Next Article in Journal
Industrial Scale Direct Liquefaction of E. globulus Biomass
Previous Article in Journal
Cu-Doped SrTiO3 Nanostructured Catalysts for CO2 Conversion into Solar Fuels Using Localised Surface Plasmon Resonance
Previous Article in Special Issue
Mechanism of Methanol Synthesis from CO2 Hydrogenation over Cu/γ-Al2O3 Interface: Influences of Surface Hydroxylation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Noble Metal Single-Atom Coordinated to Nitrogen, Oxygen, and Carbon as Electrocatalysts for Oxygen Evolution

1
School of Physics and Information Technology, Shaanxi Normal University, Xi’an 710119, China
2
Research Center for Carbon-Based Electronics, Key Laboratory for the Physics and Chemistry of Nanodevices, School of Electronics, Peking University, Beijing 100871, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(10), 1378; https://doi.org/10.3390/catal13101378
Submission received: 25 September 2023 / Revised: 16 October 2023 / Accepted: 17 October 2023 / Published: 19 October 2023
(This article belongs to the Special Issue Theory-Guided Electrocatalysis and Photocatalysis)

Abstract

:
Tuning the coordination environment centering metal atoms has been regarded as a promising strategy to promote the activities of noble metal single-atom catalysts (SACs). In the present work, first-principle calculations are employed to explore the oxygen evolution reaction (OER) performance of Ir and Ru SACs with chemical coordination being nitrogen (M-N4-C), oxygen (M-O4-C), and carbon (M-C4-C) in graphene, respectively. A “three-step” strategy was implemented by progressively investigating these metrics (stability, catalytic activity, structure–activity relationship). A volcano plot of reactivity is established by using the adsorption-free energy of O* (∆GO*) as a theoretical descriptor. The intrinsic OER activity is IrN4-C > IrO4-C > RuO4-C > RuN4-C > IrC4-C > RuC4-C. The in-depth tuning mechanism of ∆GO* can be indicated and interpreted by the d-band centers of the active sites and the crystal orbital Hamilton population analysis of metal-oxygen bonds, respectively.

1. Introduction

Electrochemical water splitting is a promising approach for producing highly purified hydrogen using renewable electricity, encompassing both the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER) [1,2,3,4,5]. Nevertheless, the large-scale application of electrolytic water splitting is hindered by the absence of highly active and cost-effective electrocatalysts to overcome the slow kinetics of the OER. While the prevalent RuO2 and IrO2 electrocatalysts are active in driving OER, their large-scale utilization in water-splitting devices is limited by the scarcity [6,7,8]. As a result, atomically dispersed Ir and Ru atoms fixed on various supporting matrices have attracted extensive attention due to their advantages in minimizing the usage of noble metals and promoting reaction kinetics in OER [9,10]. However, it is noteworthy that carbonaceous supports, renowned for their exceptional conductivity and plentiful defects, provide distinct advantages in supporting single-atom electrocatalysts. These heteroatoms (e.g., N, O, S, P) can be intentionally incorporated and controlled using experimental techniques to achieve the desirable OER performance. Additionally, these heteroatoms can establish stable coordination with the single metal atom, to a certain extent, impeding their aggregation during both the reaction and preparation processes [11]. Interestingly, in this field, the local environments of metal atoms play a crucial role in not only anchoring metal atoms, but also tuning the electronic properties of active sites [12,13,14,15]. In recent years, tuning the coordination shell of single-atom catalysts (SACs) using hetero atoms has been widely reported [16,17,18,19,20,21,22]. For instance, Wang et al. demonstrated that Co-N2C2 exhibited higher selectivity and activity than those of Co-N3C1 and Co-N4 on electroreduction of CO2 to CO [23]. Zhao et al. revealed that Mo-S2O2-C outperformed Mo-N4-C, Mo-O4-C, and Mo-S4-C sites in oxygen reduction reaction (ORR), with a low theoretical onset potential being 0.87 V [19]. Yuan et al. reported Fe-N3P as an outstanding ORR catalyst. The limiting potential (1.22 eV) was less than that of Fe-N4 (1.36 eV) [24]. Jiang et al. revealed that the Ag-N2C2 site exhibited much better H2 evolution activity than the Ag-N4 site on carbon nitride [25]. Cheng et al. indicated that the hydrogen evolution reaction (HER) activity increases in the order of Mo-O2N2, Mo-O2N1C1, and Mo-O2C2, with overpotentials being 98 mV, 71 mV, and 61 mV at 10 mA cm−2 [26]. In addition, SACs have been the subject of extensive research in the field of the OER [27,28,29]. Niu et al. investigated the OER activity of a single transition metal supported on defective g-C3N4 with N vacancy (TM/VN-CN) using DFT calculations. Their findings revealed that Ru/VN-CN exhibited exceptional performance with an overpotential being 0.32 V [27]. Wang et al. reported the OER performance of FeXYiN3−i (X, Y = B, C, O, P, and S; i = 0, 1) moiety in Fe–porphyrin with first-principle calculations, and they indicated that FeC2N2-II showed superior catalytic activity with the overpotential being 0.17 V [28]. Su et al. synthesized the Pt1–C2N2 SAC catalyst, which sustained a current density of 120 mA/cm2 at an overpotential of 405 mV and exhibited a high mass activity of 3350 Ag−1 at 232 mV [29]. Given these premises, it is imperative to explore the catalytic performance of Ir and Ru coordinated to N, O, and C as electrocatalysts for OER.
In the present work, OER catalytic activities of three distinct coordination environments of graphene-supported Ir SACs and Ru SACs were comprehensively investigated with density functional theory (DFT) calculations. The catalytic performance of these sites was compared in terms of stability, binding strength of intermediate states, and limiting potentials. A volcano plot of activity was established by using the binding free energy of the oxygen atom (ΔGO*) as a theoretical descriptor. The ΔGO* is sensitive to the micro-environments of metal atoms. The in-depth binding characteristics of the key intermediate state O* can be revealed by crystal orbital Hamilton population analysis. The quantitative trend of ΔGO* of different active sites can be indicated by the d-band center of the active sites. This work not only reports IrN4-C as an outstanding OER catalyst, but also reveals the fundamental structure–activity relationship of single Ir and Ru atoms in different coordination environments.

2. Results and Discussion

2.1. Structures and Stability of Ir, Ru Embedded N4-C/O4-C/C4-C

Figure 1 shows optimized structures of MN4-C, MO4-C, and MC4-C (M = Ir, Ru). In MN4-C and MO4-C, all atoms are nearly in the same plane, which is consistent with previous studies [30,31,32]. In MC4-C, by contrast, the C atoms in the coordination shell and the M atom protrude out of the graphene substrate. Such structural characteristics can be interpreted by the orbital geometry of sp2 (O, N) and sp3 (C) hybridizations. The lattice parameters of MN4-C, MO4-C, and MC4-C (M = Ir, Ru) are a = 12.27 Å~12.30 Å, b = 12.52 Å, and c = 19.82 Å~19.87 Å, which align with a previous study [33]. Experimentally, SACs are often prone to aggregation during synthesis and catalysis, emphasizing the significance of evaluating the thermal stability of SACs in calculations. Currently, comparing the binding energy and cohesive energy of single metal atoms is an effective approach for estimating the thermodynamic stability of SACs [27,34,35]. In this work, the binding energies (Ebind) of metal atoms were calculated as shown in Figure 2. The theoretical cohesive energies of bulk Ir and Ru are 5.15 eV and 5.17 eV, respectively [36]. Thus, MN4-C, MO4-C, and MC4-C (M = Ir, Ru) exhibit satisfied stabilities considering that their binding energies are less than −5.00 eV. As seen in Figure 2, the Ir atoms in IrO4-C and the Ru atoms in RuO4-C (Ebind = −2.19 eV and −2.73 eV, respectively) could potentially undergo aggregation during catalysis. However, the electronic structures and catalytic performance of IrO4-C and RuO4-C are still studied in this work in order to obtain the structure–activity relationship.

2.2. OER Catalytic Activity

The free energy profiles of OER on MN4-C, MO4-C, and MC4-C (M = Ir, Ru) were explored by DFT calculations. In the traditional adsorbate evolution mechanism (AEM) mechanism, OER involves four proton-coupled electron transfer (PCET) steps, as shown in Scheme 1. The elementary reactions are [37,38]:
OH + * → OH* + e
OH* + OH → O* + H2O + e
O* + OH → OOH* + e
OOH* + OH → * + O2 (g) + H2O + e
where * represents the active site of catalysts, and OH*, O*, and OOH* represent three different catalytic intermediates adsorbed on the active site. The first PCET step is the addition of one hydroxyl group to the M (M = Ir, Ru) atom, forming OH* accompanied by the transfer of one electron. Then, OH* further reacts with the second hydroxyl group to attain O*. Subsequently, the third hydroxyl group is absorbed on the surface of O* to give OOH* intermediate. Finally, OOH* combines with the fourth hydroxyl group by PCET, resulting in the release of O2 and the regeneration of the free active site.
The corresponding Gibbs free energy profiles of MN4-C, MO4-C, and MC4-C (M = Ir, Ru) catalysts are shown in Figure 3. Optimized structures of the different catalytic intermediates for OER can be depicted in Figures S1–S6. Gibbs free energy change (∆G) for each OER reaction step on MN4-C, MO4-C, and MC4-C can be seen in Table S1. In these calculations, the Gibbs free energy of the initial state of OER is defined as 0.00 eV. The cumulative free energy change of the entire OER process is 4.92 eV. By identifying ΔGmax [37,39,40], the initial potential and corresponding potential determination step (PDS) can be extracted from the free energy diagrams of the OER processes. As shown in Figure 3a, the second PCET step (OH* + OH → O* + H2O + e), which forms the O* intermediate, is the PDS of OER on IrN4-C. The free energy of all electrochemical reaction steps becomes downhill if the applied potentials are higher than 1.53 eV for IrN4-C. The η OER value of IrN4-C is 0.30 V, indicating that IrN4-C exhibits outstanding catalytic activity towards OER. The PDSs of OER on IrO4-C, IrC4-C, RuN4-C, RuO4-C, and RuC4-C are all the third PCET (O* + OH → OOH* + e) step with the η OER values being 0.51 V, 1.43 V, 0.99 V, 0.62 V, and 1.73 V, respectively (Figure 3b–f). In addition, the first PCET step (OH + * → OH* + e) of OER is exothermic for IrC4-C, RuN4-C, and RuC4-C, whereas other PCETs are endothermic, suggesting that stronger OH* binding leads to poorer activity of IrC4-C, RuN4-C, and RuC4-C. Note that the η OER value of IrN4-C is at least 0.21 V lower than that of IrO4-C and IrC4-C, indicating that the OER catalytic activity of IrN4-C is significantly better than that of IrO4-C and IrC4-C. Similarly, the OER catalytic performance of RuO4-C is superior to that of RuN4-C and RuC4-C (0.62 V vs. 0.99 V and 1.73 V). These reveal that OER activities are closely related to the coordination environment of SACs [41,42]. Moreover, the chemical identity of noble metal atoms also plays a crucial role in tuning the catalytic performance [27]. For instance, the η OER   value of IrN4-C (0.30 V) is smaller than that of RuN4-C (0.99 V). Thus, the OER activity order of MN4-C, MO4-C, and MC4-C (M = Ir, Ru) is: IrN4-C > IrO4-C > RuO4-C > RuN4-C > IrC4-C > RuC4-C.

2.3. Origin of the OER Activity

According to the Sabatier principle [43], too strong or too weak binding interactions of intermediates on catalysts will show a negative impact on energy conversion efficiency. Therefore, the catalytic activities can often be evaluated or predicted by the adsorption energies of intermediates (ΔGOH*, ΔGO*, ΔGOOH*). Table 1 shows that adsorption-free energies of intermediates vary with the metal atoms and their coordination environments. Interestingly, the variation of adsorption-free energies of OOH* and OH* on the catalysts are smaller than 1.38 eV. Nevertheless, the variation of ΔGO* on different catalysts is much more profound (about 2.48 eV). Thus, the adsorption of O atoms is more sensitive to the microenvironment of active sites than other intermediate states. The OER activity of MN4-C, MO4-C, and MC4-C (M = Ir, Ru) can probably be evaluated by the binding energy of O* (ΔGO*).
Figure 4a shows that the relationship between ΔGOH* and ΔGOOH* can be expressed as ΔGOOH* = 0.95ΔGOH* + 3.03 with a high coefficient of determination (R2 = 0.990). This suggests that ΔGOOH* rises with increasing ΔGOH*. Therefore, ΔGOOH* and ΔGOH* exhibit almost the same weight when they are used as theoretical descriptors. As shown in Figure S7, however, ΔGOOH* and ΔGOH* do not exhibit strong correlations with the catalytic activities. We then plot the d-band center with respect to the ΔGO*. Interestingly, as seen in Figure S8, the linear relationship can only be seen in the planar active sites (M-N4-C and M-O4-C). The MC4-C sites deviated from the scaling relationship due to geometrical distortions. In addition, the formation of O* from OOH* is the PDS of most catalysts. The exception is IrN4-C, in which the PDS step is the conversion of OH* to O*. These findings indicate that the η OER can be simultaneously related to the ΔGO*. The volcano plot is therefore established when ΔGO* is selected as a descriptor for OER activity (Figure 4b). It is apparent that IrN4-C stands on the top of the OER volcano plot with the ΔGO* value being 2.49 eV. IrO4-C exhibits unsatisfied catalytic efficiency owing to weak O* binding ( η OER = 0.51 V). While the IrC4-C, RuN4-C, RuO4-C, and RuC4-C exhibit small formation energies of O*, also lead to low OER activity ( η OER = 1.43 V, 0.99 V, 0.62 V, and 1.73 V, respectively).
Subsequently, crystal orbital Hamilton populations (COHP) and integrated COHP (ICOHP) are employed to describe the adsorption trend of intermediated states [44,45,46]. COHP is a prevalent method to analyze the bonding characteristics between atoms in chemical species and can often provide important information in the study of structure–activity relationships. For instance, Yan et al. demonstrated that there is a strong linear relationship between the ΔGOH* and ICOHP with an R2 of 0.91, explaining the OH* adsorption trend. Thus, the Co-N-C/GDY achieved superior OER performance due to its moderate binding strength with OH* species [45]. Wang et al. revealed that the bonding strength between the 3d metal atom and O* intermediate generally decreases with the atomic number increases for different MBene substrates by employing ICOHP calculations [46]. The binding strength of the M-O bond in MN4-C-O*, MO4-C-O*, and MC4-C-O* (M = Ir, Ru) can be further examined by COHP analysis. As depicted in Figure 5, the bonding (–COHP > 0) and antibonding (–COHP < 0) contributions in each M-O bond are shown on the right and left panels, respectively. The ICOHP values of the M-O bond in IrO4-C-O* (−4.14 eV) are more than that in IrN4-C-O* (−5.24 eV). The ICOHP values of the M-O bond in IrC4-C-O*, RuN4-C-O*, RuO4-C-O*, and RuC4-C-O* are −6.88 eV, −6.62 eV, −5.57 eV, and −8.45 eV, respectively, which is more negative compared to IrN4-C-O*. These findings reveal that the binding strength of the oxygen atom in IrO4-C is weaker than that in IrN4-C, while the strong binding of the oxygen atom is observed in IrC4-C, RuN4-C, RuO4-C, and RuC4-C compared to IrN4-C. Thus, IrN4-C exhibits excellent catalytic OER performance in these catalysts due to the moderate adsorption energy of O*. The OER performance of MN4-C, MO4-C, and MC4-C (M = Ir, Ru) could be evaluated by the binding energy of O* (ΔGO*), which is consistent with the above discussions about the volcano plot in Figure 4b.
As it has been demonstrated that ΔGO* is a theoretical descriptor to evaluate the OER activity, it is important to study the structure–property relationship between the binding energy of O* and the microstructure of the active sites. The charge density difference (CDD) of MN4-C, MO4-C, and MC4-C (M = Ir, Ru) was examined with iso-surfaces set at 0.015 e/Å3. As shown in Figure 6, electrons transfer (0.54 e~0.82 e) from M atoms to the N4-C/O4-C/C4-C substrate (Tables S3 and S4). Such charge redistributions are consistent with CDD analysis, showing the accumulation of significant positive charges (pink area) around the anchored M atoms (Figure 6). It is worth noting that formal charges carried by metal atoms are strongly dependent on coordination shells. For instance, the Bader charge of Ir atoms are +0.56|e| and +0.72|e| on the C4-C and N4-C substrates, respectively. Notably, the d-band center (εd) has been extensively employed as an effective descriptor to establish the relation between the intrinsic electronic properties of catalysts and their adsorption behaviors [27,47]. For instance, Niu et al. established a linear correlation between the εd and the adsorption energy of OER intermediates, providing a quantitative description of adsorption energies and revealing the origin of activity based on electronic structures. Consequently, Ru embedded in defective g-C3N4 exhibits exceptional performance, with an overpotential as low as 0.32 V [27]. Yang et al. demonstrated that the overpotential of CoNi in nitrogen-doped graphene shows a volcano-shaped relationship with the εd of the catalytic site Co. This suggests that the εd of the catalytic site is a critical parameter influencing the OER activity [47]. The partial density of states (PDOS) of d orbitals for M atoms on MN4-C, MO4-C, and MC4-C (M = Ir, Ru) are shown in Figure 7. As can be seen, replacing carbon with oxygen or nitrogen for coordination with M metal shifts the εd of the M atom toward the Fermi level (Ef). For example, the εd values for Ir atoms on IrN4-C, IrO4-C, and IrC4-C exhibit the following trend: IrN4-C (−1.86 eV) > IrO4-C (−2.20 eV) > IrC4-C (−2.60 eV). In addition, the εd values for Ru atoms on the C4-C, O4-C, and N4-C substrates are −1.81 eV, −1.10 eV, and −1.07 eV (Figure 7b), respectively. The closer the εd of the active site is to the Fermi level, the stronger its ability to participate in capturing reaction intermediates and forming bonds, resulting in higher catalytic activity. Specifically speaking, a more negative εd is related to weaker adsorption [48,49]. However, in Ru SACs, the activity of RuO4-C surpasses that of RuC4-C and Ru-N4-C. It is explicable that the association between the εd and catalytic performance is merely a qualitative description and not a quantitative trend. Thus, the moderate εd value of Ir atoms on IrN4-C can lead to a neither too strong nor too weak binding strength of oxygen atom (ΔGO*). As a result, IrN4-C shows excellent OER performance with the lower overpotential being 0.30 V.

2.4. Solvent Effects

Given that OER takes place in aqueous electrolytes, the impact of solvent effects in the reaction pathways of MN4-C, MO4-C, and MC4-C (M = Ir, Ru) has also been investigated using an implicit solvent model. The Gibbs free energy profiles of OER on MN4-C, MO4-C, and MC4-C (M = Ir, Ru) catalysts in a vacuum (black lines) and implicit solvent model (red lines) are shown in Figure S9. Gibbs free energy changes (∆G) for each OER reaction step on MN4-C, MO4-C, and MC4-C can be seen in Table S5. The third PCET (O* + OH → OOH* + e) is the PDS of OER on MN4-C, MO4-C, and MC4-C (M = Ir, Ru) in the implicit solvent model, which is quite consistent with those in the vacuum. As can be seen in Figure S8, the η OER value of IrN4-C is 0.33 V in the implicit solvent model, which is only slightly increased by 0.03 V ( η OER = 0.30 V in a vacuum). The η OER values of IrO4-C and RuO4-C are 0.71 V and 0.67 V in the implicit solvent model, which are 0.20 V and 0.05 V higher than that in a vacuum, respectively. The η OER values of IrC4-C and RuC4-C are 1.34 V and 1.58 V in the implicit solvent model, which is 0.09 V and 0.15 V less than that in a vacuum. These results are from the diverse influence of the implicit solvent model on ΔGO*. In addition, the adsorption-free energy of *OH on IrC4-C, RuN4-C, and RuC4-C is decreased compared with that in a vacuum, indicating stronger interactions between the O atom and the M (M = Ir, Ru) atom. While the overpotential of MN4-C, MO4-C, and MC4-C (M = Ir, Ru) increases by −0.15 V~0.20 V, the qualitative trends of the influence of coordination shells are not affected by the solvation. This finding aligns well with recent research reports [50,51,52]. Zhang et al. reported that in the implicit solvent model, the onset potential of Pt (111) increases by 0.11 V compared to that in a vacuum, and the ORR performance trend of metal surfaces is consistent with that in a vacuum [51]. Peng et al. showed that including solvent effects increases the overpotential of the Fe- or Ni-doped Co3O4(001) by around 0.20 V~0.40 V. However, the general OER activity trends of Fe- or Ni-doped Co3O4(001) remain unchanged [52]. Therefore, our results demonstrate that the implicit solvent model can enhance O* adsorption by finely tuning the OER activity. Additionally, the implicit solvent model has a minor impact on the Gibbs free energy profile. The trend of catalytic performance in the implicit solvent model closely mirrors that observed in vacuum. IrN4-C exhibits excellent OER catalytic performance compared to IrO4-C, IrC4-C, RuN4-C, RuO4-C, and RuC4-C.

2.5. Structural Distortions and OER Activity

The impact of structural distortions at the evaluated coordination sites on catalytic activity was thoroughly investigated. Among the various SACs studied, IrC4-C and Ru-C4-C exhibit the most pronounced structural distortions. Notably, these structural deformations lead to a situation where the single metal atom protrudes from the C4-C substrate, resulting in excessively strong binding with O* intermediates. Consequently, IrC4-C and Ru-C4-C show poor catalytic performance in OER. In addition, RuN4-C displays minor structural distortions in comparison to IrN4-C, IrO4-C, and RuO4-C, which shows small formation energies of O*, and also leads to low OER activity. These findings indicated that structural distortions induce a strong interaction between the metal atom and reaction intermediates, consequently impeding the efficiency of the oxygen evolution reaction. This discovery provides invaluable insights for the ongoing efforts to optimize catalyst design and achieve enhanced catalytic performance.

3. Computational Methods

All DFT calculations were carried out using a plane-wave basis set within the Vienna Ab Initio Simulation Package (VASP 5.4.4) [53,54,55]. The exchange–correlation function was described by the Perdew–Burke–Ernzerhof (PBE) parameterization of the generalized gradient approximation (GGA) [56,57,58,59]. The projector-augmented wave (PAW) approximation was adopted to describe the ion–electron interactions [60]. The plane-waves expansion of the Kohn–Sham orbitals was expanded to a kinetic energy cutoff of 400 eV. Then, 10−4 eV was set as the convergence criteria for the energy, and −0.02 eV/Å was set as the convergence criteria for the forces on each atom. The DFT-D3 method with Becke–Jonson damping was adopted to accurately describe the long-range van der Waals (vdW) interactions [61]. The solvation effect was evaluated under an implicit solvent model VASPsol [62].
MN4-C, MO4-C, and MC4-C (M = Ir, Ru) catalysts were constructed by embedding an M atom in N4-C, O4-C, and C4-C cavities, respectively. A vacuum thickness of more than 15 Å in the vertical direction was utilized to avoid interactions between periodic images. The Brillouin zone was sampled using a 2 × 2 × 1 gamma-centered k-point mesh for structural optimizations, while a denser k-point mesh of 4 × 4 × 1 was employed for electronic property computations. The binding energy (Ebind) was calculated using the equation:
E bind = E total E substrate E M
where E total , E substrate , and E M are energies of the catalysts, doped graphene substrates, and an isolated M atom in a vacuum, respectively. The adsorption energy of three different catalytic intermediates can be described as follows:
E OH * = E OH * + 1 / 2 E H 2 E * E H 2 O
E O * = E O * +   E H 2 E * E H 2 O
E OOH * = E OOH * + 3 / 2 E H 2 E * 2 E H 2 O
where E*, EOH*, EO*, and EOOH* are energies of the clean catalyst substrate, and surfaces adsorbed by OH*, O*, and OOH* species, respectively. E H 2 O and E H 2 are energies of H2O and H2 molecules in gas phases, as shown in Table S6. The adsorption-free energies can be obtained in the equation:
G ads = E ads +   E ZPE T S  
where   E ads and E ZPE are the adsorption energy and zero-point energy, respectively. T is fixed at 298.15 K in this study and S is the entropy. For all OER steps on MN4-C, MO4-C, and MC4-C (M = Ir, Ru) catalysts, the Gibbs free energy change (ΔG) can be calculated by the following expression [63]:
G = E +   E ZPE T S + G pH
where E is the energy of the reactant and product obtained from DFT computations directly. G pH = k B Tln 10 × pH represents the free energy contribution due to the variations in the H concentration, where kB is the Boltzmann constant. In this work, the pH value was set to 14. The theoretical OER overpotential ( η OER ) for a given electrocatalyst can be defined as [64]:
η OER = max Δ G 1 ,   Δ G 2 ,   Δ G 3 ,   Δ G 4 / e 1.23   V
where Δ G i (i = 1, 2, 3, 4) represents the Gibbs free energy change for each step of OER, and the value of 1.23 eV is the equilibrium overpotential of OER. The charge density difference was obtained by the following equation:
ρ =   ρ AB ρ A ρ B
where   ρ AB , ρ A , and ρ B are electron densities of complexes, substrates, and adsorbates, respectively.

4. Conclusions

In summary, by means of first-principle calculations, the theoretical increase in OER activity on MN4-C, MO4-C, and MC4-C (M = Ir, Ru) was IrN4-C > IrO4-C > RuO4-C > RuN4-C > IrC4-C > RuC4-C in the alkaline media. In particular, IrN4-C exhibits outstanding OER performance with a low overpotential of 0.30 V. Our calculations indicate that the weak binding of O* to IrO4-C results in poor catalyst activity, and the strong binding of O* is responsible for the lower OER activity observed in IrC4-C, RuN4-C, RuO4-C, and RuC4-C. Furthermore, the analysis of COHP and PDOS highlights the enhancement of electrocatalytic activity with the tuning of the coordination shells of M atoms, which effectively adjusts the electronic structure of M atoms to balance the binding strength of O*. Our study challenges traditional thinking and inspires more theoretical and experimental investigations into exploring the catalytic properties of SACs by considering different coordination environments with other metals.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13101378/s1, Figure S1: Optimized structures of (a) IrN4-C catalyst and (bd) OER intermediates along the pathway on IrN4-C. Ir, C, N, H, and O atoms are presented with blue, brown, yellow, white, and red circles, respectively. Figure S2: Optimized structure of (a) IrO4-C catalyst and (bd) OER intermediates along the pathway on IrO4-C. Ir, C, H, and O atoms are presented with blue, brown, white, and red circles, respectively. Figure S3: Optimized structure of (a) IrC4-C catalyst and (bd) OER intermediates along the pathway on IrC4-C. Ir, C, H, and O atoms are presented with blue, brown, white, and red circles, respectively. Figure S4: Optimized structure of (a) RuN4-C catalyst and (bd) OER intermediates along the pathway on RuN4-C. Ru, C, N, H, and O atoms are presented with purple, brown, yellow, white, and red circles, respectively. Figure S5: Optimized structure of (a) RuO4-C catalyst and (bd) OER intermediates along the pathway on RuO4-C. Ru, C, H, and O atoms are presented with purple, brown, white, and red circles, respectively. Figure S6: Optimized structure of (a) RuC4-C catalyst and (bd) OER intermediates along the pathway on RuC4-C. Ru, C, H, and O atoms are presented with purple, brown, white, and red circles, respectively. Figure S7: The scaling relationship between (a) ∆GOH* vs. η OER , (b) ∆GOOH* vs. η OER on MN4-C, MO4-C, and MC4-C (M = Ir, Ru). Figure S8: The scaling relationship between ∆GO* vs. d-band center on MN4-C, MO4-C, and MC4-C (M = Ir, Ru). Figure S9: Free energy diagrams of OER processes on (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C in vacuum (black lines) and implicit solvent model (red lines), respectively. H, C, O, N, Ir, and Ru atoms are represented with white, brown, red, yellow, blue, and purple circles. Table S1: Gibbs free energy change (∆G) for each OER reaction step on MN4-C, MO4-C, and MC4-C (M = Ir, Ru) in vacuum. Table S2: The d-band center of M atom on MN4-C, MO4-C, and MC4-C (M = Ir, Ru). Table S3: Bader charges of IrN4, IrO4-C, IrC4-C. The ZVAL represents the number of valent electrons in each atomic sphere. Table S4: Bader charges of RuN4, RuO4-C, and RuC4-C. The ZVAL represents the number of valent electrons in each atomic sphere. Table S5: Gibbs free energy change (∆G) for each OER reaction step on MN4-C, MO4-C, and MC4-C (M = Ir, Ru) with an implicit solvent model. Table S6: The calculated total energies (E) and thermodynamic quantities for the gas phase H2 species (T = 298.15 K, P = 1 bar), and free H2O at 298.15 K, 0.035 bar.

Author Contributions

Conceptualization, J.B.; methodology, Y.C.; investigation, J.W., J.B. and Y.C.; data curation, J.W., J.B., Y.C., Q.L., X.F. and H.L.; writing—original draft preparation, J.W.; writing—review and editing, H.L. and X.F.; supervision, Q.L., X.F. and H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grants 22373063, 22302005, 22272099, and 22072102), Fundamental Research Funds for the Central Universities of China (Grants GK202203002, GK202201001, and GK202301001), and China Postdoctoral Science Foundation (2023M730044).

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Acknowledgments

We gratefully acknowledge Shaanxi Normal University’s high-performance computing infrastructure SNNUGrid (HPC Centers: QiLin AI LAB) for providing computer facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Luyen Doan, T.L.; Tran, D.T.; Nguyen, D.C.; Tuan Le, H.; Kim, N.H.; Lee, J.H. Hierarchical three-dimensional framework interface assembled from oxygen-doped cobalt phosphide layer-shelled metal nanowires for efficient electrocatalytic water splitting. Appl. Catal. B 2020, 261, 118268. [Google Scholar] [CrossRef]
  2. Shi, Y.; Zhang, B. Recent advances in transition metal phosphide nanomaterials: Synthesis and applications in hydrogen evolution reaction. Chem. Soc. Rev. 2016, 45, 1529–1541. [Google Scholar] [CrossRef]
  3. Zheng, Y.; Jiao, Y.; Jaroniec, M.; Qiao, S.Z. Advancing the Electrochemistry of the Hydrogen-Evolution Reaction through Combining Experiment and Theory. Angew. Chem. Int. Ed. 2015, 54, 52–65. [Google Scholar] [CrossRef]
  4. Xu, X.; Su, C.; Shao, Z. Fundamental Understanding and Application of Ba0.5Sr0.5Co0.8Fe0.2O3−δ Perovskite in Energy Storage and Conversion: Past, Present, and Future. Energy Fuels 2021, 35, 13585–13609. [Google Scholar] [CrossRef]
  5. Tang, J.; Xu, X.; Tang, T.; Zhong, Y.; Shao, Z. Perovskite-Based Electrocatalysts for Cost-Effective Ultrahigh-Current-Density Water Splitting in Anion Exchange Membrane Electrolyzer Cell. Small Methods 2022, 6, 2201099. [Google Scholar] [CrossRef]
  6. Jin, H.; Guo, C.; Liu, X.; Liu, J.; Vasileff, A.; Jiao, Y.; Zheng, Y.; Qiao, S.-Z. Emerging Two-Dimensional Nanomaterials for Electrocatalysis. Chem. Rev. 2018, 118, 6337–6408. [Google Scholar] [CrossRef]
  7. Liu, J.; Zhang, T.; Waterhouse, G.I.N. Complex alloy nanostructures as advanced catalysts for oxygen electrocatalysis: From materials design to applications. J. Mater. Chem. A 2020, 8, 23142–23161. [Google Scholar] [CrossRef]
  8. Liu, J.; Zhu, D.; Guo, C.; Vasileff, A.; Qiao, S.-Z. Design Strategies toward Advanced MOF-Derived Electrocatalysts for Energy-Conversion Reactions. Adv. Energy Mater. 2017, 7, 1700518. [Google Scholar] [CrossRef]
  9. Wang, A.; Li, J.; Zhang, T. Heterogeneous single-atom catalysis. Nat. Rev. Chem. 2018, 2, 65–81. [Google Scholar] [CrossRef]
  10. Chen, Y.; Ji, S.; Chen, C.; Peng, Q.; Wang, D.; Li, Y. Single-Atom Catalysts: Synthetic Strategies and Electrochemical Applications. Joule 2018, 2, 1242–1264. [Google Scholar] [CrossRef]
  11. Chen, Y.-N.; Zhang, X.; Zhou, Z. Carbon-Based Substrates for Highly Dispersed Nanoparticle and Even Single-Atom Electrocatalysts. Small Methods 2019, 3, 1900050. [Google Scholar] [CrossRef]
  12. Wu, J.; Zhou, H.; Li, Q.; Chen, M.; Wan, J.; Zhang, N.; Xiong, L.; Li, S.; Xia, B.Y.; Feng, G.; et al. Densely Populated Isolated Single Co–N Site for Efficient Oxygen Electrocatalysis. Adv. Energy Mater. 2019, 9, 1900149. [Google Scholar] [CrossRef]
  13. Wu, J.; Xiong, L.; Zhao, B.; Liu, M.; Huang, L. Densely Populated Single Atom Catalysts. Small Methods 2020, 4, 1900540. [Google Scholar] [CrossRef]
  14. Chen, S.; Luo, T.; Li, X.; Chen, K.; Fu, J.; Liu, K.; Cai, C.; Wang, Q.; Li, H.; Chen, Y.; et al. Identification of the Highly Active Co–N4 Coordination Motif for Selective Oxygen Reduction to Hydrogen Peroxide. J. Am. Chem. Soc. 2022, 144, 14505–14516. [Google Scholar] [CrossRef] [PubMed]
  15. Huang, J.; Zhang, Q.; Ding, J.; Zhai, Y. Fe–N–C single atom catalysts for the electrochemical conversion of carbon, nitrogen and oxygen elements. Mater. Rep. Energy 2022, 2, 100141. [Google Scholar] [CrossRef]
  16. Li, Y.; Wu, Z.-S.; Lu, P.; Wang, X.; Liu, W.; Liu, Z.; Ma, J.; Ren, W.; Jiang, Z.; Bao, X. High-Valence Nickel Single-Atom Catalysts Coordinated to Oxygen Sites for Extraordinarily Activating Oxygen Evolution Reaction. Adv. Sci. 2020, 7, 1903089. [Google Scholar] [CrossRef]
  17. Zhao, Y.; Guo, Y.; Lu, X.F.; Luan, D.; Gu, X.; Lou, X.W. Exposing Single Ni Atoms in Hollow S/N-Doped Carbon Macroporous Fibers for Highly Efficient Electrochemical Oxygen Evolution. Adv. Mater. 2022, 34, 2203442. [Google Scholar] [CrossRef]
  18. Fan, L.; Liu, P.F.; Yan, X.; Gu, L.; Yang, Z.Z.; Yang, H.G.; Qiu, S.; Yao, X. Atomically isolated nickel species anchored on graphitized carbon for efficient hydrogen evolution electrocatalysis. Nat. Commun. 2016, 7, 10667. [Google Scholar] [CrossRef]
  19. Zhao, Y.; Zhang, Z.; Liu, L.; Wang, Y.; Wu, T.; Qin, W.; Liu, S.; Jia, B.; Wu, H.; Zhang, D.; et al. S and O Co-Coordinated Mo Single Sites in Hierarchically Porous Tubes from Sulfur–Enamine Copolymerization for Oxygen Reduction and Evolution. J. Am. Chem. Soc. 2022, 144, 20571–20581. [Google Scholar] [CrossRef]
  20. Ling, C.; Ouyang, Y.; Li, Q.; Bai, X.; Mao, X.; Du, A.; Wang, J. A General Two-Step Strategy–Based High-Throughput Screening of Single Atom Catalysts for Nitrogen Fixation. Small Methods 2019, 3, 1800376. [Google Scholar] [CrossRef]
  21. Zang, W.; Sun, T.; Yang, T.; Xi, S.; Waqar, M.; Kou, Z.; Lyu, Z.; Feng, Y.P.; Wang, J.; Pennycook, S.J. Efficient Hydrogen Evolution of Oxidized Ni-N3 Defective Sites for Alkaline Freshwater and Seawater Electrolysis. Adv. Mater. 2021, 33, 2003846. [Google Scholar] [CrossRef] [PubMed]
  22. Zang, W.; Yang, T.; Zou, H.; Xi, S.; Zhang, H.; Liu, X.; Kou, Z.; Du, Y.; Feng, Y.P.; Shen, L.; et al. Copper Single Atoms Anchored in Porous Nitrogen-Doped Carbon as Efficient pH-Universal Catalysts for the Nitrogen Reduction Reaction. ACS Catal. 2019, 9, 10166–10173. [Google Scholar] [CrossRef]
  23. Wang, X.; Chen, Z.; Zhao, X.; Yao, T.; Chen, W.; You, R.; Zhao, C.; Wu, G.; Wang, J.; Huang, W.; et al. Regulation of Coordination Number over Single Co Sites: Triggering the Efficient Electroreduction of CO2. Angew. Chem. Int. Ed. 2018, 57, 1944–1948. [Google Scholar] [CrossRef] [PubMed]
  24. Yuan, K.; Lützenkirchen-Hecht, D.; Li, L.; Shuai, L.; Li, Y.; Cao, R.; Qiu, M.; Zhuang, X.; Leung, M.K.H.; Chen, Y.; et al. Boosting Oxygen Reduction of Single Iron Active Sites via Geometric and Electronic Engineering: Nitrogen and Phosphorus Dual Coordination. J. Am. Chem. Soc. 2020, 142, 2404–2412. [Google Scholar] [CrossRef] [PubMed]
  25. Jiang, X.-H.; Zhang, L.-S.; Liu, H.-Y.; Wu, D.-S.; Wu, F.-Y.; Tian, L.; Liu, L.-L.; Zou, J.-P.; Luo, S.-L.; Chen, B.-B. Silver Single Atom in Carbon Nitride Catalyst for Highly Efficient Photocatalytic Hydrogen Evolution. Angew. Chem. Int. Ed. 2020, 59, 23112–23116. [Google Scholar] [CrossRef] [PubMed]
  26. Cheng, C.C.; Yeh, Y.X.; Ting, Y.C.; Lin, S.H.; Sasaki, K.; Choi, Y.; Lu, S.Y. Modulation of the coordination environment enhances the electrocatalytic efficiency of Mo single atoms toward water splitting. J. Mater. Chem. A 2022, 10, 8784–8797. [Google Scholar] [CrossRef]
  27. Niu, H.; Wan, X.; Wang, X.; Shao, C.; Robertson, J.; Zhang, Z.; Guo, Y. Single-Atom Rhodium on Defective g-C3N4: A Promising Bifunctional Oxygen Electrocatalyst. ACS Sustain. Chem. Eng. 2021, 9, 3590–3599. [Google Scholar] [CrossRef]
  28. Wang, S.; Huang, B.; Dai, Y.; Wei, W. Tuning the Coordination Microenvironment of Central Fe Active Site to Boost Water Electrolysis and Oxygen Reduction Activity. Small 2023, 19, 2205111. [Google Scholar] [CrossRef]
  29. Su, H.; Soldatov, M.A.; Roldugin, V.; Liu, Q. Platinum single-atom catalyst with self-adjustable valence state for large-current-density acidic water oxidation. eScience 2022, 2, 102–109. [Google Scholar] [CrossRef]
  30. Xue, Z.; Zhang, X.; Qin, J.; Liu, R. TMN4 complex embedded graphene as bifunctional electrocatalysts for high efficiency OER/ORR. J. Energy Chem. 2021, 55, 437–443. [Google Scholar] [CrossRef]
  31. Xiao, M.; Zhu, J.; Li, G.; Li, N.; Li, S.; Cano, Z.P.; Ma, L.; Cui, P.; Xu, P.; Jiang, G.; et al. A Single-Atom Iridium Heterogeneous Catalyst in Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2019, 58, 9640–9645. [Google Scholar] [CrossRef] [PubMed]
  32. Li, X.; Su, Z.; Zhao, Z.; Cai, Q.; Li, Y.; Zhao, J. Single Ir atom anchored in pyrrolic-N4 doped graphene as a promising bifunctional electrocatalyst for the ORR/OER: A computational study. J. Colloid Interface Sci. 2022, 607, 1005–1013. [Google Scholar] [CrossRef] [PubMed]
  33. Han, L.; Cheng, H.; Liu, W.; Li, H.; Ou, P.; Lin, R.; Wang, H.-T.; Pao, C.-W.; Head, A.R.; Wang, C.-H.; et al. A single-atom library for guided monometallic and concentration-complex multimetallic designs. Nat. Mater. 2022, 21, 681–688. [Google Scholar] [CrossRef] [PubMed]
  34. Lv, X.; Wei, W.; Huang, B.; Dai, Y.; Frauenheim, T. High-Throughput Screening of Synergistic Transition Metal Dual-Atom Catalysts for Efficient Nitrogen Fixation. Nano Lett. 2021, 21, 1871–1878. [Google Scholar] [CrossRef]
  35. Cao, N.; Zhang, N.; Wang, K.; Yan, K.; Xie, P. High-throughput screening of B/N-doped graphene supported single-atom catalysts for nitrogen reduction reaction. Chem. Synth. 2023, 3. [Google Scholar] [CrossRef]
  36. Janthon, P.; Kozlov, S.M.; Viñes, F.; Limtrakul, J.; Illas, F. Establishing the Accuracy of Broadly Used Density Functionals in Describing Bulk Properties of Transition Metals. J. Chem. Theory Comput. 2013, 9, 1631–1640. [Google Scholar] [CrossRef]
  37. Rossmeisl, J.; Qu, Z.W.; Zhu, H.; Kroes, G.J.; Nørskov, J.K. Electrolysis of water on oxide surfaces. J. Electroanal. Chem. 2007, 607, 83–89. [Google Scholar] [CrossRef]
  38. Tripković, V.; Cerri, I.; Bligaard, T.; Rossmeisl, J. The Influence of Particle Shape and Size on the Activity of Platinum Nanoparticles for Oxygen Reduction Reaction: A Density Functional Theory Study. Catal. Lett. 2014, 144, 380–388. [Google Scholar] [CrossRef]
  39. Bajdich, M.; García-Mota, M.; Vojvodic, A.; Nørskov, J.K.; Bell, A.T. Theoretical Investigation of the Activity of Cobalt Oxides for the Electrochemical Oxidation of Water. J. Am. Chem. Soc. 2013, 135, 13521–13530. [Google Scholar] [CrossRef]
  40. Valdés, Á.; Qu, Z.W.; Kroes, G.J.; Rossmeisl, J.; Nørskov, J.K. Oxidation and Photo-Oxidation of Water on TiO2 Surface. J. Phys. Chem. C 2008, 112, 9872–9879. [Google Scholar] [CrossRef]
  41. Ren, Y.; Tang, Y.; Zhang, L.; Liu, X.; Li, L.; Miao, S.; Sheng, S.; Wang, A.; Li, J.; Zhang, T. Unraveling the coordination structure-performance relationship in Pt1/Fe2O3 single-atom catalyst. Nat. Commun. 2019, 10, 4500. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, J.; Xiao, J.; Luo, B.; Tian, E.; Waterhouse, G.I.N. Central metal and ligand effects on oxygen electrocatalysis over 3d transition metal single-atom catalysts: A theoretical investigation. Chem. Eng. J. 2022, 427, 132038. [Google Scholar] [CrossRef]
  43. Medford, A.J.; Vojvodic, A.; Hummelshøj, J.S.; Voss, J.; Abild-Pedersen, F.; Studt, F.; Bligaard, T.; Nilsson, A.; Nørskov, J.K. From the Sabatier principle to a predictive theory of transition-metal heterogeneous catalysis. J. Catal. 2015, 328, 36–42. [Google Scholar] [CrossRef]
  44. Deringer, V.L.; Tchougréeff, A.L.; Dronskowski, R. Crystal orbital Hamilton population (COHP) analysis as projected from plane-wave basis sets. J. Phys. Chem. A 2011, 115, 5461–5466. [Google Scholar] [CrossRef] [PubMed]
  45. Yan, T.; Li, X.; Wang, Z.; Cai, Q.; Zhao, J. Interface engineering of transition metal-nitrogen-carbon by graphdiyne for boosting the oxygen reduction/evolution reactions: A computational study. J. Colloid Interface Sci. 2023, 649, 1–9. [Google Scholar] [CrossRef]
  46. Wang, E.; Zhang, B.; Zhou, J.; Sun, Z. High catalytic activity of MBenes-supported single atom catalysts for oxygen reduction and oxygen evolution reaction. Appl. Surf. Sci. 2022, 604, 154522. [Google Scholar] [CrossRef]
  47. Yang, C.; Wu, Y.; Wang, Y.; Zhang, H.-N.; Zhu, L.-H.; Wang, X.-C. Electronic properties of double-atom catalysts for electrocatalytic oxygen evolution reaction in alkaline solution: A DFT study. Nanoscale 2022, 14, 187–195. [Google Scholar] [CrossRef]
  48. Ling, C.; Shi, L.; Ouyang, Y.; Zeng, X.C.; Wang, J. Nanosheet Supported Single-Metal Atom Bifunctional Catalyst for Overall Water Splitting. Nano Lett. 2017, 17, 5133–5139. [Google Scholar] [CrossRef]
  49. Fu, Z.; Ling, C.; Wang, J. A Ti3C2O2 supported single atom, trifunctional catalyst for electrochemical reactions. J. Mater. Chem. A 2020, 8, 7801–7807. [Google Scholar] [CrossRef]
  50. Liu, F.; Song, L.; Liu, Y.; Zheng, F.; Wang, L.; Palotás, K.; Lin, H.; Li, Y. Using the N≡N dipole as a theoretical indicator for estimating the electrocatalytic performance of active sites in the nitrogen reduction reaction: Single transition metal atoms embedded in two dimensional phthalocyanine. J. Mater. Chem. A 2020, 8, 3598–3605. [Google Scholar] [CrossRef]
  51. Zhang, Q.; Asthagiri, A. Solvation Effects on DFT Predictions of ORR Activity on Metal Surfaces. Catal. Today 2019, 323, 35–43. Available online: https://www.sciencedirect.com/science/article/pii/S0920586118310885 (accessed on 7 September 2018). [CrossRef]
  52. Peng, Y.; Hajiyani, H.; Pentcheva, R. Influence of Fe and Ni Doping on the OER Performance at the Co3O4(001) Surface: Insights from DFT+U Calculations. ACS Catal. 2021, 11, 5601–5613. [Google Scholar] [CrossRef]
  53. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  54. Kresse, G.; Hafner, J.J.P.R.B. Ab initio molecular dynamics for open-shell transition metals. Phys. Rev. B 1993, 48, 13115–13118. [Google Scholar] [CrossRef]
  55. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  56. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  57. Perdew, J.P.; Burke, K.; Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B 1996, 54, 16533–16539. [Google Scholar] [CrossRef]
  58. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, molecules, solids, and surfaces: Applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef]
  59. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  60. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  61. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  62. Mathew, K.; Sundararaman, R.; Letchworth-Weaver, K.; Arias, T.A.; Hennig, R.G. Implicit solvation model for density-functional study of nanocrystal surfaces and reaction pathways. J. Chem. Phys. 2014, 140, 084106. [Google Scholar] [CrossRef]
  63. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  64. Man, I.C.; Su, H.-Y.; Calle-Vallejo, F.; Hansen, H.A.; Martínez, J.I.; Inoglu, N.G.; Kitchin, J.; Jaramillo, T.F.; Nørskov, J.K.; Rossmeisl, J. Universality in Oxygen Evolution Electrocatalysis on Oxide Surfaces. ChemCatChem 2011, 3, 1159–1165. [Google Scholar] [CrossRef]
Figure 1. Top and side views of optimized structures of (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C. Ir, Ru, C, N, and O atoms are presented with blue, purple, brown, yellow, and red circles, respectively.
Figure 1. Top and side views of optimized structures of (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C. Ir, Ru, C, N, and O atoms are presented with blue, purple, brown, yellow, and red circles, respectively.
Catalysts 13 01378 g001
Figure 2. Calculated binding energies of single M atoms in MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Figure 2. Calculated binding energies of single M atoms in MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Catalysts 13 01378 g002
Scheme 1. Schematic of the 4e process of OER on the MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Scheme 1. Schematic of the 4e process of OER on the MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Catalysts 13 01378 sch001
Figure 3. Gibbs free energy diagrams of the OER processes on (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C. H, C, O, N, Ir, and Ru atoms are represented with white, brown, red, yellow, blue, and purple circles.
Figure 3. Gibbs free energy diagrams of the OER processes on (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C. H, C, O, N, Ir, and Ru atoms are represented with white, brown, red, yellow, blue, and purple circles.
Catalysts 13 01378 g003
Figure 4. The scaling relationship between (a) ΔGOOH* vs. ΔGOH*, and the volcano plot of (b) − η OER vs. ΔGO* on MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Figure 4. The scaling relationship between (a) ΔGOOH* vs. ΔGOH*, and the volcano plot of (b) − η OER vs. ΔGO* on MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Catalysts 13 01378 g004
Figure 5. Projected crystal orbital Hamilton populations (COHP) between chemisorbed O* and active metal centers on (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C. The Fermi levels were set to the energy zero.
Figure 5. Projected crystal orbital Hamilton populations (COHP) between chemisorbed O* and active metal centers on (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C. The Fermi levels were set to the energy zero.
Catalysts 13 01378 g005
Figure 6. The CDD of (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C. The electron accumulation and depletion are represented with cyan and pink contours, with iso-surfaces being 0.015 e/Å3. The net charge of M atoms of MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Figure 6. The CDD of (a) IrN4-C, (b) IrO4-C, (c) IrC4-C, (d) RuN4-C, (e) RuO4-C, and (f) RuC4-C. The electron accumulation and depletion are represented with cyan and pink contours, with iso-surfaces being 0.015 e/Å3. The net charge of M atoms of MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Catalysts 13 01378 g006
Figure 7. Partial density of states (PDOS) of d orbitals for M atoms on (a) IrN4-C, IrO4-C and IrC4-C, (b) RuN4-C, RuO4-C and RuC4-C. The d-band centers (εd) are also labeled for metal atoms. The Fermi level (Ef) is set to 0.00 eV.
Figure 7. Partial density of states (PDOS) of d orbitals for M atoms on (a) IrN4-C, IrO4-C and IrC4-C, (b) RuN4-C, RuO4-C and RuC4-C. The d-band centers (εd) are also labeled for metal atoms. The Fermi level (Ef) is set to 0.00 eV.
Catalysts 13 01378 g007
Table 1. The adsorption-free energy of OH*, O*, and OOH* for MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
Table 1. The adsorption-free energy of OH*, O*, and OOH* for MN4-C, MO4-C, and MC4-C (M = Ir, Ru).
SystemsIrN4-CIrO4-CIrC4-CRuN4-CRuO4-CRuC4-C
ΔGOH* (eV)0.961.72−0.42−0.010.40−0.26
ΔGO* (eV)2.492.940.010.671.52−0.08
ΔGOOH* (eV)3.974.682.662.893.372.88
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, J.; Bai, J.; Cang, Y.; Li, Q.; Fan, X.; Lin, H. Noble Metal Single-Atom Coordinated to Nitrogen, Oxygen, and Carbon as Electrocatalysts for Oxygen Evolution. Catalysts 2023, 13, 1378. https://doi.org/10.3390/catal13101378

AMA Style

Wang J, Bai J, Cang Y, Li Q, Fan X, Lin H. Noble Metal Single-Atom Coordinated to Nitrogen, Oxygen, and Carbon as Electrocatalysts for Oxygen Evolution. Catalysts. 2023; 13(10):1378. https://doi.org/10.3390/catal13101378

Chicago/Turabian Style

Wang, Jianhua, Jiangdong Bai, Yaqi Cang, Qing Li, Xing Fan, and Haiping Lin. 2023. "Noble Metal Single-Atom Coordinated to Nitrogen, Oxygen, and Carbon as Electrocatalysts for Oxygen Evolution" Catalysts 13, no. 10: 1378. https://doi.org/10.3390/catal13101378

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop