides and CO2 adsorption were calculated by eqn (3) and (4):
|
 | (3) |
|
 | (4) |
Here

is the energy of the Co/carbon nitride complex after CO
2 adsorption, but with CO
2 removed.
ECo/Carbon Nitride is the energy of the geometry of the optimized complex with no CO
2 present.
E′ is the energy of the adsorbed CO
2 minus the Co/carbon nitride complex.
ECO2 is the energy of gas-phase CO
2. The interaction energy (
Eint) was calculated by
eqn (5):
|
Eint = ΔEads − ECo/CN deform − ECO2deform
| (5) |
CO2 deformation energies were always positive for bent adsorption (average value of 1.48 eV), indicating an energy cost to bend the molecule into its activated state. The Co/carbon nitride deformation energies for the bent CO2 case were quite varied, ranging from −0.03 to 2.00 eV, with an average value of 0.81 eV. Since deformation energies for the bent CO2 case were positive, interaction energies must be sufficiently negative for adsorption to be exothermic. Indeed, our results show this to be the case, as these interactions energies were all negative (ranging from −1.93 to −3.98 eV). This strong interaction between bent CO2 and the Co/carbon nitride overcomes geometric distortions to enable exothermic adsorption of CO2. Indeed, bent CO2 becomes anionic, further indicating strong interactions occurring with bent CO2. The significantly more exothermic adsorption energies of bent CO2 molecules on Co/melem-2, Co/2-melem-2, and Co/melem-dimer can be attributed to their more negative interaction energies (from −2.89 to −3.98 eV) compared to Co/melon-1 and Co/C3N4-2 (−2.91 eV and −2.30 eV).
On the other hand, interaction energies for linear CO2 were much weaker, ranging from −0.05 to −0.48 eV. Despite CO2 deformation energies for linear CO2 being close to zero, the interaction energies were not exothermic enough to lead to strong linear CO2 adsorption. CO2 charges were close to neutral for linear CO2, indicating little charge transfer and weak interactions between CO2 and the Co/carbon nitrides.
DOS plots also provide insight on the nature of interactions between CO2 and the Co/carbon nitride catalysts. We show in Fig. 9 example DOS plots for linear and bent CO2 interacting with different Co/carbon nitride complexes. Other DOS plots are shown in Fig. S8 and S9.† These plots show that CO2 in the bent position interacts strongly with Co. See for example the many overlapping peaks involving Co and CO2 near −2, −3.3, an −3.8 eV for Co/melem-dimer. Similar overlapping peaks occur at −1.8, −2.1, −2.5, and −2.9 eV for Co/melon-1. These interactions are due to bond formation between the CO2 molecule and the cobalt atom. On the other hand, such interactions are missing for linear CO2. The CO2 electronic levels are all much lower in energy and overlap with Co is minimized. Both the interactions energies and density of state plots confirm that bent CO2 forms strong bonds with the Co/carbon nitride while linear CO2 does not.
 |
| Fig. 9 Example density of states (DOS) plots for adsprbed CO2, indicating how linear and bent CO2 interact with Co/carbon nitrides: (a) bent CO2 + Co/melem-dimer, (b) linear CO2 + Co/melem-dimer, (c) bent CO2 + Co/melon-1, and (d) linear CO2 + Co/melon-1. The total, carbon nitride (CN), Co, and CO2 DOS are shown. The energies are relative to the Fermi level, which has been shifted to 0 eV and indicated by a dashed vertical line. Due to smearing of the electronic states, the Fermi level appears within the valence band, even though the HOMO energy level is below the Fermi level. | |
3.4 Implications for photocatalysis
Experimental work shows that addition of Co to carbon nitrides improves photoactivity.67,78,116 The exact structure of carbon nitride in these experiments is not fully resolved. We modeled non-polymerized, partially condensed, and fully condensed carbon nitrides, which should provide details on several possible carbon nitrides that could be present in real-world catalysts. Our results showed that Co/carbon nitrides have better photoexcitation compared to pristine carbon nitrides, through more narrow band gaps. This should lead to more photoexcited electrons that may enable photocatalytic reduction. Experiment also indicates that Co lowers electron–hole recombination when added to carbon nitrides.123 This lowered recombination could be facilitated by the mid-gap states created by Co, which we observed. Our results also indicate that in addition to facilitating a larger number of charge carriers to participate in photocatalytic reactions, Co/carbon nitride catalysts also are effective for CO2 activation. In particular, we found that molecular and partially condensed carbon nitride complexes with CO2 may have moderate CO2 adsorption energy (neither too strong nor too weak) to enable CO2 reduction. On the other hand, multi-layer carbon nitrides were found to have difficulties activating CO2. Future work may assess the full catalytic mechanism, but in the least, CO2 activation is a crucial reaction step which Co/carbon nitrides readily enable. All these different effects caused by Co paired with different carbon nitride supports (increased photoexcitation, decreased charge recombination, and increased reactivity) lead to better CO2 reduction and photocatalytic activity (as observed in experiments67–74).
4 Conclusions
Literature has largely focused on two-dimensional g-C3N4 as a support for SACs, despite evidence showing that partial condensation is likely common in working carbon nitride catalysts. We assessed single Co atoms supported on a variety of carbon nitrides as potential photocatalysts. We modeled single Co atoms on six different carbon nitrides with varying degrees of polymerization, including molecular (melem and melem dimer), partially condensed single-layer and two-layer melon, and fully condensed (single-layer and two-layer C3N4). Several stable Co/carbon nitrides were identified with Co forming up to four Co–N bonds, and three Co–C bonds having various Co local geometries. We also discovered correlations between Co coordination number, Co charge, and binding energy. As the Co coordination number increased, the Co atom generally became more oxidized, and the binding energy became more exothermic. Furthermore, binding of Co tended to be strongest with larger carbon nitrides.
In addition, we predicted how Co supported on a carbon nitride could function as a photocatalyst for CO2 reduction. Adding Co to the various carbon nitrides lowered the band gaps, which should lead to better photoexcitation yield. Co atoms also helped activate CO2 into a bent, anionic state, an essential step for CO2 reactivity. The activation of CO2 is facilitated by strong interactions between the CO2 molecule and Co/carbon nitride, which is required to overcome unfavorable bending of the CO2 molecule. For example, the interaction energy between a bent CO2 molecule and Co/melem-2 is −3.85 eV, while the deformation energy of the bent CO2 molecule is 1.55 eV, leading to an overall favorable activation of CO2. Exothermic bent CO2 adsorption energies varied between −2.33 and −0.38 eV, and likely a moderate CO2 adsorption energy (not too strong and not too weak, such as near −1 eV) would best facilitate CO2 reduction. Two melem molecules, a melem dimer, melon, and single-layer g-C3N4 all have moderate CO2 adsorption energies, and thus may be good candidates for CO2 photoreduction. Our work shows that carbon nitrides other than g-C3N4 when paired with single metal atoms have potential as photocatalysts, and therefore should be studied further.
Data availability
Data from the various tables and figures for this article have been included in the ESI.† Simulation files, including geometries, can be found at https://github.com/Deskins-group/Structure-Files/tree/master/Co-Carbon-Nitrides.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
We thank the National Science Foundation for support on this work (NSF Grant 2102198). We also acknowledge the Academic & Research Computing group at Worcester Polytechnic Institute for providing support and computational resources. Computational resources were also provided by the Pittsburgh Supercomputing Center and Purdue University through allocation CHE220030 from the Advanced Cyberinfrastructure Coordination Ecosystem: Services & Support (ACCESS) program, which is supported by U.S. National Science Foundation grants #2138259, #2138286, #2138307, #2137603, and #2138296.
Notes and references
- G. Dong, Y. Zhang, Q. Pan and J. Qiu, J. Photochem. Photobiol. C Photochem. Rev., 2014, 20, 33–50 CrossRef CAS
. - W. J. Ong, L. L. Tan, Y. H. Ng, S. T. Yong and S. P. Chai, Chem. Rev., 2016, 116, 7159–7329 CrossRef CAS PubMed
. - J. Wen, J. Xie, X. Chen and X. Li, Appl. Surf. Sci., 2017, 391, 72–123 CrossRef CAS
. - S. Kumar, S. Karthikeyan and A. F. Lee, Catalysts, 2018, 8, 74 CrossRef
. - S. Cao, J. Low, J. Yu and M. Jaroniec, Adv. Mater., 2015, 27, 2150–2176 CrossRef CAS PubMed
. - Y. Zheng, L. Lin, B. Wang and X. Wang, Angew. Chem., Int. Ed., 2015, 54, 12868–12884 CrossRef CAS PubMed
. - L. Jiang, X. Yuan, Y. Pan, J. Liang, G. Zeng, Z. Wu and H. Wang, Appl. Catal. B Environ., 2017, 217, 388–406 CrossRef CAS
. - X. Wang, S. Blechert and M. Antonietti, ACS Catal., 2012, 2, 1596–1606 CrossRef CAS
. - Y. Zheng, J. Liu, J. Liang, M. Jaroniec and S. Z. Qiao, Energy Environ. Sci., 2012, 5, 6717 RSC
. - A. Naseri, M. Samadi, A. Pourjavadi, A. Z. Moshfegh and S. Ramakrishna, J. Mater. Chem. A, 2017, 5, 23406–23433 RSC
. - G. Liao, Y. Gong, L. Zhang, H. Gao, G. J. Yang and B. Fang, Energy Environ. Sci., 2019, 12, 2080–2147 RSC
. - A. Mishra, A. Mehta, S. Basu, N. P. Shetti, K. R. Reddy and T. M. Aminabhavi, Carbon, 2019, 149, 693–721 CrossRef CAS
. - X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and M. Antonietti, Nat. Mater., 2009, 8, 76–80 CrossRef CAS PubMed
. - D. J. Martin, K. Qiu, S. A. Shevlin, A. D. Handoko, X. Chen, Z. Guo and J. Tang, Angew. Chem., Int. Ed., 2014, 53, 9240–9245 CrossRef CAS PubMed
. - S. Zhang, P. Gu, R. Ma, C. Luo, T. Wen, G. Zhao, W. Cheng and X. Wang, Catal. Today, 2019, 335, 65–77 CrossRef CAS
. - A. Kumar, P. Raizada, P. Singh, R. V. Saini, A. K. Saini and A. Hosseini-Bandegharaei, Chem. Eng. J., 2020, 391, 123496 CrossRef CAS
. - W. Wang, Z. Zeng, G. Zeng, C. Zhang, R. Xiao, C. Zhou, W. Xiong, Y. Yang, L. Lei, Y. Liu, D. Huang, M. Cheng, Y. Yang, Y. Fu, H. Luo and Y. Zhou, Chem. Eng. J., 2019, 378, 122132 CrossRef CAS
. - S. Bhowmik, S. J. Phukan, N. K. Sah, M. Roy, S. Garai and P. K. Iyer, ACS Appl. Nano Mater., 2021, 4, 12845–12890 CrossRef CAS
. - P. Xia, B. Zhu, J. Yu, S. Cao and M. Jaroniec, J. Mater. Chem. A, 2017, 5, 3230–3238 RSC
. - Q. Lu, K. Eid, W. Li, A. M. Abdullah, G. Xu and R. S. Varma, Green Chem., 2021, 23, 5394–5428 RSC
. - X. Huang, W. Gu, Y. Ma, D. Liu, N. Ding, L. Zhou, J. Lei, L. Wang and J. Zhang, Res. Chem. Intermed., 2020, 46, 5133–5164 CrossRef CAS
. - M. Shen, L. Zhang, M. Wang, J. Tian, X. Jin, L. Guo, L. Wang and J. Shi, J. Mater. Chem. A, 2019, 7, 1556–1563 RSC
. - P. Xia, M. Antonietti, B. Zhu, T. Heil, J. Yu and S. Cao, Adv. Funct. Mater., 2019, 29, 1900093 CrossRef
. - A. Thomas, A. Fischer, F. Goettmann, M. Antonietti, J.-O. O. Müller, R. Schlögl and J. M. Carlsson, J. Mater. Chem., 2008, 18, 4893–4908 RSC
. - T. S. Miller, A. B. Jorge, T. M. Suter, A. Sella, F. Corà and P. F. McMillan, Phys. Chem. Chem. Phys., 2017, 19, 15613–15638 RSC
. - Z. Syrgiannis and K. C. Christoforidis, in Handbook of Carbon-Based Nanomaterials, ed. S. Thomas, C. Sarathchandran, S. A. Ilangovan and J. C. Moreno-Pirajan, Elsevier, 2021, pp. 629–670 Search PubMed
. - D. Adekoya, S. Qian, X. Gu, W. Wen, D. Li, J. Ma and S. Zhang, Nano-Micro Lett., 2021, 13, 13 CrossRef PubMed
. - V. W.-h. Lau and B. V. Lotsch, Adv. Energy Mater., 2022, 12, 2101078 CrossRef CAS
. - B. Jürgens, E. Irran, J. Senker, P. Kroll, H. Müller and W. Schnick, J. Am. Chem. Soc., 2003, 125, 10288–10300 CrossRef PubMed
. - J. Dong, Y. Zhang, M. I. Hussain, W. Zhou, Y. Chen and L.-N. Wang, Nanomaterials, 2022, 12, 121 CrossRef CAS PubMed
. - S. T. Melissen, S. N. Steinmann, T. Le Bahers and P. Sautet, J. Phys. Chem. C, 2016, 120, 24542–24550 CrossRef CAS
. - T. Botari, W. P. Huhn, V. W.-h. Lau, B. V. Lotsch and V. Blum, Chem. Mater., 2017, 29, 4445–4453 CrossRef CAS
. - C. Im, B. Kirchhoff, I. Krivtsov, D. Mitoraj, R. Beranek and T. Jacob, Chem. Mater., 2023, 35, 1547–1559 CrossRef CAS
. - Y. Zhang, J. Liu, G. Wu and W. Chen, Nanoscale, 2012, 4, 5300 RSC
. - K. Akaike, K. Aoyama, S. Dekubo, A. Onishi and K. Kanai, Chem. Mater., 2018, 30, 2341–2352 CrossRef CAS
. - V. W.-h. Lau, M. B. Mesch, V. Duppel, V. Blum, J. Senker and B. V. Lotsch, J. Am. Chem. Soc., 2015, 137, 1064–1072 CrossRef CAS PubMed
. - F. Fina, S. K. Callear, G. M. Carins and J. T. S. Irvine, Chem. Mater., 2015, 27, 2612–2618 CrossRef CAS
. - B. Lotsch, M. Döblinger, J. Sehnert, L. Seyfarth, J. Senker, O. Oeckler and W. Schnick, Chem.–Eur. J., 2007, 13, 4969–4980 CrossRef CAS PubMed
. - S. Melissen, T. Le Bahers, S. N. Steinmann and P. Sautet, J. Phys. Chem. C, 2015, 119, 25188–25196 CrossRef CAS
. - S. T. Melissen, T. Le Bahers, P. Sautet and S. N. Steinmann, Phys. Chem. Chem. Phys., 2021, 23, 2853–2859 RSC
. - Y. Li, T. Kong and S. Shen, Small, 2019, 15, 1900772 CrossRef PubMed
. - J. Fu, S. Wang, Z. Wang, K. Liu, H. Li, H. Liu, J. Hu, X. Xu, H. Li and M. Liu, Front. Phys., 2020, 15, 33201 CrossRef
. - S. K. Kaiser, Z. Chen, D. Faust Akl, S. Mitchell and J. Pérez-Ramírez, Chem. Rev., 2020, 120, 11703–11809 CrossRef CAS PubMed
. - R. Kavitha, P. Nithya and S. Girish Kumar, Appl. Surf. Sci., 2020, 508, 145142 CrossRef CAS
. - X. Xiao, L. Zhang, H. Meng, B. Jiang and H. Fu, Sol. RRL, 2021, 5, 1–20 Search PubMed
. - M. Zhao, J. Feng, W. Yang, S. Song and H. Zhang, ChemCatChem, 2021, 13, 1250–1270 CrossRef CAS
. - G. F. R. Rocha, M. A. da Silva, A. Rogolino, G. A. Diab, L. F. Noleto, M. Antonietti and I. F. Teixeira, Chem. Soc. Rev., 2023, 4878–4932 RSC
. - W. A. Qureshi, S. N.-U.-Z. Haider, A. Naveed, A. Ali, Q. Liu and J. Yang, Int. J. Hydrogen Energy, 2023, 48, 19459–19485 CrossRef CAS
. - T. Xiong, W. Cen, Y. Zhang and F. Dong, ACS Catal., 2016, 6, 2462–2472 CrossRef CAS
. - Y. Li, Z. Wang, T. Xia, H. Ju, K. Zhang, R. Long, Q. Xu, C. Wang, L. Song, J. Zhu, J. Jiang and Y. Xiong, Adv. Mater., 2016, 28, 6959–6965 CrossRef CAS PubMed
. - Y. Wang, X. Zhao, D. Cao, Y. Wang and Y. Zhu, Appl. Catal. B Environ., 2017, 211, 79–88 CrossRef CAS
. - L. Liu, X. Wu, L. Wang, X. Xu, L. Gan, Z. Si, J. Li, Q. Zhang, Y. Liu, Y. Zhao, R. Ran, X. Wu, D. Weng and F. Kang, Commun. Chem., 2019, 2, 18 CrossRef
. - F. Zhang, J. Zhang, H. Wang, J. Li, H. Liu, X. Jin, X. Wang and G. Zhang, Chem. Eng. J., 2021, 424, 130004 CrossRef CAS
. - X.-H. Jiang, L.-S. Zhang, H.-Y. Liu, D.-S. Wu, F.-Y. Wu, L. Tian, L.-L. Liu, J.-P. Zou, S.-L. Luo and B.-B. Chen, Angew. Chem., Int. Ed., 2020, 59, 23112–23116 CrossRef CAS PubMed
. - C. Li, N. Su, H. Wu, C. Liu, G. Che and H. Dong, Inorg. Chem., 2022, 61, 13453–13461 CrossRef CAS PubMed
. - J. Shen, C. Luo, S. Qiao, Y. Chen, Y. Tang, J. Xu, K. Fu, D. Yuan, H. Tang, H. Zhang and C. Liu, ACS Catal., 2023, 13, 6280–6288 CrossRef CAS
. - M. Yang, J. Mei, Y. Ren, J. Cui, S. Liang and S. Sun, J. Energy Chem., 2023, 81, 502–509 CrossRef CAS
. - W. Zhang, Q. Peng, L. Shi, Q. Yao, X. Wang, A. Yu, Z. Chen and Y. Fu, Small, 2019, 15, 1905166 CrossRef CAS PubMed
. - X. Xiao, Y. Gao, L. Zhang, J. Zhang, Q. Zhang, Q. Li, H. Bao, J. Zhou, S. Miao, N. Chen, J. Wang, B. Jiang, C. Tian and H. Fu, Adv. Mater., 2020, 32, 2003082 CrossRef CAS PubMed
. - X. Li, W. Bi, L. Zhang, S. Tao, W. Chu, Q. Zhang, Y. Luo, C. Wu and Y. Xie, Adv. Mater., 2016, 28, 2427–2431 CrossRef CAS PubMed
. - Y. Cao, S. Chen, Q. Luo, H. Yan, Y. Lin, W. Liu, L. Cao, J. Lu, J. Yang, T. Yao and S. Wei, Angew. Chem., Int. Ed., 2017, 56, 12191–12196 CrossRef CAS PubMed
. - G. Gao, Y. Jiao, E. R. Waclawik and A. Du, J. Am. Chem. Soc., 2016, 138, 6292–6297 CrossRef CAS PubMed
. - Y. Li, B. Li, D. Zhang, L. Cheng and Q. Xiang, ACS Nano, 2020, 14, 10552–10561 CrossRef CAS PubMed
. - P. Sharma, S. Kumar, O. Tomanec, M. Petr, J. Zhu Chen, J. T. Miller, R. S. Varma, M. B. Gawande and R. Zbořil, Small, 2021, 17, 2006478 CrossRef CAS PubMed
. - L. Li, I. M. U. Hasan, J. Qiao, R. He, L. Peng, N. Xu, N. K. Niazi, J.-N. Zhang and F. Farwa, Nano Res. Energy, 2022, 9120015 Search PubMed
. - P. Chen, B. Lei, X. Dong, H. Wang, J. Sheng, W. Cui, J. Li, Y. Sun, Z. Wang and F. Dong, ACS Nano, 2020, 14, 15841–15852 CrossRef CAS PubMed
. - P. Huang, J. Huang, S. Pantovich, A. Carl, T. Fenton, C. A. Caputo, R. Grimm, A. I. Frenkel and G. Li, J. Am. Chem. Soc., 2018, 140, 16042–16047 CrossRef CAS PubMed
. - J. Jiang, D. Duan, J. Ma, Y. Jiang, R. Long, C. Gao and Y. Xiong, Appl. Catal. B Environ., 2021, 295, 120261 CrossRef CAS
. - L. He, W. Zhang, S. Liu and Y. Zhao, Appl. Catal. B Environ., 2021, 298, 120546 CrossRef CAS
. - P. Huang, S. A. Pantovich, N. O. Okolie, N. A. Deskins and G. Li, ChemPhotoChem, 2020, 4, 420–426 CrossRef CAS
. - P. Huang, J. Huang, J. Li, L. Zhang, J. He, C. A. Caputo, A. I. Frenkel and G. Li, ChemNanoMat, 2021, 7, 1051–1056 CrossRef CAS
. - J. Fu, L. Zhu, K. Jiang, K. Liu, Z. Wang, X. Qiu, H. Li, J. Hu, H. Pan, Y. R. Lu, T. S. Chan and M. Liu, Chem. Eng. J., 2021, 415, 128982 CrossRef CAS
. - J. Li, P. Huang, F. Guo, J. Huang, S. Xiang, K. Yang, N. A. Deskins, V. S. Batista, G. Li and A. I. Frenkel, J. Phys. Chem. C, 2023, 127, 3626–3633 CrossRef CAS
. - P. Huang, J. Huang, J. Li, T. D. Pham, L. Zhang, J. He, G. W. Brudvig, N. A. Deskins, A. I. Frenkel and G. Li, J. Phys. Chem. C, 2022, 126, 8596–8604 CrossRef CAS
. - Y. Zheng, Y. Jiao, Y. Zhu, Q. Cai, A. Vasileff, L. H. Li, Y. Han, Y. Chen and S.-Z. Qiao, J. Am. Chem. Soc., 2017, 139, 3336–3339 Search PubMed
. - X. Wang, Z. Chen, X. Zhao, T. Yao, W. Chen, R. You, C. Zhao, G. Wu, J. Wang, W. Huang, J. Yang, X. Hong, S. Wei, Y. Wu and Y. Li, Angew. Chem., Int. Ed., 2018, 57, 1944–1948 CrossRef CAS PubMed
. - W. Liu, L. Cao, W. Cheng, Y. Cao, X. Liu, W. Zhang, X. Mou, L. Jin, X. Zheng and W. Che, et al., Angew. Chem., 2017, 129, 9440–9445 CrossRef
. - X. Tang, W. Shen, D. Li, B. Li, Y. Wang, X. Song, Z. Zhu and P. Huo, J. Alloys Compd., 2023, 954, 170044 Search PubMed
. - K. Homlamai, T. Maihom, S. Choomwattana, M. Sawangphruk and J. Limtrakul, Appl. Surf. Sci., 2020, 499, 143928 CrossRef CAS
. - C. Ao, B. Feng, S. Qian, L. Wang, W. Zhao, Y. Zhai and L. Zhang, J. CO2 Util., 2020, 36, 116–123 CrossRef CAS
. - D. Ghosh, G. Periyasamy, B. Pandey and S. K. Pati, J. Mater. Chem. C, 2014, 2, 7943–7951 RSC
. - J. Lin, Y. Wang, W. Tian, H. Zhang, H. Sun and S. Wang, ACS Catal., 2023, 13, 11711–11722 CrossRef CAS
. - F. Yu, T. Huo, Q. Deng, G. Wang, Y. Xia, H. Li and W. Hou, Chem. Sci., 2022, 13, 754–762 Search PubMed
. - M. Zheng, H. Xu, Y. Li, K. Ding, Y. Zhang, C. Sun, W. Chen and W. Lin, J. Phys. Chem. C, 2021, 125, 13880–13888 CrossRef CAS
. - G. Kresse and J. Hafner, Phys. Rev. B:Condens. Matter Mater. Phys., 1994, 49, 14251–14269 CrossRef CAS PubMed
. - G. Kresse and J. Hafner, Phys. Rev. B:Condens. Matter Mater. Phys., 1993, 47, 558–561 CrossRef CAS PubMed
. - G. Kresse and J. Furthmüller, Phys. Rev. B:Condens. Matter Mater.
Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed
. - G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS
. - J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed
. - P. E. Blöchl, Phys. Rev. B:Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef PubMed
. - G. Kresse and D. Joubert, Phys. Rev. B:Condens. Matter Mater. Phys., 1999, 59, 1758–1775 CrossRef CAS
. - S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed
. - E. R. Johnson and A. D. Becke, J. Chem. Phys., 2006, 124, 174104 CrossRef PubMed
. - G. Henkelman, A. Arnaldsson and H. Jónsson, Comput. Mater. Sci., 2006, 36, 354–360 CrossRef
. - E. Sanville, S. D. Kenny, R. Smith and G. Henkelman, J. Comput. Chem., 2007, 28, 899–908 CrossRef CAS PubMed
. - W. Tang, E. Sanville and G. Henkelman, J. Phys.: Condens. Matter, 2009, 21, 084204 CrossRef CAS PubMed
. - J. Heyd, G. E. Scuseria and M. Ernzerhof, J. Chem. Phys., 2003, 118, 8207–8215 CrossRef CAS
. - J. Heyd, G. E. Scuseria and M. Ernzerhof, J. Chem. Phys., 2006, 124, 219906 CrossRef
. - A. V. Krukau, O. A. Vydrov, A. F. Izmaylov and G. E. Scuseria, J. Chem. Phys., 2006, 125, 224106 CrossRef PubMed
. - V. Wang, N. Xu, J.-C. Liu, G. Tang and W.-T. Geng, Comput. Phys. Commun., 2021, 267, 108033 CrossRef CAS
. - K. Momma and F. Izumi, J. Appl. Crystallogr., 2011, 44, 1272–1276 CrossRef CAS
. - H. Inoki, G. Seo and K. Kanai, Appl. Surf. Sci., 2020, 534, 147569 CrossRef CAS
. - N. A. Deskins, J. Du and P. Rao, Phys. Chem. Chem. Phys., 2017, 19, 18671–18684 RSC
. - V. Fung and D.-e. Jiang, J. Phys. Chem. C, 2017, 121, 10796–10802 CrossRef CAS
. - A. Zitolo, N. Ranjbar-Sahraie, T. Mineva, J. Li, Q. Jia, S. Stamatin, G. F. Harrington, S. M. Lyth, P. Krtil, S. Mukerjee, E. Fonda and F. Jaouen, Nat. Commun., 2017, 8, 957 CrossRef PubMed
. - L. Cao, Q. Luo, W. Liu, Y. Lin, X. Liu, Y. Cao, W. Zhang, Y. Wu, J. Yang, T. Yao and S. Wei, Nat. Catal., 2018, 2, 134–141 CrossRef
. - J. Lin, L. Jiang, W. Tian, Y. Yang, X. Duan, Y. Jiao, H. Zhang and S. Wang, J. Mater. Chem. A, 2023, 11, 13653–13664 RSC
. - M. Qian, X. Wu, M. Lu, L. Huang, W. Li, H. Lin, J. Chen, S. Wang and X. Duan, Adv. Funct. Mater., 2023, 33, 2208688 CrossRef CAS
. - J. Ding, Z. Teng, X. Su, K. Kato, Y. Liu, T. Xiao, W. Liu, L. Liu, Q. Zhang, X. Ren, J. Zhang, Z. Chen, O. Teruhisa, A. Yamakata, H. Yang, Y. Huang, B. Liu and Y. Zhai, Chem, 2023, 9, 1017–1035 CAS
. - G. F. S. R. Rocha, M. A. R. Da Silva, A. Rogolino, G. A. A. Diab, L. F. G. Noleto, M. Antonietti and I. F. Teixeira, Chem. Soc. Rev., 2023, 52, 4878–4932 RSC
. - E. Vorobyeva, V. C. Gerken, S. Mitchell, A. Sabadell-Rendón, R. Hauert, S. Xi, A. Borgna, D. Klose, S. M. Collins, P. A. Midgley, D. M. Kepaptsoglou, Q. M. Ramasse, A. Ruiz-Ferrando, E. Fako, M. A. Ortuño, N. López, E. M. Carreira and J. Pérez-Ramírez, ACS Catal., 2020, 10, 11069–11080 CrossRef CAS
. - E.-J. Lin, Y.-B. Huang, P.-K. Chen, J.-W. Chang, S.-Y. Chang, W.-T. Ou, C.-C. Lin, Y.-H. Wu, J.-L. Chen, C.-W. Pao, C.-J. Su, C.-H. Wang, U.-S. Jeng and Y.-H. Lai, Appl. Surf. Sci., 2023, 615, 156372 CrossRef CAS
. - M. Wang and Z. Feng, Curr. Opin. Electrochem., 2021, 30, 100803 CrossRef CAS
. - G. D. Fao and J.-C. Jiang, Mol. Catal., 2022, 526, 112402 CrossRef CAS
. - J. Wang, T. Heil, B. Zhu, C.-W. Tung, J. Yu, H. M. Chen, M. Antonietti and S. Cao, ACS Nano, 2020, 14, 8584–8593 CrossRef CAS PubMed
. - Y.-C. Huang, J. Chen, Y.-R. Lu, K. T. Arul, T. Ohigashi, J.-L. Chen, C.-L. Chen, S. Shen, W.-C. Chou, W.-F. Pong and C.-L. Dong, J. Electron Spectrosc. Relat. Phenom., 2023, 264, 147319 CrossRef CAS
. - Y. Xiong, W. Sun, Y. Han, P. Xin, X. Zheng, W. Yan, J. Dong, J. Zhang, D. Wang and Y. Li, Nano Res., 2021, 14, 2418–2423 CrossRef CAS
. - Y. Yang, G. Zeng, D. Huang, C. Zhang, D. He, C. Zhou, W. Wang, W. Xiong, B. Song, H. Yi, S. Ye and X. Ren, Small, 2020, 16, 2001634 CrossRef CAS PubMed
. - S. Wang, J. Li, Q. Li, X. Bai and J. Wang, Nanoscale, 2020, 12, 364–371 RSC
. - J. Ding, Z. Teng, X. Su, K. Kato, Y. Liu, T. Xiao, W. Liu, L. Liu, Q. Zhang, X. Ren, J. Zhang, Z. Chen, O. Teruhisa, A. Yamakata, H. Yang, Y. Huang, B. Liu and Y. Zhai, Chem, 2023, 9, 1017–1035 CAS
. - B. Babu, J. Shim and K. Yoo, Ceram. Int., 2020, 46, 16422–16430 CrossRef CAS
. - X. Yang, Z. Tian, Y. Chen, H. Huang, J. Hu and B. Wen, J. Alloys Compd., 2021, 859, 157754 CrossRef CAS
. - P.-W. Chen, K. Li, Y.-X. Yu and W.-D. Zhang, Appl. Surf. Sci., 2017, 392, 608–615 CrossRef CAS
. - S. Agrawal, D. Casanova, D. J. Trivedi and O. V. Prezhdo, J. Phys. Chem. Lett., 2024, 15, 2202–2208 CrossRef CAS PubMed
. - H.-J. Freund and M. Roberts, Surf. Sci. Rep., 1996, 25, 225–273 CrossRef
. - B. Zhu, L. Zhang, D. Xu, B. Cheng and J. Yu, J. CO2 Util., 2017, 21, 327–335 CrossRef CAS
. - S. K. Iyemperumal and N. A. Deskins, Phys. Chem. Chem. Phys., 2017, 19, 28788–28807 RSC
. - J. Ye, C. Liu and Q. Ge, J. Phys. Chem. C, 2012, 116, 7817–7825 CrossRef CAS
. - W. Zhong, R. Sa, L. Li, Y. He, L. Li, J. Bi, Z. Zhuang, Y. Yu and Z. Zou, J. Am. Chem. Soc., 2019, 141, 7615–7621 CrossRef CAS PubMed
. - L. M. Azofra, D. R. MacFarlane and C. Sun, Phys. Chem. Chem. Phys., 2016, 18, 18507–18514 RSC
. - S. Liu, Y. Li, Y. Zhang and W. Lin, Phys. Chem. Chem. Phys., 2023, 25, 9901–9908 RSC
. - H.-Z. Wu, S. Bandaru, X.-L. Huang, J. Liu, L.-L. Li and Z. Wang, Phys. Chem. Chem. Phys., 2019, 21, 1514–1520 RSC
. - B. Zhu, S. Wageh, A. A. Al-Ghamdi, S. Yang, Z. Tian and J. Yu, Catal. Today, 2019, 335, 117–127 CrossRef CAS
. - Q.-L. Tang and Q.-H. Luo, J. Phys. Chem. C, 2013, 117, 22954–22966 CrossRef CAS
. - C. Morin, D. Simon and P. Sautet, J. Phys. Chem. B, 2003, 107, 2995–3002 CrossRef CAS
. - M. Valero, P. Raybaud and P. Sautet, J. Catal., 2007, 247, 339–355 CrossRef CAS
.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.