DOI:
10.1039/D4TA00152D
(Paper)
J. Mater. Chem. A, 2024,
12, 7724-7731
Effect of ferroelectric polarization on the oxygen evolution reaction: a theoretical study of MIrSn2S6 (M = Bi, Mn, and Sb)†
Received
8th January 2024
, Accepted 13th February 2024
First published on 14th February 2024
Abstract
Ferroelectric polarization plays important roles in catalytic reactions, but the mechanism is still under debate. In this work, 2D ferroelectric MIrSn2S6 is systematically investigated for revealing the effects of ferroelectric polarization on the oxygen evolution reaction (OER) based on density-functional-theory (DFT) calculations. We find that: (1) the adsorption configurations of intermediates, protons, and water strongly depend on the polarization, which in turn affects their bonding patterns and adsorption energies; (2) the paraelectric (PE) state generally shows the highest OER activity (overpotential = 0.54 V) due to its most favorable free energy change; and (3) the capacitances are varied under different polarization states, which affects the reaction kinetics. Our findings illustrate the full picture of the OER process under different polarizations, which may provide insightful understanding on the ferroelectric-enhanced catalytic activity and guide the design of novel ferroelectric catalysts.
1 Introduction
The development of highly active, long-term stable, and low-cost catalysts has gained increasing attention because of their pivotal role in energy conversion and storage.1–5 To improve the catalytic performances, various strategies, such as defect engineering,6,7 composition engineering,8–10 morphology control,11–13 strain engineering,14,15 manipulation of the spin state,11,16 and modulation of the external field,17,18 have been developed recently. Especially, the utilization of ferroelectric polarization has been widely explored.19 For example, perovskites like BaTiO3 showed polarization-dependent adsorption, leading to improved hydrogen evolution reaction (HER).20–23 The adsorption energies of molecules and the efficiency of CO2 reduction reaction (CO2RR) were dependent on the polarization of In2Se3.24,25 Ferroelectricity is also favorable for photocatalysis. For example, perovskites,26,27 Bi2WO6,28 and CuInP2S6
29–31 were applied in photocatalytic water splitting and CO2RR.
The mechanism of ferroelectric-enhanced catalytic activity has been investigated to guide the design of novel catalysts and improve the efficiency. The polarization changes the surface properties,22,32 which should affect the adsorption energy of molecules/ions and charge transfer,20,31,33–35 possibly leading to improved selectivity.25,36 The tilted band induced by the built-in electric field can significantly accelerate carrier migration and separation, which is beneficial to (photo)catalytic applications, especially in multilayer materials.29,37,38 In addition, the carrier transportation can be modulated by the ferroelectric polarization as well because of the modified electrode–catalyst contact.20 The charged surface affects the reaction kinetics too, as indicated by the Tafel slope and reaction impedance.20,38,39 However, there are still important issues to be addressed. The ferroelectric substrate can only provide limited polarization effect.40 In contrast, the surface electronic structure strongly depends on the direction and size of polarization in bulk. The stability of polarization in a catalytic environment is unclear. In addition, the coupling among the polarization, spin, and electronic state needs to be discussed. To address these issues, a simple model is necessary.
Recently, a family of two-dimensional (2D) ferroelectric materials, MIMIIP2X6 (MI and MII = transition metal, X = S, Se, and Te), has attracted much attention because of its rich composition,41–43 diverse magnetic and electronic properties,44–46 and potential application in catalysis.29,47 When Ge replaces P in MIMIIP2X6, ferroelectric MIMIIGe2X6 (X = S, Se, Te) can be obtained, which further expands the materials' family.48,49 Especially, the surfaces of paraelectric and ferroelectric phases of MIMIIGe2X6 are very similar because the Ge ions protrude from both sides in the paraelectric (PE) phase, and only one Ge ion moves towards the center in the ferroelectric (FE) phase (Fig. 1a and b). MIMIIGe2X6 also exhibits rich magnetic and electronic properties,48,49 which may be used for studying the effect of ferroelectric polarization on the catalytic performance. However, the Ge ions are catalytically inactive because the Gibbs free energy change of *OH adsorption is too positive or negative (Fig. S1†). To obtain a series of catalytically active materials with varied electronic properties, we replace Ge with Sn because Sn has strong metallic properties, and been widely used as catalysts in the oxygen evolution reaction (OER). Meanwhile, S and Ir are selected for X and MII, respectively, because this combination shows plentiful electronic properties in MIMIIGe2X6.48,49 Therefore, MIMIISn2X6 is designed to be an ideal model catalyst for the purpose.
 |
| Fig. 1 (a) Top and (b) side views of PE and FE MnIrSn2S6, and the charge density difference between FE- and PE-MnIrSn2S6. The arrow represents the direction of polarization. (c) Tilted view of AFE MnIrSn2S6, and the charge density difference between AFE- and PE-MnIrSn2S6. (d) Phonon spectrum and (e) AIMD simulation of FE-MnIrSn2S6. (f) Minimum energy pathway for the phase change of MnIrSn2S6. Band structures and PDOSs of (g) FE, (h) PE, and (i) AFE-MnIrSn2S6. | |
In this work, the PE, FE, and anti-ferroelectric (AFE) phases of MIrSn2S6 (M = Bi, Mn, Sb, and In) are comprehensively studied to determine the role of ferroelectricity in catalytic reactions. We find that the ferroelectricity of PE-InIrSn2S6 changes in the reaction. For the other three systems, the adsorption configurations in the different phases are distinct. The PE phase generally has the best OER performance due to the most suitable adsorption energy for intermediates, which is attributed to the different bonding patterns between Sn and intermediates. The adsorption energies of H2O and protons are generally high on the AFE and FE phases. We further show that the capacitances of the three phases are different because their surface charge responds differently to the applied voltage, indicating that the ion transportation in the reaction may be affected by ferroelectricity.
2 Results and discussion
2.1 Structure, stability, magnetic, and electronic properties
The primitive cells of PE- and FE-MIrSn2S6 (M = Bi, Mn, Sb, and In) are optimized first (Fig. 1a and b). The S ions are located on both sides of MIrSn2S6. Mn and Ir are sandwiched between the two S layers. The positions of the two Sn ions determine the polarization of MIrSn2S6: one Sn protrudes from the S layer in the FM phase, while the PE phase has both the Sn ions above the surfaces (Fig. 1b). The electric polarization is also confirmed by the charge density difference between PE- and FE phases (Fig. 1b). The lattice constants of FE and PE phases are similar (Table S1†). The AFE phase is then obtained by using a 2 × 2 × 1 supercell of MIrSn2S6 (Fig. 1c). PE-MIrSn2S6 has the highest energy among the three phases, which is consistent with their ferroelectric properties. The AFE state has slightly lower energy than the FE state, except for SbIrSn2S6 (Table S2†). The formation energies of MIrSn2S6 are negative, indicating that it is possible to fabricate them (Table S1†). The stabilities of MIrSn2S6 were also tested. FE-MIrSn2S6 are all dynamically stable as there are negligible imaginary frequencies around the Γ points in the phonon dispersions (Fig. 1d, and S2a–c†). Meanwhile, huge imaginary frequencies can be seen in PE-MIrSn2S6, indicating the possible phase change (Fig. S2d†). The calculated elastic constants (Table S3†) and AIMD simulations (Fig. 1e and S3†) further prove that FE-MIrSn2S6 are mechanically and thermodynamically stable. The minimum energy pathways between FE- and −FE-MIrSn2S6 show that the energy barrier is low for the transition from PE to FE phase (Fig. 1f and S4†).
The magnetic and electronic properties of MIrSn2S6 were then investigated. Only MnIrSn2S6 is magnetic because of the huge energy difference between the total energies with (EM) and without (ENM) spin (Table S1†). The magnetic ground states of FE-, PE-, and AFE-MnIrSn2S6 are ferromagnetic with an exchange energy of over 70 meV per atom (Table S4†). Interestingly, the Mn ion shows a magnetic moment of 2.00 and 2.08 μB, respectively, in FE-/AFE- and PE-MnIrSn2S6 (Table S1†) because of different electronic properties. FE-MnIrSn2S6 is semiconductive with a band gap of ∼1.01/0.50 eV for the spin-up/down channel (Fig. 1g). Mn contributes mostly to the density of states (DOS) around the Fermi level (EF). PE-MnIrSn2S6 is conductive (Fig. 1h), and both Mn and S dominate the DOS around EF. Sn has little contribution to DOS in both FE- and PE-MnIrSn2S6. AFE-MnIrSn2S6 is semiconductive and has similar electronic properties to FE-MnIrSn2S6 (Fig. 1i). The band structure of InIrSn2S6 is similar to that of MnIrSn2S6, in which the FE and AFE phases are semiconductive and the PE phase is conductive (Fig. S5†). The three phases of BiIrSn2S6 and SbIrSn2S6 are all semiconductive. The PE phase of BiIrSn2S6 has a much smaller (∼0.19 eV) band gap compared to its FE (∼1.02 eV) and AFE (∼0.89) phase (Fig. S6a–c†). The three phases of SbIrSn2S6 have a similar band gap of ∼0.71 eV (Fig. S6d–f†). The flat bands are mainly contributed by S and Ir around EF in the band structures for the FE and AFE phases of BiIrSn2S6 and SbIrSn2S6 (Fig. S6b, c, e and f†). It is worth noting that the heavy atom Bi leads to strong spin-orbital coupling (SOC), which results in clear band splitting and a band gap of only 0.02 eV in PE-BiIrSn2S6 (Fig. S7†). Overall, MIrSn2S6 with special structure and rich electronic properties should provide an ideal platform to study the role of ferroelectricity in catalysis.
2.2 Adsorption under different polarizations
To investigate the OER activity of the PE-, FE-, and AFE-MIrSn2S6, a 2 × 2 × 1 supercell with a lattice constant of over 12 Å was used. The active sites for the OER were systematically tested first. We find that the protruding Sn ion is the only active site, as *OH always connects to Sn after relaxation whatever the initial adsorption configuration is (Fig. S8†). As Sn is the active sites after testing, the OER cannot happen on the surface of the −FE phase. We find that all three phases for MnIrSn2S6, BiIrSn2S6, and SbIrSn2S6 can maintain their structures during the OER process. The adsorption configurations of *OO and *H on PE-, FE-, and AFE-MIrSn2S6 are similar. Take MnIrSn2S6 as an example, the Sn ion moves downward in the PE phase, but has little change in the FE and AFE phases after adsorbing *OO (Fig. 2a). The Sn ion also moves downward slightly after adsorbing *H (Fig. S9†). H2O cannot adsorb strongly on the surface of MIrSn2S6 as the distance between Sn and O is over 3 Å (Fig. 2c and S12†). However, the adsorption configurations of *OH, *O, and *OOH are clearly different on PE-, FE-, and AFE-MIrSn2S6. The Sn ion in the PE phase moves slightly downward, but is pulled towards another (a neighboring) Sn ion in the FE and AFE phases after *OOH, *O, or *OH is adsorbed (Fig. 2b, S10 and S11†). Specifically, the *O drops on the surface and bonds with S in AFE-MnIrSn2S6, and connects to the neighboring Sn in FE-MnIrSn2S6, leading to the movement of Sn (Fig. S11†). The distance between Sn and intermediates (i.e., Sn–O or Sn–H bond) shows the biggest variation in the case of *OO and *H2O, which are all steps without involving charge transfer. For example, the *OO can weakly adsorb on the top of the Sn ion with a Sn–O bond length of ∼2.46 to 3.17 Å in MnIrSn2S6 (Fig. 2a and c), but the Sn–H and Sn–O bonds in *H and *OH have similar bond length of ∼1.75 Å and ∼2.00 Å in MnIrSn2S6 respectively (Fig. 2c, S9 and S10†). The Sn–O bond for the *OOH intermediate in FE-BiIrSn2S6 is longer than those in the other two phases (Fig. S12†).
 |
| Fig. 2 The adsorption configuration of (a) *OO and (b) *OOH. (c) The comparison of bond lengths between Sn and intermediates on PE, FE, and AFE-MnIrSn2S6 surfaces. (d) The initial and final structures of *O and *OH adsorption on the PE-InIrSn2S6 surface. | |
Different from MIrSn2S6 (M = Bi, Mn, and Sb), the PE-InIrSn2S6 cannot hold its structure in the OER process. The Sn ions bonded with intermediates remain unchanged, but the other Sn atoms on the upper surface move towards the center of the lattice after adsorption, indicating the instability and change of ferroelectric properties in the OER process (Fig. 2d). In addition, the Sn ion connecting with intermediates moves towards the center of the lattice when *OH is adsorbed, while it moves little when *O is adsorbed (Fig. 2d), which may be due to the interaction between the substrate and the intermediates. It is also worth noting that the energy barriers of PE–FE phase change of InIrSn2S6 and BiIrSn2S6 are both small (<0.2 eV), but phase change does not occur on PE-BiIrSn2S6 in the OER process, which may be due to their intrinsic properties (Fig. S4†).
2.3 Adsorption energy and reaction free energy
To obtain the adsorption energies (Ead) of the proton, water molecule, and intermediates in the OER process (*OH, *O, *OOH, and *OO), the spin state is carefully checked.50 We find that only *OO shows clear spin polarization (Fig. S13 and S19†). The net magnetic moment of *OO is generally 2 μB, such as in PE-MnIrSn2S6, which is consistent with our previous study.50 However, the net magnetic moment of *OO is ∼1.5 μB in one excited state in FE-MnIrSn2S6 (Fig. S13†), which is different from those in PE-MnIrSn2S6 and the other two substrates. Here, the Ead is calculated by using the ground state energy.
The Ead values of *OOH, *O, and *OH are distinct in MIrSn2S6 (M = Bi, Mn, and Sb). The Ead values of *O on PE- and FE-MnIrSn2S6 are nearly 0, but negative on AFE-MnIrSn2S6 (Fig. 3a). The Ead of *O on PE-BiIrSn2S6 is still nearly 0, but is −0.57 eV and −1.90 eV on AFE- and FE-BiIrSn2S6, respectively (Fig. S14a†). FE- and AFE-SbIrSn2S6 can tightly bond with *O with an Ead of −2.5 eV, but the Ead of *O on PE-SbIrSn2S6 is only slightly negative (Fig. S15a†). The other intermediates have distinct Ead on different surfaces too. For example, the Ead values of *OH on FE- (−2.08 eV) and AFE-MnIrSn2S6 (−1.79 eV) are only ∼89% and 76% of that on PE-MnIrSn2S6 (−2.34 eV), respectively (Fig. 3a). The Ead of O2 on the three phases is similar for MnIrSn2S6 and SbIrSn2S6. For the adsorption of O2 on BiIrSn2S6, the Ead (−0.20 eV) is negative on the PE surface, but positive on the FE (0.30 eV) and AFE (0.33 eV) surfaces (Fig. S14a†). The adsorption of *H and *H2O on MnIrSn2S6 is also significantly different. The *H and *H2O are adsorbed slightly on the PE-MnIrSn2S6 surface with an Ead close to ∼0 eV. On the other hand, *H and *H2O show much higher Ead on the FE- and AFE-MnIrSn2S6 surfaces, indicating difficult adsorption (Fig. 3a). Such differences may also affect the OER performance. However, the Ead values of *H2O are similar on BiIrSn2S6 and SbIrSn2S6. The Ead of *H on PE-BiIrSn2S6 is significantly lower than that on FE- and AFE-BiIrSn2S6, but similar on all three phases of SbIrSn2S6 (Fig. S14a and S15a†).
 |
| Fig. 3 (a) Ead of intermediates and (b) diagrams of Gibbs free energy change in the OER process on PE-, FE-, and AFE-MnIrSn2S6. The overpotential is marked in the figure. | |
The contrast Ead corresponds to the distinct Gibbs free energy change in the OER process on the three surfaces. We take MnIrSn2S6 as an example, the higher Ead values of *OH and *O result in lower Gibbs free energy changes (ΔG) in the first (* + OH− → *OH + e−) and second (*OH + OH− → *O + H2O + e−) steps on the FE and AFE surfaces. However, the huge ΔG in the third (*O + OH− → *OOH + e−) step leads to the worse OER performances of FE- and AFE-MnIrSn2S6, which is also the potential determining step (PDS). In contrast, the PDS of PE-MnIrSn2S6 is the second step, and the three phases of MnIrSn2S6 have a similar overpotential (∼1 V) (Fig. 3b). The second and third steps are the PDS for PE- and AFE-BiIrSn2S6, respectively, and their overpotentials are similar (∼1 V). Differently, the OER is hard to happen on FE-BiIrSn2S6 due to the huge ΔG in the third step (Fig. S14b†). FE- and AFE-SbIrSn2S6 also need to overcome a huge energy barrier in the third step. However, the PE-SbIrSn2S6 only needs an overpotential of 0.54 V in this step, leading to a good OER performance (Fig. S15b†), which is comparable to those of recognized excellent OER catalysts (Table S5†).
To understand the mechanism, we investigated the bonding of each intermediate by the crystal orbital Hamilton population (COHP) analysis, which illustrates that polarization can affect the bonding patterns. Take the *OH intermediate as an example, the COHP patterns of the Sn–O bond on the FE and AFE surfaces are similar, but distinct from that on the PE state. For example, one antibonding state around −6 eV in the FE and AFE states changes into two in the PE state (Fig. 4a–c). The integration of COHP (ICOHP) on PE-MnIrSn2S6 is slightly lower than those on FE- and AFE-MnIrSn2S6, corresponding to the higher Ead of *OH on the PE surface (Fig. 4a–c). Similarly, the Sn–O bonds for the *OOH intermediate in FE and AFE states are clearly distinct from that in the PE state (Fig. S17a†). Specifically, the COHP patterns of Sn–O bonds for the *O intermediate in all three states are different, and the bonding states of the Sn–O bond in AFE-MnIrSn2S6 are higher than those in FE or PE states, corresponding to the different adsorption configurations of *O. The O–O bonding in *OOH is also affected by polarization (Fig. S17b†). The O–H bonding in *OH and *OOH is also affected in the three states. For example, the shapes of two antibonding states around −7 eV are different in *OH (Fig. 4d–f and S17c†). The Sn–O bonds for the *OO intermediate on the three phases are all different, while the O–O bond of *OO is similar to that in O2, corresponding to the weak adsorption of O2 (Fig. S18 and S19†). Spin polarization can be observed in the Sn–O and O–O bonding for *OO, consistent with our previous statements. We also investigated the Sn–H bonding for *H. The COHP patterns of *H are distinct in all three phases. Although the Ead values of *H on the three phases are clearly different, the ICOHPs of Sn–H bonding are similar (Fig. 4g–i).
 |
| Fig. 4 The COHP of (a–c) Sn–O bonding for the *OH, (d–f) O–H bonding for the *OH, and (g–i) Sn–H bonding for the *H intermediate on (left to right) PE, FE, and AFE-MnIrSn2S6, respectively. | |
2.4 Potential–energy relationship and reaction kinetic analysis
Under an external circuit, the surface charge of the electrode is unbalanced, which shall greatly affect the catalytic activity and the kinetics of the OER.51 When PE-, FE-, and AFE-MnIrSn2S6 are under a balanced surface charge, the corresponding potentials are also different, and intermediates have complex effects on the surface charge (Fig. 5a–c and S20†). For example, the surface of AFE-MnIrSn2S6 is negatively charged, but those of FE- and PE-MnIrSn2S6 are positively charged naturally (Fig. 5a). However, AFE-MnIrSn2S6 is close to a zero-potential state (−0.007 V vs. SHE) when the *OH is adsorbed (Fig. 5b), but a negative potential (−0.23 V vs. SHE) can balance the surface charge when *OOH is adsorbed (Fig. 5c). Meanwhile, negative (−0.41 V vs. SHE) and positive (+0.37 V vs. SHE) potentials are needed, respectively, for FE and PE-MnIrSn2S6 to have a balanced surface charge when *OH is adsorbed (Fig. 5b), but both positive (+0.41 and +0.86 V vs. SHE for PE and FE, respectively) when *OOH is adsorbed (Fig. 5c).
 |
| Fig. 5 The potential vs. free energy and their fitting of (a) the pure surface, (b) *OH, and (c) *OOH intermediates on PE, FE, and AFE-MnIrSn2S6. The gray lines indicate the potential when the system has a balanced surface charge. (d) The capacitance of the surface and surface with intermediates. The number represents the percentage relative to the capacitance on the PE surface. | |
By simulating the relationship between the potential and free energy on MnIrSn2S6 with each intermediate adsorbed (see also the supplementary notes), we can see the effect of applied potential. When a huge negative potential is applied, the free energies of systems increase dramatically, but the free energy has a good quadratic relationship with potential in the range ∼ −1.5 to 3 V. Thus, the surface capacity can be obtained by fitting the curve (Fig. 5a–c). We can see that the capacitance of the pure surface of MnIrSn2S6 follows the sequence PE < FE < AFE. However, when the surface is covered by intermediates, the capacitance of PE-MnIrSn2S6 may get larger than those of the other two phases, such as when *OH is adsorbed (Fig. 5d). Despite the similar OER performance and the COHP pattern of Sn–O, the capacitances of FE- and AFE-MnIrSn2S6 are also distinct. Specifically, FE-MnIrSn2S6 has larger surface capacitance for *O, but lower for *OOH and *OO than AFE-MnIrSn2S6. Take the PE-MnIrSn2S6 as the reference, the capacitance differences of the three phases can be up to ∼73% and 181% when intermediates are adsorbed, which would affect the kinetic process in the OER greatly (Fig. 5d).
3 Conclusions
In summary, the OER performances of PE-, FE-, and AFE-MIrSn2S6 are systematically studied. We show that the polarization may change during the OER. The polarization affects the adsorption configuration, which further decides the bonding patterns and adsorption energies of intermediates. PE-MIrSn2S6 shows the best performance among the three phases, especially for SbIrSn2S6, which only has an overpotential of 0.54 V. In addition, the adsorption energies of protons and H2O are significantly different, which may affect the OER activity as well. Finally, the surface charges on PE-, FE-, and AFE-MIrSn2S6 have a distinct response to potential, resulting in the different capacitances of them, indicating that the kinetic process in the OER may be affected by polarization. Our work not only provided highly active OER catalysts, but also revealed the effect of ferroelectric polarization in catalytic reactions in multiple aspects, which may provide insightful understanding of the reaction mechanism.
4 Computational methods
The first-principles calculations were performed with the Vienna Ab initio Simulation Package (VASP) code.52–54 The Perdew–Burke–Ernzerhof (PBE) version of the generalized gradient approximation (GGA) was used for the exchange–correlation functional.55 A vacuum of over 15 Å along the z-direction was used to avoid the interaction between neighboring images. The primitive cells of MIrSn2S6 (M = Bi, Mn, Sb, and In) were first optimized with a plane-wave cutoff energy of 550 eV and a Γ-centered Monkhorst–Pack k point mesh of 5 × 5 × 1.56 The convergence criterion for energy and the force tolerance for ionic relaxation were set as 10−6 eV and 0.005 eV Å−1, respectively. The formation energies of MIrSn2S6 were described as Ef = (E0 − EIr − EM − 2ESe − 6ES), where E0 is the total energy for MIrSn2S6, and EIr, EM, ESe and ES are the chemical potentials determined by the most stable Ir, M, Sn crystals and S8 amorphous, respectively.57 A 3 × 3 × 1 k-mesh was used to obtain the PE, FE, and AFE phases of MIrSn2S6 in the 2 × 2 × 1 supercell, which is also used for investigation of catalytic performances. The Density-Functional Perturbation Theory (DFPT) method was used to obtain phonon dispersion spectra of MIrSn2S6 with phonopy code,58,59 in which the 3 × 3 × 1 supercell and 3 × 3 × 1 Monkhorst–Pack k-point sampling were adopted, and the kinetic-energy cutoff was set to 500 eV. The ab initio molecular dynamics (AIMD) simulations were carried out with a 3 × 3 × 1 supercell and Γ point at 300 K with a time step of 2 fs.60 The climbing-image nudged-elastic band (cI-NEB) method is employed to get the ferroelectric switching pathways.61
Spin-polarized DFT calculation was performed to study the OER performance of MIrSn2S6 (M = Bi, Mn, and Sb). The vdW correction of Grimme's D3 scheme was chosen to treat the weak interactions in the system.62 A k-point grid of 3 × 3 × 1, and cut-off energy of 450 eV were used. The convergence thresholds of energy and force were 10−5 eV and 0.02 eV Å−1, respectively. The adsorption energy of gas molecules was determined by Ead = Esub+gas − Esub − Egas, where Esub+gas, Esub and Egas are the total energies of the adsorbed system, the substrate and gas molecules, respectively. The Gibbs free energy was calculated as G = E + ZPE − TS, where E is the total energy obtained by DFT calculation, and ZPE is zero-point energy, which is calculated as ZPE = 1/2∑hνi, where νi is the frequency of intermediates. T is the temperature of reaction, which is set as 298.15 K, and S is entropy. The overpotential of OER vs. RHE is defined as ηORR = max{ΔGi}/e − 1.23, where e is the electron charge and ΔG refers to the change of Gibbs free energies in each reaction step. The dipole correction was found to have negligible effect on the trends of the OER performance of systems (Fig. S21†).
Conflicts of interest
The authors declare no conflict of interest.
Acknowledgements
This work was supported by the Science and Technology Development Fund (FDCT) from Macau SAR (0050/2023/RIB2, 0023/2023/AFJ, 006/2022/ALC, and 0111/2022/A2), Multi-Year Research Grants (MYRG-GRG2023-00010-IAPME, and MYRG2022-00026-IAPME) from Research & Development Office at University of Macau, and Shenzhen-Hong Kong-Macao Science and Technology Research Programme (Type C) (SGDX20210823103803017) from Shenzhen. The DFT calculations were performed at the High Performance Computing Cluster (HPCC) of the Information and Communication Technology Office (ICTO) at the University of Macau.
References
- Z. Y. Yu, Y. Duan, X. Y. Feng, X. Yu, M. R. Gao and S. H. Yu, Adv. Mater., 2021, 2007100 CrossRef CAS PubMed.
- D. Liu, L. Qiao, S. Peng, H. Bai, C. Liu, W. F. Ip, K. H. Lo, H. Liu, K. W. Ng, S. Wang, X. Yang and H. Pan, Adv. Funct. Mater., 2023, 33, 2303480 CrossRef CAS.
- D. Liu, R. Tong, Y. Qu, Q. Zhu, X. Zhong, M. Fang, K. Ho Lo, F. Zhang, Y. Ye, Y. Tang, S. Chen, G. Xing and H. Pan, Appl. Catal., B, 2020, 267, 118721 CrossRef CAS.
- M. Li, Y. Liu, L. Dong, C. Shen, F. Li, M. Huang, C. Ma, B. Yang, X. An and W. Sand, Sci. Total Environ., 2019, 668, 966–978 CrossRef CAS PubMed.
- M. I. Din, A. G. Nabi, Z. Hussain, R. Khalid, M. Iqbal, M. Arshad, A. Mujahid and T. Hussain, Int. J. Energy Res., 2021, 45, 8370–8388 CrossRef.
- Y. Zhu, X. Liu, S. Jin, H. Chen, W. Lee, M. Liu and Y. Chen, J. Mater. Chem. A, 2019, 7, 5875–5897 RSC.
- J. Feng, X. Zhong, M. Chen, P. Zhou, L. Qiao, H. Bai, D. Liu, D. Liu, Y.-Y. Chen, W. F. Ip, S. Chen, J. Ni, D. Liu and H. Pan, J. Power Sources, 2023, 561, 232700 CrossRef CAS.
- J. Ma, F. Zhu, Y. Pan, H. Zhang, K. Xu, Y. Wang and Y. Chen, Sep. Purif. Technol., 2022, 288, 120657 CrossRef CAS.
- M. Qu, X. Ding, Z. Shen, M. Cui, F. E. Oropeza, G. Gorni, V. A. D. L. P. O'Shea, W. Li, D.-C. Qi and K. H. L. Zhang, Chem. Mater., 2021, 33, 2062–2071 CrossRef CAS.
- Y. Tian, S. Wang, E. Velasco, Y. Yang, L. Cao, L. Zhang, X. Li, Y. Lin, Q. Zhang and L. Chen, iScience, 2020, 23, 100756 CrossRef CAS PubMed.
- Y. Tong, Y. Guo, P. Chen, H. Liu, M. Zhang, L. Zhang, W. Yan, W. Chu, C. Wu and Y. Xie, Chem, 2017, 3, 812–821 CAS.
- Y. Kong, T. He, A. R. P. Santiago, D. Liu, A. Du, S. Wang and H. Pan, ChemSusChem, 2021, 14, 3257–3266 CrossRef CAS PubMed.
- B. Yang, Y. Wang, B. Gao, L. Zhang and L. Guo, ACS Catal., 2023, 13, 10364–10374 CrossRef CAS.
- F. Li, H. Ai, D. Liu, K. H. Lo and H. Pan, J. Mater. Chem. A, 2021, 9(33), 17749–17759 RSC.
- D. Liu, H. Ai, W. T. Lou, F. Li, K. H. Lo, S. Wang and H. Pan, Sustainable Energy Fuels, 2020, 4, 3773–3779 RSC.
- Z. Duan and G. Henkelman, ACS Catal., 2020, 10, 12148–12155 CrossRef CAS.
- J. Yao, W. Huang, W. Fang, M. Kuang, N. Jia, H. Ren, D. Liu, C. Lv, C. Liu, J. Xu and Q. Yan, Small Methods, 2020, 4, 2000494 CrossRef CAS.
- H. Bai, J. Feng, D. Liu, P. Zhou, R. Wu, C. T. Kwok, W. F. Ip, W. Feng, X. Sui, H. Liu and H. Pan, Small, 2023, 19(5), 2205638 CrossRef CAS PubMed.
- W. Ding, J. Lu, X. Tang, L. Kou and L. Liu, ACS Omega, 2023, 8, 6164–6174 CrossRef CAS PubMed.
- P. Abbasi, M. R. Barone, M. d. l. P. Cruz-Jauregui, D. Valdespino-Padilla, H. Paik, T. Kim, L. Kornblum, D. G. Schlom, T. A. Pascal and D. P. Fenning, Nano Lett., 2022, 22, 4276–4284 CrossRef CAS PubMed.
- E. S. Beh, S. A. Basun, X. Feng, I. U. Idehenre, D. R. Evans and M. W. Kanan, Chem. Sci., 2017, 8, 2790–2794 RSC.
- A. Kakekhani and S. Ismail-Beigi, ACS Catal., 2015, 5, 4537–4545 CrossRef CAS.
- M. H. Zhao, D. A. Bonnell and J. M. Vohs, Surf. Sci., 2008, 602, 2849–2855 CrossRef CAS.
- X. Tang, J. Shang, Y. Gu, A. Du and L. Kou, J. Mater. Chem. A, 2020, 8, 7331–7338 RSC.
- L. Ju, X. Tan, X. Mao, Y. Gu, S. Smith, A. Du, Z. Chen, C. Chen and L. Kou, Nat. Commun., 2021, 12(1), 5128 CrossRef CAS PubMed.
- G. Huang, G. Zhang, Z. Gao, J. Cao, D. Li, H. Yun and T. Zeng, J. Alloys Compd., 2019, 783, 943–951 CrossRef CAS.
- S. Ninova and U. Aschauer, J. Mater. Chem. A, 2017, 5, 11040–11046 RSC.
- G. Zhang, J. Cao, G. Huang, J. Li, D. Li, W. Yao and T. Zeng, Catal. Sci. Technol., 2018, 8, 6420–6428 RSC.
- Y. Fan, X. Song, H. Ai, W. Li and M. Zhao, ACS Appl. Mater. Interfaces, 2021, 13, 34486–34494 CrossRef CAS PubMed.
- S. Huang, Z. Shuai and D. Wang, J. Mater. Chem. A, 2021, 9, 2734–2741 RSC.
- L. Ju, J. Shang, X. Tang and L. Kou, J. Am. Chem. Soc., 2020, 142, 1492–1500 CrossRef CAS PubMed.
- K. Garrity, A. Kakekhani, A. Kolpak and S. Ismail-Beigi, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 88(4), 045401 CrossRef.
- S. Sanna, R. Hölscher and W. G. Schmidt, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 86(20), 205407 CrossRef.
- J. H. Lee and A. Selloni, Phys. Rev. Lett., 2014, 112(19), 196102 CrossRef PubMed.
- I. Efe, N. A. Spaldin and C. Gattinoni, J. Chem. Phys., 2021, 154, 024702 CrossRef CAS PubMed.
- N. Saito, H. Nishiyama, K. Sato and Y. Inoue, Surf. Sci., 2000, 454–456, 1099–1103 CrossRef CAS.
- M. A. Khan, M. A. Nadeem and H. Idriss, Surf. Sci. Rep., 2016, 71, 1–31 CrossRef CAS.
- X. Li, Y. Du, L. Ge, C. Hao, Y. Bai, Z. Fu, Y. Lu and Z. Cheng, Adv. Funct. Mater., 2022, 2210194 Search PubMed.
- X. Li, H. Liu, Z. Chen, Q. Wu, Z. Yu, M. Yang, X. Wang, Z. Cheng, Z. Fu and Y. Lu, Nat. Commun., 2019, 10(1), 1409 CrossRef PubMed.
- K. Garrity, A. M. Kolpak, S. Ismail-Beigi and E. I. Altman, Adv. Mater., 2010, 22, 2969–2973 CrossRef CAS PubMed.
- R. R. Ma, D. D. Xu, Q. L. Zhong, C. R. Zhong, R. Huang, P. H. Xiang, N. Zhong and C. G. Duan, Adv. Mater. Interfaces, 2021, 2101769 Search PubMed.
- S. A. Tawfik, J. R. Reimers, C. Stampfl and M. J. Ford, J. Phys. Chem. C, 2018, 122, 22675–22687 CrossRef CAS.
- L. Chen, C. Mao, J.-H. Chung, M. B. Stone, A. I. Kolesnikov, X. Wang, N. Murai, B. Gao, O. Delaire and P. Dai, Nat. Commun., 2022, 13(1), 4037 CrossRef CAS PubMed.
- M. Chyasnavichyus, M. A. Susner, A. V. Ievlev, E. A. Eliseev, S. V. Kalinin, N. Balke, A. N. Morozovska, M. A. Mcguire and P. Maksymovych, Appl. Phys. Lett., 2016, 109, 172901 CrossRef.
- Y. Lai, Z. Song, Y. Wan, M. Xue, C. Wang, Y. Ye, L. Dai, Z. Zhang, W. Yang, H. Du and J. Yang, Nanoscale, 2019, 11, 5163–5170 RSC.
- X. Ma, L. Sun, J. Liu, X. Feng, W. Li, J. Hu and M. Zhao, Phys. Status Solidi RRL, 2020, 14, 2000008 CrossRef CAS.
- L. Ji, L. Chang, Y. Zhang, S. Mou, T. Wang, Y. Luo, Z. Wang and X. Sun, ACS Catal., 2019, 9, 9721–9725 CrossRef CAS.
- K.-R. Hao, X.-Y. Ma, Z. Zhang, H.-Y. Lyu, Q.-B. Yan and G. Su, J. Phys. Chem. Lett., 2021, 12, 10040–10051 CrossRef CAS PubMed.
- X. Y. Ma, H. Y. Lyu, K. R. Hao, Y. M. Zhao, X. Qian, Q. B. Yan and G. Su, Sci. Bull., 2021, 66, 233–242 CrossRef CAS PubMed.
- H. Bai, D. Liu, P. Zhou, J. Feng, X. Sui, Y. Lu, H. Liu and H. Pan, J. Mater. Chem. A, 2022, 10(47), 25262–25271 RSC.
- H. N. Nong, L. J. Falling, A. Bergmann, M. Klingenhof, H. P. Tran, C. Spöri, R. Mom, J. Timoshenko, G. Zichittella, A. Knop-Gericke, S. Piccinin, J. Pérez-Ramírez, B. R. Cuenya, R. Schlögl, P. Strasser, D. Teschner and T. E. Jones, Nature, 2020, 587, 408–413 CrossRef CAS PubMed.
- W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133–A1138 CrossRef.
- P. Hohenberg and W. Kohn, Phys. Rev., 1964, 136, B864–B871 CrossRef.
- G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
- M. M. Obeid, A. Bafekry, S. U. Rehman and C. V. Nguyen, Appl. Surf. Sci., 2020, 534, 147607 CrossRef CAS.
- A. Togo, F. Oba and I. Tanaka, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78(13), 134106 CrossRef.
- S. Baroni, S. D. Gironcoli, A. D. Corso and P. Giannozzi, Rev. Mod. Phys., 2001, 73, 515–562 CrossRef CAS.
- W. G. Hoover, Phys. Rev. A, 1985, 31, 1695–1697 CrossRef PubMed.
- D. Sheppard, P. Xiao, W. Chemelewski, D. D. Johnson and G. Henkelman, J. Chem. Phys., 2012, 136, 074103 CrossRef PubMed.
- S. Grimme, J. Comput. Chem., 2006, 27, 1787–1799 CrossRef CAS PubMed.
|
This journal is © The Royal Society of Chemistry 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.