Next Article in Journal
Preparation of Cordierite Monolith Catalysts with the Coating of K-Modified Spinel MnCo2O4 Oxide and Their Catalytic Performances for Soot Combustion
Previous Article in Journal
Polyaromatic Carboxylate Ligands Based Zn(II) Coordination Polymers for Ultrasound-Assisted One-Pot Tandem Deacetalization–Knoevenagel Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

β-Arsenene Monolayer: A Promising Electrocatalyst for Anodic Chlorine Evolution Reaction

1
Centre for Catalysis and Clean Energy, School of Environment and Science, Griffith University, Gold Coast, QLD 4222, Australia
2
Institute of Solid-State Physics, Chinese Academy of Sciences, Hefei 230031, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(3), 296; https://doi.org/10.3390/catal12030296
Submission received: 18 January 2022 / Revised: 2 March 2022 / Accepted: 3 March 2022 / Published: 5 March 2022
(This article belongs to the Special Issue Two-Dimensional Materials for Electrocatalysis)

Abstract

:
Materials innovation plays an essential role to address the increasing demands of gaseous chlorine from anodic chlorine evolution reaction (CER) in chlor-alkali electrolysis. In this study, two-dimensional (2D) semiconducting group-VA monolayers were theoretically screened for the electrochemical CER by means of the density functional theory (DFT) method. Our results reveal the monolayered β-arsenene has the ultralow thermodynamic overpotential of 0.068 V for CER, which is close to that of the commercial Ru/Ir-based dimensionally stable anode (DSA) of 0.08 V @ 10 mA cm−2 and 0.13 V from experiments and theory, respectively. The change of CER pathways via Cl* intermediate on 2D β-arsenene also efficiently suppresses the parasitical oxygen gas production because of a high theoretical oxygen evolution reaction (OER) overpotential of 1.95 V. Our findings may therefore expand the scope of the electrocatalysts design for CER by using emerging 2D materials.

Graphical Abstract

1. Introduction

The chlor-alkali process as the primary means for chlorine (Cl2) manufacture is one of the largest industrial electrochemical technologies [1,2]. Electrocatalysis is the heart of the cost-intensive chlor-alkali industry since it has the demonstrated capacity for a series of energy-related applications including chlorine evolution reaction (CER), oxygen reduc-tion reaction (ORR), and hydrogen evolution reaction (HER) [3,4,5,6]. The anodic CER can be achieved through electrolyzing a concentrated brine solution by applying a direct electric current (2Cl → Cl2 + 2e, UCER = 1.36 V vs. SHE) [3,7]. The development of a high-performance electrocatalyst for CER is essential for its commercialization. Mixed-metal oxides (MMOs) based on noble metals Ru or Ir, such as dimensionally stable anodes (DSAs), have been predominantly used as CER catalysts [8,9]. For example, the TiO2-sup-ported mixtures of RuO2 and IrO2 are the most commonly used electrocatalysts in indus-trial chlorine processes [10,11]. However, the DSA catalysts have a high oxygen evolution reaction (OER) activity [1,12,13,14,15,16]. This is because the OER (2H2O → O2 + 2H+ + 2e, UOER = 1.23 V vs. SHE) is competitive at the anode under typical reaction conditions [9,13,17,18]. In the same potential window of the CER, the concomitant O2 reduces CER selectivity [19]. As such, the purification of gaseous Cl2 with the high selectivity of CER becomes a major challenge in chlor-alkali electrolysis. To solve the encountered selectivity issue, one of the most promising strategies is the acidification of electrolyte solutions, as the OER perfor-mance of most electrode materials can be suppressed in an acidic media, while the reversible electrode potential of CER is pH-independent [9,19]. To this end, the screening of novel high-performance CER with high reactivity and selectivity in acidic solution be-comes imperative.
To date, two-dimensional (2D) layered nanostructures have attracted increasing at-tention using as electrocatalysts, since they expose more active sites and achieve much higher catalytic efficiency [20,21,22,23,24,25,26,27]. Recently, the successful fabrication of few-layer black phosphorus brings group-VA elements (P, As, Sb, and Bi) into the family of 2D materials and inspires research interest on other layered allotropes [28]. For example, Bat-Erdene et al. reported that 2D antimonene nanosheets is an efficient electrocatalyst for the nitrogen reduction reaction (NRR) [29]. Additionally, Ren et al. adopt a favorable liquid exfoliation approach to produce few-layer antimonene and implement a metal-free electrocatalyst for water splitting [30]. Thus, the 2D semiconductor family, composed of group-VA elements, has a great potential for electrocatalysis. Yet, the investigations of this class of materials for the CER are still rare.
In this study, the density functional theory (DFT) was employed to comparatively investigate the CER electrocatalytic performance of 2D group-VA monolayers including phosphorene, arsenene, antimonene, and bismuthene. Here, considering the energetic sta-bility and possible fabrication in experiments, we mainly focus on the α and β phases of group-VA monolayers [31,32,33]. Our DFT results predict that the 2D β-arsenene monolayer is a promising candidate for CER with the ultralow thermodynamic overpotential of 0.068 V and high selectivity in terms of OER.

2. Results and Discussion

2.1. Cl Adsorption on Group-VA Monolayers

The atomic structures of α and β phases of group-VA monolayers are shown in Figure 1. Their structural parameters are listed in Table 1. Group-VA monolayers with α and β allotropes possess honeycomb structures, where α and β phases are derived from orthorhombic and rhombohedral bulk structures, respectively. The lattice constants and layer thicknesses of all studied group-VA monolayers increase from P to Bi regardless of phase because of the increased atomic radius. For four monolayers with α phase, there are two types of bonds, namely in-plane bond (Bond 1) and out-plane bond (Bond 2), as shown in Figure 1a. It is found that bond lengths of both types gradually increase along the periodic table. For another allotrope β phases, only one type of bond, X−X, can be observed with bond lengths increasing from β-phosphorene (2.26 Å) to β-bismuthene (3.04 Å). All these structural parameters are consistent with reported data in previous studies [31,32].
Previous theoretical and experimental studies of the electrocatalytic CER over low-dimensional anodes suggest that the formation of the Cl* intermediate instead of ClO* intermediate would significantly promote the chlorine generation with high selectivity [16,17,18,34]. Herein, to evaluate the CER performance of group-VA 2D monolayers, the adsorption properties of Cl atoms on both α and β phases were first investigated, and the results are listed in Table 2. As illustrated in Figure 1, one adsorption site (top) of α phase and three different adsorption sites including Site 1 (top), Site 2 (hollow), and Site 3 (hollow) of β phase were considered here to examine the regioselectivity of Cl atoms. It was found that the adsorption on the Site 2 of β-antimonene and Site 3 of β-phosphorene and β-arsenene is unpreferred because the adsorbed Cl atom would migrate to the Site 1 after structural optimization. Moreover, for 2D β-phosphorene and β-arsenene, the Cl adsorptions at Site 1 are much less positive than that of Site 2, indicating that Site 1 is thermodynamically preferred for chlorine adsorption. As a comparison, the Gibbs free energies of Cl* at the Site 3 on β-antimonene and β-bismuthene are much more negative than others, which implies that this hollow site is energetically favored for adsorbing Cl atoms.

2.2. CER Activity of Group-VA Monolayers

The electrocatalytic CER is a two-electron process through the Volmer-Heyrovsky mechanism [12,34]. By using the Cl* intermediate, this mechanism can be explained by that Cl* forms first via the adsorption and discharge of a chloride anion in the Volmer step (* + 2Cl → Cl* + Cl + e). Thereafter, the Cl* intermediate directly combines with another chloride anion from the electrolyte solution to release gaseous chlorine in the Heyrovsky step (Cl* + Cl + e → * + Cl2 + 2e). Figure 2 depicts the Gibbs free energy changes for CER with the most stable configurations of all monolayers with α and β phases at the equilibrium potential of 1.36 V according to Volmer and Heyrovsky steps. The theoretical overpotential for CER (ηCER) can be defined by the ΔGCl*, namely ηCER = |ΔGCl*|/e. As shown in Figure 2, the interactions between Cl atoms and α-monolayers are relatively stronger than that of β-monolayers due to the much more negative values of ΔGCl*. In the case of α-monolayers, the energy wells of α-arsenene, α-antimonene, and α-bismuthene indicate that the Cl* can form spontaneously. However, the interactions between As, Sb, and Bi and Cl atoms are too strong for efficient Cl2 desorption. Consequently, the formation of Cl2 in the Heyrovsky step is energy-demanding for α-arsenene, α-antimonene, and α-bismuthene with the thermodynamic overpotentials of −0.232 V, −0.856 V, and −0.966 V, respectively. As a comparison, the small energy barrier for α-phosphorene implies that the formation of Cl* in the Volmer step is more energy-demanding, leading to a theoretical overpotential of 0.055 V. It demonstrates that the CER mechanisms on α-group VA monolayers can be adjusted by using different elements, and α-phosphorene possesses the highest activity for Cl2 evolution among these materials.
In addition, the CER performances of β-monolayers are quite distinct from α-monolayers due to the totally different atomic structures. Specifically, β-phosphorene has relatively low activity ascribed to its weak binding ability with Cl, whereas β-antimonene and β-bismuthene are anticipated to possess low activity owing to the overly strong adsorption of Cl*. It is notable that β-arsenene has a moderate ΔGCl* of 0.068 eV to compromise the reaction barriers in the Cl adsorption and desorption steps, which is beneficial to generate Cl2 gas. As reported by Exner et al., the CER overpotential of the traditional single-crystalline RuO2(110) electrocatalyst is 0.13 V vs. SHE via the ClO* precursor [12,14,35]. For better comparison, a summary table including the CER performance for reported electrocatalysts is listed in Table 3. Therefore, our DFT results suggest that the monolayered α-phosphorene and β-arsenene exhibit comparable CER activity compared to that of benchmark RuO2(110) electrocatalyst.
To deeply understand the CER performances of different group-VA monolayers with α and β phases, the bonding mechanism between the adsorbed Cl and P, As, Sb, or Bi atoms was investigated through the COHP and Mulliken charge analyses. The corresponding COHP images are shown in Figure 3, where covalent bonding and antibonding states are characterized by the positive and negative overlap population, respectively. For obtaining the quantitative description of covalent bonding strength between P, As, Sb, or Bi and Cl atoms, the integral of the −pCOHP (−IpCOHP) up to Fermi energy level was also calculated, shown in Figure 3. The more negative −IpCOHP value suggests that the corresponding active site is more reactive towards the adsorption of Cl via the covalent bonding. However, our results reveal that there is an inverse relationship between the −IpCOHP and ΔGCl*, suggesting that the covalent bonding strength is not a reasonable descriptor for CER activity of α and β phases of group-VA monolayers. Therefore, we fur-ther investigate their ionic bonding strength, which can be described by the electrostatic attractions (Fes) between the P, As, Sb, or Bi and Cl atoms, as well as their bond length, as follows:
F es = K | q X   ×   q Cl | r Cl - X 2
where K is Coulomb’s constant, and qX and qCl are the Mulliken charges of group VA element X (X = P, As, Sb, or Bi) and Cl, respectively, and the rCl-X is the bond length between Cl and X atoms. As listed in Table 4, the trend of −IpCOHP values follows the rule of P < As < Sb < Bi for all α- and β-monolayers, agreeing with the corresponding binding affinity of Cl atom. It demonstrates that the electrostatic attraction between group VA elements and Cl can be a reliable descriptor to predict the CER performance of all studied α- and β-monolayers, which is in agreement with our previous study [42].
To clearly visualize the X−Cl interaction, the electron localization function (ELF) of α- and β-monolayers with the Cl* was calculated, as displayed in Figure 4. The color denotes the renormalized ELF values, with the values 1.0 and 0.5 representing fully localized and fully delocalized electrons, respectively, while the value 0 means very low charge density. Since the electrons are gradually highly delocalized, it is found that the ionic bonding characters between X and Cl atoms increase with the increase of the atomic number. This can be ascribed to the gradually enhanced metallicity from P to Bi atoms, resulting from their decreased electronegativity. The atom with relatively strong metallicity would give rise to much stronger X−Cl interaction, such as Sb and Bi, which is detrimental to the desorption of Cl2 gas. In contrast, the monolayers comprised of the non-metallic atom, such as α-phosphorene and β-arsenene, have the weaker X−Cl interaction to benefit the Cl desorption, resulting in the relatively high activity of Cl2 generation.

2.3. CER Selectivity

The evolution of gaseous oxygen at the anode is more thermodynamically preferred on account of the relatively lower potential UOER of 1.23 V vs. SHE, which regrettably causes the selectivity problem of Cl2 gas production at the anode. Therefore, it should combine the high activity with selectivity in an acidic electrolyte for a desired CER electrocatalyst. Since the α-phosphorene and β-arsenene are predicted to have the best performance towards Cl2 generation, the selectivity of these two monolayers for CER is further analyzed. According to the CHE method, the OER proceeds in the four-electron transfer steps, as follows:
* + H2O → HO* + H+ + e
HO* → O* + H+ + e
O* + H2O → HOO* + H+ + e
HOO* → * + O2 + H+ + e
During the OER process, intermediates of HO*, O*, and HOO* are formed in turn following corresponding elementary steps. In this study, we used the adsorption free energy of HO* intermediate forming in the first elementary step of OER to evaluate the se-lectivity for α-phosphorene and β-arsenene monolayers. Figure 5a shows the Gibbs free energy changes of HO* and Cl* species of α-phosphorene and β-arsenene as a function of applied potential USHE to determine the more energetically stable structures at pH = 0. It is observed that the value of ΔGHO* of α-phosphorene monolayer is smaller than that of ΔGCl*, implying that the formation of HO* intermediate on the α-phosphorene is more energetically favorable in acidic solution. This indicates that the OER process is preferred on the α-phosphorene monolayer at the potential region of UCER, leading to a poor selectivity of Cl2. As a comparison, the value of ΔGCl* of β-arsenene monolayer is much lower than that of ΔGHO*, suggesting that the CER process priors to the OER with the more thermodynamically favorable Cl* precursor.
To better understand the selectivity of β-arsenene 2D monolayer towards CER, the free energy diagram of the detrimental OER was also calculated, as shown in Figure 5b. The theoretical overpotential of OER is derived from the most energy-demanding step of four elementary reactions Equations (2)–(5). The reaction free energy of Equations (2)–(5) for OER can be defined as follows:
ΔG1 = ΔGHO*
ΔG2 = ΔGO* − ΔGHO*
ΔG3 = ΔGHOO* − ΔGO*
ΔG4 = 4.92 − ΔGHOO*
Therefore, the ηOER can be calculated by
η OER = max   [ Δ G 1 ,   Δ G 2 ,   Δ G 3 ,   Δ G 4 ] e   1.23
The formation of HOO* intermediate from O* is found to be the most endothermic step, resulting in the potential determining step (PDS) with an extremely large theoretical overpotential of 2.08 V at UOER = 1.23 V. To be referenced to the same potential of CER (UCER = 1.36 V), the corrected theoretical overpotential of OER for β-arsenene monolayer is 1.95 V, i.e., ηOER = 2.08 – (1.36 – 1.23) = 1.95 V. It reveals that the OER is efficiently suppressed on the 2D β-arsenene monolayer even in the absence of Cl. In comparison, previous studies demonstrate that the widely used CER electrocatalyst RuO2(110) is also highly active for oxygen evolution [43,44]. Our DFT results show that the thermodynamic OER overpotential of β-arsenene monolayer is five times larger than that of RuO2(110) [12], indicating that the selectivity problem can be solved by using β-arsenene monolayer as the CER electrocatalysts. Such an improved selectivity of Cl2 generation can be ascribed to the inherently different pathways for CER. Unlike RuO2(110), by virtue of the ClO* intermediate, the CER on β-arsenene monolayer is via Cl* species, which is beneficial to the selective chlorine evolution.

3. Computational Methods

All first principles DFT calculations were performed using the Vienna Ab initio Simulation Package (VASP) based on the projector-augmented wave (PAW) method [45,46]. The exchange-correlation energy was treated by the Perdew-Burke-Ernzerhof (PBE) functional at the generalized gradient approximation (GGA) level [47]. The electron-ion interaction was described using the PAW potentials [48], 3s23p3 for P, 4s24p3 for As, 5s25p3 for Sb, 6s25d106p3 for Bi, 3s23p5 for Cl, 2s22p4 for O, and 1s1 for H, respectively. In order to incorporate the effects of nonlocal van der Waals interactions that are not included correctly in conventional DFT calculations, the DFT-D3 method was adopted for dispersion corrections here [49,50,51,52,53]. A plane-wave basis set with the cut-off kinetic energy of 520 eV is employed to expand the smooth part of wave functions. The gamma-centered k-point meshes with a reciprocal space resolution of 2π × 0.03 Å−1 and 2π × 0.02 Å−1 were utilized for structural optimization and static self-consistent calculations, respectively. To model the α and β phases of group-VA monolayers in this study, a (3 × 3) supercell was adopted, including 36 and 18 group-VA element atoms for α and β phases monolayers, respectively. A 20 Å vacuum along z-direction was applied to prevent spurious interaction between the periodically repeated images. All atoms were allowed to relax until the Hellmann−Feynman forces were smaller than 0.01 eV/Å, and the convergence criterion for the electronic self-consistent loop was set to 10−5 eV. The projected crystal orbital Hamilton population (pCOHP) method was used via the LOBSTER program to understand the chemical bonding between group-VA elements (P, As, Sb or Bi) and adsorbed Cl atoms [54,55,56,57,58].
The adsorption energy (ΔE) of all considered adsorbates (i.e., Cl*, HO*, O*, and HOO*, * refers to the corresponding catalytic site) can be calculated by following equations:
* + H2O → O* + 2(H+ + e)
* + H2O → HO* + H+ + e
* + 2H2O → HOO* + 3(H+ + e)
* + Cl → Cl* + e
Hence, ΔE for each species was calculated as follows:
Δ E O * = E O * + E H 2   E E H 2 O
Δ E HO * = E HO * + 0.5 E H 2   E E H 2 O
Δ E HOO * = E HOO * + 1.5 E H 2   E 2 E H 2 O
Δ E Cl * = E Cl *   E   0.5 E Cl 2  
The free energy change (ΔG) of the considered adsorbates (i.e., Cl*, HO*, O*, and HOO*) was obtained by using the computational hydrogen electrode (CHE) method [59], which can be explained as follows:
ΔG = ΔE + ΔZPE − TΔS
where ΔZPE is the change in zero-pint vibrational energy and −TΔS is the entropy contribution at room temperature. In this study, we considered the standard conditions of a Cl   = 1 and pH = 0. Therefore, the ΔG for each species as a function of applied potential USHE can be defined as follows at 298 K:
ΔG (U) = ΔE + ΔZPE − TΔS − ν(e)·e·USHE − ν(Cl)·e·UCl
where ν(e), and ν(Cl) are the values for stoichiometric coefficient of transferred electrons and chloride of corresponding adsorption intermediates, respectively; UCl is the standard potential of the reversible chlorine electrode (1.36 V at 298 K); and USHE denotes the applied electrode potential referenced to standard hydrogen electrode (SHE).

4. Conclusions

In summary, the first-principle DFT calculations were performed to investigate the CER performance of 2D α and β phases of group-VA monolayers. Our calculated results reveal that β-arsenene monolayer exhibits high activity and selectivity of gaseous Cl2 generation by virtue of the expected Cl* precursor, with the thermodynamic overpotential of 0.068 V and 1.95 V for the CER and OER, respectively. This 2D β-arsenene monolayered catalyst may therefore be a promising candidate for CER in the acidic medium. Moreover, our results found that the COHP analyses fail to predict the CER performance of these α and β phases of group-VA monolayers, since the covalent bonding state between adsorbent and Cl atoms becomes weaker from P to Bi. The electrostatic attraction between adsorbent and Cl atoms is a better descriptor for these systems. The theoretical prediction of this study may broaden the scope of CER electrocatalysts design using 2D materials.

Author Contributions

Methodology, software, formal analysis, writing—original draft preparation, J.L.; writing—review and editing, J.J.H.; writing—review and editing, H.Y.; writing—review and editing, P.L.; project administration, writing—review and editing, H.Z.; project administration, supervision, conceptualization, methodology, resources, writing—review and editing, Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This research was undertaken on the supercomputers in National Computational Infrastructure (NCI) in Canberra, Australia, which is supported by the Australian Commonwealth Government, and Pawsey Supercomputing Centre in Perth with funding from the Australian government and the Government of Western Australia.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sohrabnejad-Eskan, I.; Goryachev, A.; Exner, K.S.; Kibler, L.A.; Hensen, E.J.M.; Hofmann, J.P.; Over, H. Temperature-Dependent Kinetic Studies of the Chlorine Evolution Reaction over RuO2(110) Model Electrodes. ACS Catal. 2017, 7, 2403–2411. [Google Scholar] [CrossRef]
  2. Karlsson, R.K.; Cornell, A. Selectivity between Oxygen and Chlorine Evolution in the Chlor-Alkali and Chlorate Processes. Chem. Rev. 2016, 116, 2982–3028. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, Y.; Liu, Y.; Wiley, D.; Zhao, S.; Tang, Z. Recent advances in electrocatalytic chloride oxidation for chlorine gas production. J. Mater. Chem. A 2021, 9, 18974–18993. [Google Scholar] [CrossRef]
  4. Bhalothia, D.; Lin, C.-Y.; Yan, C.; Yang, Y.-T.; Chen, T.-Y. Effects of Pt metal loading on the atomic restructure and oxygen reduction reaction performance of Pt-cluster decorated Cu@Pd electrocatalysts. Sustain. Energy Fuels 2019, 3, 1668–1681. [Google Scholar] [CrossRef]
  5. Bhalothia, D.; Krishnia, L.; Yang, S.-S.; Yan, C.; Hsiung, W.-H.; Wang, K.-W.; Chen, T.-Y. Recent Advancements and Future Prospects of Noble Metal-Based Heterogeneous Nanocatalysts for Oxygen Reduction and Hydrogen Evolution Reactions. Appl. Sci. 2020, 10, 7708. [Google Scholar] [CrossRef]
  6. Bhalothia, D.; Huang, T.-H.; Chang, C.-W.; Lin, T.-H.; Wu, S.-C.; Wang, K.-W.; Chen, T.-Y. High-Performance and Stable Hydrogen Evolution Reaction Achieved by Pt Trimer Decoration on Ultralow-Metal Loading Bimetallic PtPd Nanocatalysts. ACS Appl. Energy Mater. 2020, 3, 11142–11152. [Google Scholar] [CrossRef]
  7. Novak, D.; Tilak, B.; Conway, B. Fundamental and applied aspects of anodic chlorine production. Mod. Asp. Electrochem. 1982, 195–318. [Google Scholar]
  8. Trasatti, S. Electrocatalysis: Understanding the success of DSA®. Electrochim. Acta 2000, 45, 2377–2385. [Google Scholar] [CrossRef]
  9. Vos, J.G.; Wezendonk, T.A.; Jeremiasse, A.W.; Koper, M.T.M. MnOx/IrOx as Selective Oxygen Evolution Electrocatalyst in Acidic Chloride Solution. J. Am. Chem. Soc. 2018, 140, 10270–10281. [Google Scholar] [CrossRef] [Green Version]
  10. Vos, J.G.; Koper, M.T.M. Measurement of competition between oxygen evolution and chlorine evolution using rotating ring-disk electrode voltammetry. J. Electroanal. Chem. 2018, 819, 260–268. [Google Scholar] [CrossRef]
  11. Exner, K.S.; Anton, J.; Jacob, T.; Over, H. Microscopic Insights into the Chlorine Evolution Reaction on RuO2(110): A Mechanistic Ab Initio Atomistic Thermodynamics Study. Electrocatalysis 2015, 6, 163–172. [Google Scholar] [CrossRef]
  12. Exner, K.S.; Anton, J.; Jacob, T.; Over, H. Controlling selectivity in the chlorine evolution reaction over RuO2-based catalysts. Angew. Chem. Int. Ed. 2014, 53, 11032–11035. [Google Scholar] [CrossRef] [PubMed]
  13. Vos, J.G.; Liu, Z.; Speck, F.D.; Perini, N.; Fu, W.; Cherevko, S.; Koper, M.T.M. Selectivity Trends Between Oxygen Evolution and Chlorine Evolution on Iridium-Based Double Perovskites in Acidic Media. ACS Catal. 2019, 9, 8561–8574. [Google Scholar] [CrossRef] [Green Version]
  14. Exner, K.S. Controlling Stability and Selectivity in the Competing Chlorine and Oxygen Evolution Reaction over Transition Metal Oxide Electrodes. ChemElectroChem 2019, 6, 3401–3409. [Google Scholar] [CrossRef]
  15. Exner, K.S.; Anton, J.; Jacob, T.; Over, H. Full Kinetics from First Principles of the Chlorine Evolution Reaction over a RuO2 (110) Model Electrode. Angew. Chem. Int. Ed. 2016, 55, 7501–7504. [Google Scholar] [CrossRef]
  16. Lim, T.; Kim, J.H.; Kim, J.; Baek, D.S.; Shin, T.J.; Jeong, H.Y.; Lee, K.-S.; Exner, K.S.; Joo, S.H. General Efficacy of Atomically Dispersed Pt Catalysts for the Chlorine Evolution Reaction: Potential-Dependent Switching of the Kinetics and Mechanism. ACS Catal. 2021, 11, 12232–12246. [Google Scholar] [CrossRef]
  17. Exner, K.S. Beyond Dimensionally Stable Anodes: Single-Atom Catalysts with Superior Chlorine Selectivity. ChemElectroChem 2020, 7, 1528–1530. [Google Scholar] [CrossRef]
  18. Exner, K.S. Design criteria for the competing chlorine and oxygen evolution reactions: Avoid the OCl adsorbate to enhance chlorine selectivity. Phys. Chem. Chem. Phys. 2020, 22, 22451–22458. [Google Scholar] [CrossRef]
  19. Exner, K.S. Beyond the Traditional Volcano Concept: Overpotential-Dependent Volcano Plots Exemplified by the Chlorine Evolution Reaction over Transition-Metal Oxides. J. Phys. Chem. C 2019, 123, 16921–16928. [Google Scholar] [CrossRef]
  20. Gupta, U.; Rao, C.N.R. Hydrogen generation by water splitting using MoS2 and other transition metal dichalcogenides. Nano Energy 2017, 41, 49–65. [Google Scholar] [CrossRef]
  21. Faber, M.S.; Jin, S. Earth-abundant inorganic electrocatalysts and their nanostructures for energy conversion applications. Energy Environ. Sci. 2014, 7, 3519–3542. [Google Scholar] [CrossRef]
  22. Qiu, W.; Xie, X.Y.; Qiu, J.; Fang, W.H.; Liang, R.; Ren, X.; Ji, X.; Cui, G.; Asiri, A.M.; Cui, G.; et al. High-performance artificial nitrogen fixation at ambient conditions using a metal-free electrocatalyst. Nat. Commun. 2018, 9, 3485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, L.; Ji, X.; Ren, X.; Ma, Y.; Shi, X.; Tian, Z.; Asiri, A.M.; Chen, L.; Tang, B.; Sun, X. Electrochemical Ammonia Synthesis via Nitrogen Reduction Reaction on a MoS2 Catalyst: Theoretical and Experimental Studies. Adv. Mater. 2018, 30, e1800191. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, F.; Zhu, Y.; Tang, C.; Chen, Y.; Qian, B.; Hu, Z.; Chang, Y.C.; Pao, C.W.; Lin, Q.; Kazemi, S.A.; et al. High-Efficiency Electrosynthesis of Hydrogen Peroxide from Oxygen Reduction Enabled by a Tungsten Single Atom Catalyst with Unique Terdentate N1O2 Coordination. Adv. Funct. Mater. 2021, 2110224. [Google Scholar] [CrossRef]
  25. Khan, K.; Yan, X.; Yu, Q.; Bae, S.-H.; White, J.J.; Liu, J.; Liu, T.; Sun, C.; Kim, J.; Cheng, H.-M.; et al. Stone-Wales defect-rich carbon-supported dual-metal single atom sites for Zn-air batteries. Nano Energy 2021, 90, 106488. [Google Scholar] [CrossRef]
  26. Ramalingam, V.; Varadhan, P.; Fu, H.C.; Kim, H.; Zhang, D.; Chen, S.; Song, L.; Ma, D.; Wang, Y.; Alshareef, H.N.; et al. Heteroatom-Mediated Interactions between Ruthenium Single Atoms and an MXene Support for Efficient Hydrogen Evolution. Adv. Mater. 2019, 31, e1903841. [Google Scholar] [CrossRef] [Green Version]
  27. Xiong, Q.; Wang, Y.; Liu, P.F.; Zheng, L.R.; Wang, G.; Yang, H.G.; Wong, P.K.; Zhang, H.; Zhao, H. Cobalt Covalent Doping in MoS2 to Induce Bifunctionality of Overall Water Splitting. Adv. Mater. 2018, 30, e1801450. [Google Scholar] [CrossRef]
  28. Li, L.; Yu, Y.; Ye, G.J.; Ge, Q.; Ou, X.; Wu, H.; Feng, D.; Chen, X.H.; Zhang, Y. Black phosphorus field-effect transistors. Nat Nanotechnol 2014, 9, 372–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Bat-Erdene, M.; Xu, G.; Batmunkh, M.; Bati, A.S.R.; White, J.J.; Nine, M.J.; Losic, D.; Chen, Y.; Wang, Y.; Ma, T.; et al. Surface oxidized two-dimensional antimonene nanosheets for electrochemical ammonia synthesis under ambient conditions. J. Mater. Chem. A 2020, 8, 4735–4739. [Google Scholar] [CrossRef]
  30. Ren, X.; Li, Z.; Qiao, H.; Liang, W.; Liu, H.; Zhang, F.; Qi, X.; Liu, Y.; Huang, Z.; Zhang, D.; et al. Few-Layer Antimonene Nanosheet: A Metal-Free Bifunctional Electrocatalyst for Effective Water Splitting. ACS Appl. Energy Mater. 2019, 2, 4774–4781. [Google Scholar] [CrossRef]
  31. Zhang, S.; Xie, M.; Li, F.; Yan, Z.; Li, Y.; Kan, E.; Liu, W.; Chen, Z.; Zeng, H. Semiconducting Group 15 Monolayers: A Broad Range of Band Gaps and High Carrier Mobilities. Angew. Chem. Int. Ed. 2016, 55, 1666–1669. [Google Scholar] [CrossRef] [PubMed]
  32. Pumera, M.; Sofer, Z. 2D Monoelemental Arsenene, Antimonene, and Bismuthene: Beyond Black Phosphorus. Adv. Mater. 2017, 29, 1605299. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, J.H.; Wang, E.G.; Ting, C.S.; Su, W.P. Tight-binding theory of the electronic structures for rhombohedral semimetals. Phys. Rev. B 1993, 48, 17271–17279. [Google Scholar] [CrossRef] [PubMed]
  34. Lim, T.; Jung, G.Y.; Kim, J.H.; Park, S.O.; Park, J.; Kim, Y.T.; Kang, S.J.; Jeong, H.Y.; Kwak, S.K.; Joo, S.H. Atomically dispersed Pt-N4 sites as efficient and selective electrocatalysts for the chlorine evolution reaction. Nat. Commun. 2020, 11, 412. [Google Scholar] [CrossRef]
  35. Exner, K.S.; Sohrabnejad-Eskan, I.; Over, H. A Universal Approach To Determine the Free Energy Diagram of an Electrocatalytic Reaction. ACS Catal. 2018, 8, 1864–1879. [Google Scholar] [CrossRef]
  36. Exner, K.S.; Anton, J.; Jacob, T.; Over, H. Chlorine Evolution Reaction on RuO2(110): Ab initio Atomistic Thermodynamics Study-Pourbaix Diagrams. Electrochim. Acta 2014, 120, 460–466. [Google Scholar] [CrossRef]
  37. Menzel, N.; Ortel, E.; Mette, K.; Kraehnert, R.; Strasser, P. Dimensionally Stable Ru/Ir/TiO2-Anodes with Tailored Mesoporosity for Efficient Electrochemical Chlorine Evolution. ACS Catal. 2013, 3, 1324–1333. [Google Scholar] [CrossRef]
  38. Jiang, M.; Wang, H.; Li, Y.; Zhang, H.; Zhang, G.; Lu, Z.; Sun, X.; Jiang, L. Superaerophobic RuO2 -Based Nanostructured Electrode for High-Performance Chlorine Evolution Reaction. Small 2017, 13, 1602240. [Google Scholar] [CrossRef]
  39. Chen, R.; Trieu, V.; Natter, H.; Stöwe, K.; Maier, W.F.; Hempelmann, R.; Bulan, A.; Kintrup, J.; Weber, R. In situ Supported Nanoscale RuxTi1−xO2 on Anatase TiO2 with Improved Electroactivity. Chem. Mater. 2010, 22, 6215–6217. [Google Scholar] [CrossRef]
  40. Zhu, X.; Wang, P.; Wang, Z.; Liu, Y.; Zheng, Z.; Zhang, Q.; Zhang, X.; Dai, Y.; Whangbo, M.-H.; Huang, B. Co3O4 nanobelt arrays assembled with ultrathin nanosheets as highly efficient and stable electrocatalysts for the chlorine evolution reaction. J. Mater. Chem. A 2018, 6, 12718–12723. [Google Scholar] [CrossRef]
  41. Moreno-Hernandez, I.A.; Brunschwig, B.S.; Lewis, N.S. Crystalline nickel, cobalt, and manganese antimonates as electrocatalysts for the chlorine evolution reaction. Energy Environ. Sci. 2019, 12, 1241–1248. [Google Scholar] [CrossRef]
  42. Liu, J.; Hinsch, J.J.; Yin, H.; Liu, P.; Zhao, H.; Wang, Y. TMN4 complex embedded graphene as efficient and selective electrocatalysts for chlorine evolution reactions. J. Electroanal. Chem. 2022, 907, 116071. [Google Scholar] [CrossRef]
  43. Dickens, C.F.; Nørskov, J.K. A Theoretical Investigation into the Role of Surface Defects for Oxygen Evolution on RuO2. J. Phys. Chem. C 2017, 121, 18516–18524. [Google Scholar] [CrossRef]
  44. Rao, R.R.; Kolb, M.J.; Halck, N.B.; Pedersen, A.F.; Mehta, A.; You, H.; Stoerzinger, K.A.; Feng, Z.; Hansen, H.A.; Zhou, H.; et al. Towards identifying the active sites on RuO2(110) in catalyzing oxygen evolution. Energy Environ. Sci. 2017, 10, 2626–2637. [Google Scholar] [CrossRef] [Green Version]
  45. Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558. [Google Scholar] [CrossRef]
  46. 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]
  47. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  48. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  49. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [Green Version]
  50. Dobson, J.F.; Gould, T. Calculation of dispersion energies. J. Phys. Condens. Matter. 2012, 24, 073201. [Google Scholar] [CrossRef]
  51. Bjorkman, T.; Gulans, A.; Krasheninnikov, A.V.; Nieminen, R.M. Are we van der Waals ready? J. Phys. Condens. Matter 2012, 24, 424218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Dobson, J.F.; White, A.; Rubio, A. Asymptotics of the Dispersion Interaction: Analytic Benchmarks for van der Waals Energy Functionals. Phys. Rev. Lett. 2006, 96, 073201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Sun, T.; Wang, Y.; Zhang, H.; Liu, P.; Zhao, H. Adsorption and oxidation of oxalic acid on anatase TiO2 (001) surface: A density functional theory study. J. Colloid Interface Sci. 2015, 454, 180–186. [Google Scholar] [CrossRef] [PubMed]
  54. Dronskowski, R.; Bloechl, P.E. Crystal orbital Hamilton populations (COHP): Energy-resolved visualization of chemical bonding in solids based on density-functional calculations. J. Phys. Chem. 1993, 97, 8617–8624. [Google Scholar] [CrossRef]
  55. Deringer, V.L.; Tchougreeff, 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]
  56. Maintz, S.; Deringer, V.L.; Tchougreeff, A.L.; Dronskowski, R. Analytic projection from plane-wave and PAW wavefunctions and application to chemical-bonding analysis in solids. J. Comput. Chem. 2013, 34, 2557–2567. [Google Scholar] [CrossRef]
  57. Maintz, S.; Deringer, V.L.; Tchougreeff, A.L.; Dronskowski, R. LOBSTER: A tool to extract chemical bonding from plane-wave based DFT. J. Comput. Chem. 2016, 37, 1030–1035. [Google Scholar] [CrossRef] [Green Version]
  58. Nelson, R.; Ertural, C.; George, J.; Deringer, V.L.; Hautier, G.; Dronskowski, R. LOBSTER: Local orbital projections, atomic charges, and chemical-bonding analysis from projector-augmented-wave-based density-functional theory. J. Comput. Chem. 2020, 41, 1931–1940. [Google Scholar] [CrossRef]
  59. Nørskov, J.K.; Bligaard, T.; Logadottir, A.; Kitchin, J.R.; Chen, J.G.; Pandelov, S.; Stimming, U. Trends in the Exchange Current for Hydrogen Evolution. J. Electrochem. Soc. 2005, 152, J23. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Atomic configurations of all group-VA 2D monolayers. (a) α and (b) β phase. Red circles represent the different adsorption sites for Cl atoms. Top panel—top view; bottom panel—side view.
Figure 1. Atomic configurations of all group-VA 2D monolayers. (a) α and (b) β phase. Red circles represent the different adsorption sites for Cl atoms. Top panel—top view; bottom panel—side view.
Catalysts 12 00296 g001
Figure 2. Gibbs free energy diagrams for CER over (a) α and (b) β phase of group-VA monolayers.
Figure 2. Gibbs free energy diagrams for CER over (a) α and (b) β phase of group-VA monolayers.
Catalysts 12 00296 g002
Figure 3. COHP analyses of the chemical bonding between the adsorbed chlorine and group-VA element for (a) α-phosphorene; (b) α-arsenene; (c) α-antimonene; (d) α-bismuthene; (e) β-phosphorene; (f) β-arsenene; (g) β-antimonene; and (h) β-bismuthene monolayers, with the adsorbed Cl atom. The bonding and antibonding states are positive and negative of −pCOHP, respectively. The corresponding integral of −pCOHP (−IpCOHP) values are also provided.
Figure 3. COHP analyses of the chemical bonding between the adsorbed chlorine and group-VA element for (a) α-phosphorene; (b) α-arsenene; (c) α-antimonene; (d) α-bismuthene; (e) β-phosphorene; (f) β-arsenene; (g) β-antimonene; and (h) β-bismuthene monolayers, with the adsorbed Cl atom. The bonding and antibonding states are positive and negative of −pCOHP, respectively. The corresponding integral of −pCOHP (−IpCOHP) values are also provided.
Catalysts 12 00296 g003
Figure 4. The electron localization functions for (a) α-phosphorene; (b) α-arsenene; (c) α-antimonene; (d) α-bismuthene; (e) β-phosphorene; (f) β-arsenene; (g) β-antimonene; and (h) β-bismuthene monolayers with the adsorbed Cl atom.
Figure 4. The electron localization functions for (a) α-phosphorene; (b) α-arsenene; (c) α-antimonene; (d) α-bismuthene; (e) β-phosphorene; (f) β-arsenene; (g) β-antimonene; and (h) β-bismuthene monolayers with the adsorbed Cl atom.
Catalysts 12 00296 g004
Figure 5. (a) The adsorption free energy for Cl* and HO* on α-phosphorene and β-arsenene monolayers as a function of the theoretical standard hydrogen electrode potential (USHE) at pH = 0; (b) Gibbs free energy diagram for OER on β-arsenene monolayer.
Figure 5. (a) The adsorption free energy for Cl* and HO* on α-phosphorene and β-arsenene monolayers as a function of the theoretical standard hydrogen electrode potential (USHE) at pH = 0; (b) Gibbs free energy diagram for OER on β-arsenene monolayer.
Catalysts 12 00296 g005
Table 1. Structural parameters of all group-VA 2D monolayers. a and b are the lattice constants, d is the thickness of the monolayer, and L denotes the bond lengths.
Table 1. Structural parameters of all group-VA 2D monolayers. a and b are the lattice constants, d is the thickness of the monolayer, and L denotes the bond lengths.
Phasesa (Å)b (Å)d (Å)BondsL (Å)
Pα4.593.302.11in-plane 2.22
out-plane 2.26
β3.273.271.24P−P2.26
Asα4.703.692.41in-plane2.51
out-plane2.50
β3.603.601.40As−As2.51
Sbα4.764.382.83in-plane2.94
out-plane2.86
β4.114.111.65Sb−Sb2.89
Biα4.904.582.99in-plane3.09
out-plane3.02
β4.334.331.73Bi−Bi3.04
Table 2. Calculated average distance between Cl and group-VA atoms (rCl−X), adsorption energy (ΔECl*), and Gibbs free energy change (ΔGCl*) on different sites of group-VA 2D monolayers with α and β phases.
Table 2. Calculated average distance between Cl and group-VA atoms (rCl−X), adsorption energy (ΔECl*), and Gibbs free energy change (ΔGCl*) on different sites of group-VA 2D monolayers with α and β phases.
StructuresrCl−X (Å)ΔECl* (eV)ΔGCl* (eV)
αP2.24−0.3150.055
As2.39−0.602−0.232
Sb2.55−1.226−0.856
Bi2.70−1.336−0.966
βPSite 12.21−0.1340.236
Site 23.040.5820.952
AsSite 12.37−0.3020.068
Site 23.070.2710.641
SbSite 12.51−0.585−0.215
Site 32.98−0.602−0.232
BiSite 12.59−0.797−0.427
Site 23.20−0.799−0.429
Site 33.08−1.091−0.721
Table 3. Summary of the experimental (ηexp) and thermodynamic overpotentials(ηtd) of reported electrocatalysts for chlorine evolution a.
Table 3. Summary of the experimental (ηexp) and thermodynamic overpotentials(ηtd) of reported electrocatalysts for chlorine evolution a.
Electrocatalystηexp (mV)
@ 10 mA cm−2
ηtd (V)Ref.
commercial Ru/Ir-based105N/A[34]
RuO2 (110)820.13[1,12,14,36]
Ru/Ir/TiO2>240N/A[37]
RuO2@TiO2~80N/A[38]
Ru0.3Ti0.7O2~80N/A[39]
Co3O4 200N/A[40]
CoSb2Ox~300N/A[41]
Pt1/CNT500.09[16,34]
PtO2 (110)N/A0.20[34]
PtNT/CNT120N/A[34]
β-arseneneN/A0.068This work
a N/A—not available; CNT—carbon nanotube; NP—nanoparticle.
Table 4. The bond lengths of Cl and group-VA atoms (rCl−X), Mulliken charges of the adsorbed Cl atom (qCl) and group-VA atoms (qX), and electrostatic attractions (Fes) of all α- and β-monolayers with the adsorbed Cl.
Table 4. The bond lengths of Cl and group-VA atoms (rCl−X), Mulliken charges of the adsorbed Cl atom (qCl) and group-VA atoms (qX), and electrostatic attractions (Fes) of all α- and β-monolayers with the adsorbed Cl.
StructuresrCl−X (Å)qCl (e)qX (e)Fes (eV/Å)
AP2.24−0.350.190.19
As2.39−0.430.200.22
Sb2.55−0.510.260.29
Bi2.70−0.560.290.32
ΒP2.21−0.290.250.21
As2.37−0.370.250.24
Sb2.51−0.440.350.35
Bi2.59−0.480.380.39
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, J.; Hinsch, J.J.; Yin, H.; Liu, P.; Zhao, H.; Wang, Y. β-Arsenene Monolayer: A Promising Electrocatalyst for Anodic Chlorine Evolution Reaction. Catalysts 2022, 12, 296. https://doi.org/10.3390/catal12030296

AMA Style

Liu J, Hinsch JJ, Yin H, Liu P, Zhao H, Wang Y. β-Arsenene Monolayer: A Promising Electrocatalyst for Anodic Chlorine Evolution Reaction. Catalysts. 2022; 12(3):296. https://doi.org/10.3390/catal12030296

Chicago/Turabian Style

Liu, Junxian, Jack Jon Hinsch, Huajie Yin, Porun Liu, Huijun Zhao, and Yun Wang. 2022. "β-Arsenene Monolayer: A Promising Electrocatalyst for Anodic Chlorine Evolution Reaction" Catalysts 12, no. 3: 296. https://doi.org/10.3390/catal12030296

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