Next Article in Journal
Improved Operation of Chloralkaline Reversible Cells with Mixed Metal Oxide Electrodes Made Using Microwaves
Previous Article in Journal
Optical Force Effects of Rayleigh Particles by Cylindrical Vector Beams
Previous Article in Special Issue
Feature-Assisted Machine Learning for Predicting Band Gaps of Binary Semiconductors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rational Design of Non-Noble Metal Single-Atom Catalysts in Lithium–Sulfur Batteries through First Principles Calculations

1
Department of Physics, College of Science, Yanbian University, Yanji 133002, China
2
Institute of Zhejiang University-Quzhou, Quzhou 324000, China
3
Key Laboratory of Biomass Chemical Engineering of Ministry of Education, College of Chemical and Biological Engineering, Zhejiang University, Hangzhou 310027, China
*
Authors to whom correspondence should be addressed.
These authors contribute equally to this work.
Nanomaterials 2024, 14(8), 692; https://doi.org/10.3390/nano14080692
Submission received: 19 March 2024 / Revised: 13 April 2024 / Accepted: 15 April 2024 / Published: 17 April 2024
(This article belongs to the Special Issue Theoretical Chemistry and Computational Simulations in Nanomaterials)

Abstract

:
Lithium–sulfur (Li–S) batteries with a high theoretical energy density of 2600 Wh·kg−1 are hindered by challenges such as low S conductivity, the polysulfide shuttle effect, low S reduction conversion rate, and sluggish Li2S oxidation kinetics. Herein, single-atom non-noble metal catalysts (SACs) loaded on two-dimensional (2D) vanadium disulfide (VS2) as the potential host materials for the cathode in Li–S batteries were investigated systematically by using first-principles calculations. Based on the comparisons of structural stability, the ability to immobilize sulfur, electrochemical reactivity, and the kinetics of Li2S oxidation decomposition between these non-noble metal catalysts and noble metal candidates, Nb@VS2 and Ta@VS2 were identified as the potential candidates of SACs with the decomposition energy barriers for Li2S of 0.395 eV (Nb@VS2) and of 0.162 eV (Ta@VS2), respectively. This study also identified an exothermic reaction for Nb@VS2 and the Gibbs free energy of 0.218 eV for Ta@VS2. Furthermore, the adsorption and catalytic mechanisms of the VS2-based SACs in the reactions were elucidated, presenting a universal case demonstrating the use of unconventional graphene-based SACs in Li–S batteries. This study presents a universal surface regulation strategy for transition metal dichalcogenides to enhance their performance as host materials in Li–S batteries.

Graphical Abstract

1. Introduction

With the growing energy density demand in fields such as electronic devices and electric vehicles, conventional commercial lithium-ion batteries cannot meet the requirements of development. Consequently, scientists are now dedicated to developing new energy storage candidates. Among various battery systems, Li–S batteries are regarded as a promising energy storage candidate based on the abundant reserves, low cost, and non-toxic nature of sulfur, especially their impressively high theoretical specific energy density (2600 W∙h∙kg−1) and volumetric energy density (2800 W∙h∙L−1) [1,2,3,4]. However, the large-scale commercial application of Li–S batteries is still impeded because of the following reasons: (i) low sulfur utilization resulting from the poor conductivity of the sulfur (S8) cathode material and the insoluble discharging products of lower-order lithium polysulfides (LiPSs) such as Li2S2 and Li2S; (ii) the continuous loss of active substances due to the notorious “shuttle effect” of soluble higher-order LiPSs Li2Sn (4 ≤ n ≤ 8) in the charging/discharging process, wherein soluble intermediate higher-order LiPSs species in the electrolyte reach the lithium anode and react with Li metal to form insoluble Li2Sn (1 ≤ n < 4) and deposit on it, which results in the sluggish kinetics of LiPSs, quick reductions in capacity, and cycle stability issues; (iii) the pulverization of the sulfur cathode caused by large volumetric expansion during the lithiation process [5]. In 2D materials, the 100 cycle capacity retention rate using catalytic materials could be increased by about 20% compared to not using catalytic materials; for example, the capacity retention rate of MnO increased from 76.1% to 94.4% after the addition of catalytic materials [6].
In recent years, some strategies have been proposed to address the above problems of Li–S batteries. Particularly, the composite material formed by adding additives to the cathodes of Li–S batteries is expected to solve the problem of low conductivity [7,8,9]. Some framework materials with physical confinement functions are expected to solve the problem of excessive volume changes to the sulfur cathode during charging and discharging processes. Meanwhile, a lot of effort has been invested in developing suitable materials to mitigate the “shuttle effect”. The so-called suitable materials should bind LiPSs to inhibit their migration and accelerate their redox kinetic process, that is, the suitable materials should be materials with anchoring functions [10]. Anchoring materials have been regarded as electrocatalysts which play a crucial role in LiPSs conversion and reduce the reaction barriers, which has attracted great attention. For example, some 2D materials [11,12], including graphene [13,14], MXenes [15], transition metal carbides [16,17], and metal sulfides [18,19], have been elucidated as promising anchoring materials for Li–S batteries due to large specific surface areas and unique physical and chemical properties. Many studies have focused on enhancing the binding of the anchoring material and LiPSs by predicting the adsorption energy based on density functional theory (DFT) simulations. However, most of these 2D materials exhibit incomplete LiPSs conversion to Li2S2 or Li2S due to the sluggish redox reaction kinetics.
Recently, SAC materials consisting of monodispersed transition metal (TM) atoms supported on the surfaces of 2D materials have received much attention. Apart from the maximum metal atom utilization efficiency, SACs can offer outstanding catalytic activity with low metal loading, tunable structure, flexible selectivity, and low cost. In 2021, Cheng’s group proposed that single-atom Ti as the sulfur cathode catalyst could promote battery performance for Li–S batteries, which had the lowest electrochemical barrier to LiPSs reduction/Li2S oxidation and the highest catalytic activity. Because of the highly active catalytic center of single-atom Ti on the conductive transport network, high sulfur utilization was achieved with low catalyst loading (1 wt.%) and a high area of sulfur loading (8 mg·cm−2) [20,21]. Compared with 2D materials, SACs for Li–S batteries can provide suitable adsorption strength with LiPSs and the exceptional ability to improve reaction kinetics [22,23,24,25]. It can be seen that the adsorption energy between Co and N/G (−7.99 eV) is more negative than those of CoS (−6.39 eV) and CoS2 (−7.01 eV) [26]. Among the reported 2D materials in Li–S batteries, 2D transition metal chalcogenides have been widely researched and applied because of their abundant reserves, environmentally friendly nature, and fast reaction kinetics. In particular, T-phase 2D VS2 has good intrinsic conductivity compared with FeS2, MoS2, NiS2, NbS2, TiS2, and ZrS2, as well as moderate adsorption strength for Li2Sn [25,27,28]. For example, to further validate that the adsorption capacities of LiPSs on the surfaces of FeS2 were moderate for Li2Sn, the adsorption energies of Li2S4 on FeS2, Ti3C2, CoP, and graphene surfaces were compared and measured as −4.25, −4.16, −4.91, and −1.21 eV, respectively, indicating a moderate adsorption ability of FeS2 [29].
2D VS2 as a host material loaded the monodispersed TM atom (hereafter denoted as TM@VS2) to constitute SAC materials, which is expected to have the dual effect of suppressing the shuttle effect and accelerating reaction kinetics [22,26,30,31,32]. In this work, 17 unique SACs TM@VS2 (TM = 4d and 5d TMs) for sulfur cathodes were constructed and investigated systematically via DFT simulations. By comparing the thermal stability values of 17 TM@VS2, the adsorption behavior of S8 and five LiPSs intermediates (Li2Sn, n = 1, 2, 4, 6, and 8), and the conversion mechanism of Li2S2 to Li2S on the TM@VS2, the promising SAC candidates Nb@VS2 and Ta@VS2 were screened out. This study provided an atomic-level understanding about the reaction mechanism of this composite material TM@VS2 in Li–S batteries and clarified how to inhibit the shuttle effect and accelerate the reaction kinetics, which is of significance for the rational design of SACs.

2. Methods

All the DFT simulations were performed using the Vienna ab initio simulation package (VASP) [33]. Projector augmented wave (PAW) pseudopotentials and Perdew–Bruke–Ernzerhof (PBE) functions with the generalized gradient approximation (GGA) were employed to investigate the electron–ion interactions and the electron–electron exchange correlations [20,34]. The kinetic energy cutoff was selected as 400 eV for the plane wave basis calculations, the energy convergence was set to 10−5 eV/Å, and the force convergence criterion for optimization was set to 0.05 eV/Å. A vacuum space of 20 Å was set to avoid the interactions between the periodic boundaries. The van der Waals interactions were calculated using the DFT-D3 method to provide better accuracy for the adsorption strength of polysulfides with TM@VS2. A 4 × 4 × 1 supercell of VS2 monolayer was built and a 2 × 2 × 1 K-point was set in the first Brillouin zone for the geometric optimizations and the density of states (DOS) calculation [35]. DOS and electrostatics potential were calculated by DS-PAW software (2023A) integrated in the Device Studio program [36]. In addition, ab initio molecular dynamics (AIMD) simulations of Pd@VS2, Rh@VS2, and Ta@VS2 were carried out to confirm material stabilities on the 8 × 4 × 1 supercell at 298.15 K within 10 ps. The decomposition properties of Li2S were determined using the climbing-image nudged elastic band (CI-NEB) method. The SACs among 17 single TM atoms loaded on the VS2 monolayer were first screened by calculating the adsorption energy following Equation (1):
E a d 1 = E T M @ V S 2 E T M E V S 2
where E T M @ V S 2 and E V S 2 are the energies of loaded single atoms and the bare VS2 monolayer, respectively. ETM is the energy per single atom in the metallic phase. A more negative value of E a d 1 represents a stronger adsorption.
The adsorption energies E a d 2 of Li2Sn and S8 molecules adsorbed on the VS2 and TM@VS2 monolayer are calculated based on the following Equation (2):
E a d 2 = E t o t a l E s u b E m o l
where E t o t a l represents the total energy of adsorbed system Li2Sn (or S8) on TM@VS2 or VS2 monolayer, E s u b represents the energy of TM@VS2 or VS2 monolayer, and E m o l represents the energy of Li2Sn (n = 1, 2, 4, 6, 8) and the S8 molecule.
To visualize the charge transfer of Li2Sn molecules on TM@VS2 monolayers, the charge density differences (CDDs) of the adsorbed configurations were displayed by calculating from the following Equation (3):
ρ = ρ t o t a l ρ s u b ρ m o l
where ρ t o t a l , ρ T M , and ρ m o l represent the charge density of the whole adsorbed system, TM@VS2 or VS2 monolayer, and Li2Sn and S8 molecule, respectively.
The Gibbs free energies can be defined as follows (4):
G = E + Z P E T S
where E represents the electronic energy of the product minus that of the reactant, and Z P E and S represent the zero-point vibrational energy and entropy change at the standard temperature and pressure, respectively.

3. Results and Discussion

3.1. Structure and Stability of the TM@VS2 SACs

The configuration of T-phase VS2 is shown in Figure 1a. There are three possible sites where a single TM atom can deposit on the surface of the VS2 monolayer, including Hollow-V (H-V), Top-S (T-S), and Hollow-S (H-S). In order to screen the most suitable loaded single TM atoms on the VS2 monolayer to constitute SACs, the adsorption energies of 17 4d and 5d TM atoms loaded on the VS2 surface at three possible sites were calculated based on Equation (1). Remarkably, the adsorption of single atoms at the T-S site after the optimization became higher than the H-S or H-V sites, such that the adsorption energies were higher than those of the H-S or H-V sites, even to positive values (about 0.5–7.1 eV), which indicated that the adsorption of the T-S site was unstable thermodynamically. Therefore, all the TM@VS2 at the T-S site will not be considered in the following. The adsorption energies at H-V and H-S sites are depicted in Figure 1b. The results revealed that the single TM atoms of four species (Y, Zr, Nb, and Hf) were strong in affinity and stably adsorbed on the VS2 surface due to the negative adsorption energies. In addition, Pd@VS2, Rh@VS2, and Ta@VS2 displayed slightly positive adsorption energies, suggesting their comparatively weaker interactions. Further, the thermal stability of Pd@VS2, Rh@VS2, and Ta@VS2 was considered by AIMD simulation. The final configurations of Pd@VS2, Rh@VS2, and Ta@VS2 after AIMD simulations are depicted in Figure S1a–c. The simulation profiles revealed a consistent trend of converging energy, with the structures maintaining their original coordination environment even after 10 ps. No structural deformations or atomic clustering were observed, demonstrating the thermal stability of Pd@VS2, Rh@VS2, and Ta@VS2. Furthermore, Figure 1 shows that Hf, Rh, Y, and Zr atoms were preferentially adsorbed at the H-S site, while Pd, Ta, and Nb atoms preferred the H-V site. For a single TM atom, adsorption at the H-V and H-S sites on the VS2 surface site are comparable, and the more favorable sites can be screened by calculating the COHP of TM–S bonds at the H-S and H-V sites based on a comparison of adsorption energies, with sites showing more negative adsorption energies being considered optimal. Compared to the other site, Nb–V, Ta–V, Zr–S, and Rh–S demonstrate higher ICOHP values, indicating stronger bonding strength with single atoms (Figure S2). This disparity in site behavior arises from differences in the types of single atoms and the adsorption strengths of various adsorption sites on the VS2 substrate. The lowest adsorption energy corresponds to the most stable configuration, as shown in Figure S3a. Furthermore, different coordination numbers have different geometric structures and chemical properties. The optimized configurations of the SACs substantiate a TM–S with three coordinates on the VS2 substrate. Typically, compared to four coordinates in TM–N4–graphene [37,38], TM–S with three coordinates on a VS2 surface has fewer coordination numbers, resulting in a single atom with more unbonded electrons, and these sites can adsorb and activate reactants, facilitating the reaction.
In addition, in order to better understand the bonding properties between the TM atom and VS2 monolayer, the projected density of states (PDOS) of TM@VS2 was calculated based on fully relaxed configurations, as shown in Figure 2. The results provided compelling evidence of significant d-p orbital hybridization between the TM and S atoms, indicating that there were stable covalent bonds between TM and S atoms. Furthermore, a continuous state density across the Fermi level indicated that these TM@VS2 were metallic. This observation carried crucial implications for enhancing the conductivity of the cathode. Specifically, the adsorption energy of the TM and VS2 monolayer is often linked to the electron state’s density, as the d orbitals of Y, Zr, and Hf exhibit more electron states above the Fermi level than other metal atoms, therefore exhibiting stronger adsorption energy.

3.2. Anchoring Effect of TM@VS2 SACs

The primary objective of catalysts of the sulfur cathode for Li–S batteries is to enhance the immobilization of sulfur and polysulfides. To investigate the anchoring effect of TM@VS2, the adsorption energies of VS2 and the TM@VS2 monolayer for the intermediate polysulfides Li2Sn (n = 1, 2, 4, 6, 8) and S8 during charge–discharge processes were calculated based on Equation (2). The values of all adsorption energies were shown in Figure 3a and Table S1. The adsorption energies of Li2Sn and S8 on the surface of VS2 were as follows: Li2S (−4.613 eV), Li2S2 (−3.158 eV), Li2S4 (−2.525 eV), Li2S6 (−1.915 eV), Li2S8 (−1.918 eV), and S8 (−0.957 eV). The corresponding optimal configurations of VS2 are given in Figure 3b. Nb shows obvious higher electrostatic potential than that of Pd (Figure S4), which is an electrophilic center to adsorb polysulfides favorably in the reaction process, and the adsorption strength was enhanced to inhibit the shuttle effect effectively compared to the VS2 substrate. The corresponding optimal configurations of Ta@VS2, Pd@VS2, Hf@VS2, Rh@VS2, Y@VS2, Nb@VS2, and Zr@VS2 are given in Figure 3c and Figure S3b–g. There is no significant structural deformation observed after the introduction of single atoms. The results indicated that most of the adsorption energies were enhanced after the introduction of TM single atoms on the VS2 monolayer, where the adsorption energies of Ta@VS2, Hf@VS2, Y@VS2, Nb@VS2, and Zr@VS2 could all be enhanced for Li2S2, Li2S6, Li2S8, and S8 compared with VS2, which exhibited the sulfur immobilization ability of these non-noble metal SACs. However, noble metal atoms Pd and Rh enhanced adsorption ability partially for S8 and LiPSs compared with VS2. In addition, the adsorption strengths of these five non-noble metal SACs for soluble higher-order LiPSs Li2Sn (4 ≤ n ≤ 8) were significantly higher than the interactions between common electrolyte components and soluble intermediates [39]; for example, the adsorption energy of 1,3-dioxolane (DOL) for Li2S6 was −0.75 eV, and that of 1,2-dimethoxyethane (DME) was −0.77 eV. Differently, the adsorption energies for Li2S6 were −3.802 eV (Nb@VS2) and −3.860 eV (Hf@VS2), which effectively prevented the shuttle effect of LiPSs. Therefore, non-noble metal atoms Hf, Nb, Ta, Y, and Zr loaded on VS2 monolayer have the potential to enhance sulfur immobilization capability compared to noble metal atoms Pd and Rh.
Taking Ta@VS2 as an example, the optimized adsorption configurations are shown in Figure 3b, which reveals that Ta@VS2 forms both Li–S and Ta–S bonds during the adsorption process. Compared to the Li–S bond on the surface of VS2, the presence of more bonding sites and species offers the potential to enhance the adsorption effect for TM@VS2. To gain a further insight into the mechanisms between the adsorption strength and bonding, the CDD between TM@VS2 and the adsorbed intermediates were investigated, and the results are shown in Figure S5 for VS2, Pd@VS2, Hf@VS2, Rh@VS2, Y@VS2, Ta@VS2, Nb@VS2, and Zr@VS2, respectively. The results revealed that the electron transfer primarily occurred between Li atoms and S atoms on the surface of VS2 monolayer. However, for non-noble SACs such as Hf@VS2, Nb@VS2, Ta@VS2, Zr@VS2, and Y@VS2, the electron transfer occurred not only between Li atoms and S atoms but also between the single TM atoms and the S atoms in Li2Sn, forming TM–S bonds. This implies there was a more significant electron transfer between TM@VS2 and the S atoms in Li2Sn clusters. For all the non-noble SACs, the electron transfer from S8 to Li2S gradually increases, and the largest charge transfer occurred between the SACs and Li2S, corresponding to the strongest adsorption energy. The enhanced adsorption of S8 can also be attributed to the redistribution of charge between individual TM atoms. Furthermore, there exists a competitive bonding interaction between Li–TM@VS2 and Li–S/S–S, wherein TM atoms interact with S atoms to form charge density accumulations during chemical bonding. Meanwhile, varying degrees of charge loss are observed in the S–S/Li–S bonds within the Li2Sn molecules. As lithiation progresses, the strength of Li–S bonds within the Li2Sn clusters weakened, while the adsorption capability of Li2Sn on TM@VS2 increased. In addition, the weakened Li–S bonds are expected to promote the reaction kinetics of Li2Sn conversion.

3.3. The Catalytic Mechanism of TM@VS2

To ascertain the catalytic performance of TM@VS2 (TM = Y, Zr, Hf, Nb, Ta, Rh, Pd) in the sulfur reduction reaction, the Gibbs free energy for the conversion of S8 to Li2S was calculated. The energy landscape was depicted in Figure 4 and Figure S6. The research results revealed that the Gibbs free energies of the reaction-determining step were 0.344 (VS2), 0.176 (Y@VS2), 0.320 (Zr@VS2), 0.408 (Hf@VS2), 0.218 (Ta@VS2), and 0.154 eV (Pd@VS2), respectively. The Gibbs free energy of Rh@VS2 and Pd@VS2 are shown in Figure S6. The reactions of Rh@VS2 and Nb@VS2 were exothermic, indicating that the reactions proceeded readily. According to the above, Hf@VS2 has a higher Gibbs free energy, which is unfavorable for the reaction thermodynamics. Therefore, Y@VS2, Pd@VS2, Zr@VS2, and Ta@VS2 accelerated the sulfur lithiation reaction compared to VS2, and Nb@VS2 and Rh@VS2 experienced exothermic reactions, which highlighted their relatively strong catalytic activity in the sulfur reduction reaction.
In Li–S batteries, the generation of Li2S as the final discharge product causes the challenge of sluggish reaction kinetics due to its low electronic conductivity and Li+ diffusion rates [40]. A crucial approach to address the slow oxidation rate during Li–S battery charging is to decrease the decomposition energy barrier of Li2S. Therefore, it is necessary to evaluate the catalytic oxidation performance of TM@VS2 for the charging process by examining the decomposition energy barriers of Li2S. The calculated results are shown in Figure 5 and Figure S7, and significant differences were observed in Li2S decomposition energy barriers on various surfaces. The decomposition energy barrier of Li2S on the VS2 surface was 0.366 eV. The corresponding decomposition energy barriers of Li2S on TM@VS2 were 0.162 (Ta@VS2), 0.322 (Hf@VS2), 0.396 (Nb@VS2), 0.496 (Zr@VS2), 0.682 (Rh@VS2), 0.736 (Pd@VS2), and 0.744 eV (Y@VS2). The results indicated that TM@VS2 (TM = Nb, Ta, Hf) could significantly reduce the decomposition energy barrier of Li2S compared to the VS2 surface. Among all the investigated seven TM@VS2, Ta@VS2 displayed the lowest decomposition energy barrier. This could be attributed to the formation of chemical bonds between the Ta and the S of Li2S, weakening the Li–S bonds and, therefore, lowering the decomposition energy barrier of Li2S. Taking this into account, crystal orbital Hamilton population (COHP) analysis was introduced to explore the bonding strength of Li–S and TM–S bonds in Li2S on TM@VS2, as shown in Figure 6. The COHPs of Ta@VS2 and Pd@VS2 were selected for comparative analysis as an example. Compared to Pd@VS2, the bonding state of the Ta–S bond below the Fermi level is significantly lower in Ta@VS2–Li2S than that of Y–S, with ICOHP values of 5.379 and 4.706 eV, respectively. This is accompanied by an increase in bond length from 2.425 Å for Ta–S to 2.497 Å for Y–S. Meanwhile, the Li–S bonds in the adsorbed state of Li2S exhibit an opposite ICOHP trend, with the Li–S bond in Ta@VS2–LiS at 0.887 and in Y@VS2–Li2S at 1.145. The corresponding Li–S bond lengths are 2.595 and 2.500 Å for Ta@VS2–Li2S and 2.343 and 2.368 Å for Y@VS2–Li2S. This indicates a significant weakening of the Li–S bonds in Li2S on Ta@VS2, promoting the catalytic oxidation of Li2S and aligning with the trend in decomposition energy barriers. Benefiting from high electronic conductivity and excellent bifunctional catalytic ability, Nb@VS2 and Ta@VS2 can effectively accelerate the Li2Sn conversion reaction and improve the electrode kinetics of Li–S batteries.

4. Conclusions

In summary, this study presents a systematic first-principles calculation of seven VS2-based SACs as host materials of the cathode for Li–S batteries. The calculation results show that among the seven selected monatomic catalysts, two noble metals Pd and Rh are removed, and five non-noble metals catalysts are obtained. By comparing the adsorption properties, metal properties, and reaction kinetics, it is found that the addition of Nb and Ta into VS2 has the potential to improve the performance of Li–S batteries. This enhancement is mainly attributed to the metallic characteristics of TM@VS2, which improve the electronic conductivity of the sulfur cathode. The enhanced adsorption of intermediate discharge products helps suppress the dissolution and shuttle effect of lithium polysulfides. Additionally, the lower reaction-determining step and reduced energy barrier for Li2S decomposition contribute to accelerated reaction kinetics. This work provides a universal approach for the atom-scale modification of transition metal sulfides, offering valuable insights and guidance for the development of sustainable and efficient Li–S battery technologies.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano14080692/s1. Figure S1. The final configurations of (a) Pd@VS2, (b) Rh@VS2 and (c) Ta@VS2. Figure S2. The ICOHP of TM-S bonds (TM@VS2, TM = Nb, Ta, Zr, Rh) for two kind of different site (H-V, H-S). Figure S3. (a) The optimal configurations of TM@VS2 (TM = Pd, Hf, Rh, Y, Nb, Zr). (b–g) Optimal configurations of Li2Sn and S8 on TM@VS2 (TM = Pd, Hf, Rh, Y, Nb, Zr). Figure S4. The electrostatics potential of (a) VS2, (b) Nb@VS2, (c) Pd@VS2 by DS-PAW method, respectively. Figure S5. Differential charge density of Li2S, Li2S6 and S8 on (a) VS2, (b) Pd@VS2, (c) Hf@VS2, (d) Rh@VS2, (e) Y@VS2, (f) Ta@VS2, (g) Nb@VS2 and (h) Zr@VS2, the iso-surface is set to 0.0025 e/A3 (Orange region indicates electron concentration and green region indicates electron deficiency). Figure S6. Gibbs Free Energy of S8–to–Li2S reaction on TM@VS2 (TM = Rh, Pd). Figure S7. Decomposition energy Barriers of Li2S on TM@VS2. Table S1. The adsorption energies of polysulfides on TM@VS2 (The units is electron volts (eV)). Table S2. The optimized coordinates of polysulfides on TM@VS2.

Author Contributions

Conceptualization, Y.L. (Yao Liu) and J.Z.; Investigation, Y.L. (Yang Li); Data curation, Y.L. (Yao Liu), J.Z. and J.X.; Writing—original draft, Y.L. (Yang Li), D.W. and J.X.; Supervision, J.X.; Project administration, D.W. and J.X.; Funding acquisition, D.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the research funds of the Institute of Zhejiang University—Quzhou (IZQ2021RCZX040) and the Yanbian University PhD start-up funds (602021029). We gratefully acknowledge HZWTECH for providing computation facilities.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bruce, P.G.; Freunberger, S.A.; Hardwick, L.J.; Tarascon, J.M. Li–O2 and Li–S batteries with high energy storage. Nat. Mater. 2012, 11, 19–29. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, X.; Hou, T.; Persson, K.A.; Zhang, Q. Combining theory and experiment in lithium–sulfur batteries: Current progress and future perspectives. Mater. Today 2019, 22, 142–158. [Google Scholar] [CrossRef]
  3. Evers, S.; Nazar, L.F. New Approaches for High Energy Density Lithium–Sulfur Battery Cathodes. Acc. Chem. Res. 2013, 46, 1135–1143. [Google Scholar] [CrossRef] [PubMed]
  4. Demir-Cakan, R.; Morcrette, M.; Nouar, F.; Davoisne, C.; Devic, T.; Gonbeau, D.; Dominko, R.; Serre, C.; Férey, G.; Tarascon, J.-M. Cathode Composites for Li–S Batteries via the Use of Oxygenated Porous Architectures. J. Am. Chem. Soc. 2011, 133, 16154–16160. [Google Scholar] [CrossRef] [PubMed]
  5. Feng, S.; Fu, Z.H.; Chen, X.; Zhang, Q. A review on theoretical models for lithium–sulfur battery cathodes. InfoMat 2022, 4, e12304. [Google Scholar] [CrossRef]
  6. Qian, X.; Jin, L.; Zhao, D.; Yang, X.; Wang, S.; Shen, X.; Rao, D.; Yao, S.; Zhou, Y.; Xiao, X. Ketjen Black-MnO Composite Coated Separator for High Performance Rechargeable Lithium-Sulfur Battery. Electrochim. Acta 2016, 192, 346–356. [Google Scholar] [CrossRef]
  7. Lu, Y.Q.; Wu, Y.J.; Sheng, T.; Peng, X.X.; Gao, Z.G.; Zhang, S.J.; Deng, L.; Nie, R.; Światowska, J.; Li, J.-T.; et al. Novel Sulfur Host Composed of Cobalt and Porous Graphitic Carbon Derived from MOFs for the High-Performance Li–S Battery. ACS Appl. Mater. Interfaces 2018, 10, 13499–13508. [Google Scholar] [CrossRef]
  8. Ji, X.; Lee, K.T.; Nazar, L.F. A highly ordered nanostructured carbon–sulphur cathode for lithium–sulphur batteries. Nat. Mater. 2009, 8, 500–506. [Google Scholar] [CrossRef] [PubMed]
  9. Xiao, L.; Cao, Y.; Xiao, J.; Schwenzer, B.; Engelhard, M.H.; Saraf, L.V.; Nie, Z.; Exarhos, G.J.; Liu, J. A Soft Approach to Encapsulate Sulfur: Polyaniline Nanotubes for Lithium-Sulfur Batteries with Long Cycle Life. Adv. Mater. 2012, 24, 1176–1181. [Google Scholar] [CrossRef]
  10. Wang, L.; Hua, W.; Wan, X.; Feng, Z.; Hu, Z.; Li, H.; Niu, J.; Wang, L.; Wang, A.; Liu, J.Y. Design rules of a sulfur redox electrocatalyst for lithium–sulfur batteries. Adv. Mater. 2022, 34, 2110279. [Google Scholar] [CrossRef]
  11. Yang, Y.; Liu, X.; Zhu, Z.; Zhong, Y.; Bando, Y.; Golberg, D.; Yao, J.; Wang, X. The Role of Geometric Sites in 2D Materials for Energy Storage. Joule 2018, 2, 1075–1094. [Google Scholar] [CrossRef]
  12. Chang, C.; Chen, W.; Chen, Y.; Chen, Y.; Chen, Y.; Ding, F.; Fan, C.; Fan, H.J.; Fan, Z.; Gong, C. Recent progress on two-dimensional materials. Acta Phys. -Chim. Sin. 2021, 37, 2108017. [Google Scholar] [CrossRef]
  13. Andritsos, E.I.; Lekakou, C.; Cai, Q. Single-Atom Catalysts as Promising Cathode Materials for Lithium–Sulfur Batteries. J. Phys. Chem. C 2021, 125, 18108–18118. [Google Scholar] [CrossRef]
  14. Su, Y.S.; Manthiram, A. Lithium–sulphur batteries with a microporous carbon paper as a bifunctional interlayer. Nat. Commun. 2012, 3, 1166. [Google Scholar] [CrossRef] [PubMed]
  15. Liang, X.; Rangom, Y.; Kwok, C.Y.; Pang, Q.; Nazar, L.F. Interwoven MXene nanosheet/carbon-nanotube composites as Li–S cathode hosts. Adv. Mater. 2017, 29, 1603040. [Google Scholar] [CrossRef] [PubMed]
  16. Li, H.; Ma, S.; Cai, H.; Zhou, H.; Huang, Z.; Hou, Z.; Wu, J.; Yang, W.; Yi, H.; Fu, C. Ultra-thin Fe3C nanosheets promote the adsorption and conversion of polysulfides in lithium-sulfur batteries. Energy Storage Mater. 2019, 18, 338–348. [Google Scholar] [CrossRef]
  17. Sim, E.S.; Yi, G.S.; Je, M.; Lee, Y.; Chung, Y.C. Understanding the anchoring behavior of titanium carbide-based MXenes depending on the functional group in LiS batteries: A density functional theory study. J. Power Sources 2017, 342, 64–69. [Google Scholar] [CrossRef]
  18. Ghazi, Z.A.; He, X.; Khattak, A.M.; Khan, N.A.; Liang, B.; Iqbal, A.; Wang, J.; Sin, H.; Li, L.; Tang, Z. MoS2/celgard separator as efficient polysulfide barrier for long-life lithium–sulfur batteries. Adv. Mater. 2017, 29, 1606817. [Google Scholar] [CrossRef] [PubMed]
  19. Cheng, Z.; Xiao, Z.; Pan, H.; Wang, S.; Wang, R. Elastic Sandwich-Type rGO–VS2/S Composites with High Tap Density: Structural and Chemical Cooperativity Enabling Lithium–Sulfur Batteries with High Energy Density. Adv. Energy Mater. 2018, 8, 1702337. [Google Scholar] [CrossRef]
  20. Yang, X.F.; Wang, A.; Qiao, B.; Li, J.; Liu, J.; Zhang, T. Single-atom catalysts: A new frontier in heterogeneous catalysis. Acc. Chem. Res. 2013, 46, 1740–1748. [Google Scholar] [CrossRef]
  21. Han, Z.; Zhao, S.; Xiao, J.; Zhong, X.; Sheng, J.; Lv, W.; Zhang, Q.; Zhou, G.; Cheng, H.M. Engineering d-p orbital hybridization in single-atom metal-embedded three-dimensional electrodes for Li–S batteries. Adv. Mater. 2021, 33, 2105947. [Google Scholar] [CrossRef]
  22. Lin, H.; Yang, L.; Jiang, X.; Li, G.; Zhang, T.; Yao, Q.; Zheng, G.W.; Lee, J.Y. Electrocatalysis of polysulfide conversion by sulfur-deficient MoS 2 nanoflakes for lithium–sulfur batteries. Energy Environ. Sci. 2017, 10, 1476–1486. [Google Scholar] [CrossRef]
  23. Yuan, Z.; Peng, H.J.; Hou, T.Z.; Huang, J.Q.; Chen, C.M.; Wang, D.W.; Cheng, X.B.; Wei, F.; Zhang, Q. Powering lithium–sulfur battery performance by propelling polysulfide redox at sulfiphilic hosts. Nano Lett. 2016, 16, 519–527. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, T.; Liang, J.; Ye, S.; Zhang, Q.; Liu, J. Fundamental, application and opportunities of single atom catalysts for Li-S batteries. Energy Storage Mater. 2023, 55, 322–355. [Google Scholar] [CrossRef]
  25. Boyjoo, Y.; Shi, H.; Olsson, E.; Cai, Q.; Wu, Z.S.; Liu, J.; Lu, G.Q. Molecular-level design of pyrrhotite electrocatalyst decorated hierarchical porous carbon spheres as nanoreactors for lithium–sulfur batteries. Adv. Energy Mater. 2020, 10, 2000651. [Google Scholar] [CrossRef]
  26. Du, Z.; Chen, X.; Hu, W.; Chuang, C.; Xie, S.; Hu, A.; Yan, W.; Kong, X.; Wu, X.; Ji, H.; et al. Cobalt in nitrogen-doped graphene as single-atom catalyst for high-sulfur content lithium–sulfur batteries. J. Am. Chem. Soc. 2019, 141, 3977–3985. [Google Scholar] [CrossRef]
  27. Zhang, Q.; Wang, Y.; Seh, Z.W.; Fu, Z.; Zhang, R.; Cui, Y.l. Understanding the anchoring effect of two-dimensional layered materials for lithium–sulfur batteries. Nano Lett. 2015, 15, 3780–3786. [Google Scholar] [CrossRef]
  28. He, J.; Hartmann, G.; Lee, M.; Hwang, G.S.; Chen, Y.; Manthiram, A. Freestanding 1T MoS2/graphene heterostructures as a highly efficient electrocatalyst for lithium polysulfides in Li–S batteries. Energy Environ. Sci. 2019, 12, 344–350. [Google Scholar] [CrossRef]
  29. Ren, Y.; Wang, B.; Liu, H.; Wu, H.; Bian, H.; Ma, Y.; Lu, H.; Tang, S.; Meng, X. CoP nanocages intercalated MXene nanosheets as a bifunctional mediator for suppressing polysulfide shuttling and dendritic growth in lithium-sulfur batteries. Chem. Eng. J. 2022, 450, 138046. [Google Scholar] [CrossRef]
  30. Park, J.; Yu, B.C.; Park, J.S.; Choi, J.W.; Kim, C.; Sung, Y.E.; Goodenough, J.B. Tungsten disulfide catalysts supported on a carbon cloth interlayer for high performance Li–S battery. Adv. Energy Mater. 2017, 7, 1602567. [Google Scholar] [CrossRef]
  31. Lin, H.; Zhang, S.; Zhang, T.; Cao, S.; Ye, H.; Yao, Q.; Zheng, G.W.; Lee, J.Y. A cathode-integrated sulfur-deficient Co9S8 catalytic interlayer for the reutilization of “lost” polysulfides in lithium–sulfur batteries. ACS Nano 2019, 13, 7073–7082. [Google Scholar] [CrossRef] [PubMed]
  32. Xu, J.; Zhang, W.; Fan, H.; Cheng, F.; Su, D.; Wang, G. Promoting lithium polysulfide/sulfide redox kinetics by the catalyzing of zinc sulfide for high performance lithium-sulfur battery. Nano Energy 2018, 51, 73–82. [Google Scholar] [CrossRef]
  33. Kresse, G.; Hafner, J.J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, J. Catalysis by supported single metal atoms. ACS Catal. 2017, 7, 34–59. [Google Scholar] [CrossRef]
  35. 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. [Google Scholar] [CrossRef]
  36. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, Z.; Zhou, L.; Ge, Q.; Chen, R.; Ni, M.; Utetiwabo, W.; Zhang, X.; Yang, W. Atomic Iron Catalysis of Polysulfide Conversion in Lithium–Sulfur Batteries. ACS Appl. Mater. Interfaces 2018, 10, 19311–19317. [Google Scholar] [CrossRef]
  38. Zeng, P.; Yuan, C.; Liu, G.; Gao, J.; Li, Y.; Zhang, L. Recent progress in electronic modulation of electrocatalysts for high-efficient polysulfide conversion of Li-S batteries. Chin. J. Catal. 2022, 43, 2946–2965. [Google Scholar] [CrossRef]
  39. Henkelman, G.; Uberuaga, B.P.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113, 9901–9904. [Google Scholar] [CrossRef]
  40. Zhou, L.; Danilov, D.L.; Qiao, F.; Wang, J.; Li, H.; Eichel, R.A.; Notten, P.H. Sulfur reduction reaction in lithium–sulfur batteries: Mechanisms, catalysts, and characterization. Adv. Energy Mater. 2022, 12, 2202094. [Google Scholar] [CrossRef]
Figure 1. (a) Top and side views of VS2 monolayer with three different adsorption sites for single atoms. (b) The adsorption energies of single atoms at H-V and H-S sites. The dashed line indicates the equal adsorption energies of H-V and H-S sites.
Figure 1. (a) Top and side views of VS2 monolayer with three different adsorption sites for single atoms. (b) The adsorption energies of single atoms at H-V and H-S sites. The dashed line indicates the equal adsorption energies of H-V and H-S sites.
Nanomaterials 14 00692 g001
Figure 2. PDOS of VS2 and TM@VS2 (TM = Hf, Nb, Pd, Ta, Y, Zr, or Rh, where the d orbitals of each TM atom are multiplied by 10).
Figure 2. PDOS of VS2 and TM@VS2 (TM = Hf, Nb, Pd, Ta, Y, Zr, or Rh, where the d orbitals of each TM atom are multiplied by 10).
Nanomaterials 14 00692 g002
Figure 3. (a) The difference in adsorption energies between Li2Sn/S8 on VS2 and TM@VS2. The adsorption configurations of Li2Sn/S8 on (b) VS2 and (c) Ta@VS2.
Figure 3. (a) The difference in adsorption energies between Li2Sn/S8 on VS2 and TM@VS2. The adsorption configurations of Li2Sn/S8 on (b) VS2 and (c) Ta@VS2.
Nanomaterials 14 00692 g003
Figure 4. Gibbs free energy of S8-to-Li2S reaction on VS2 and TM@VS2 (TM = Y, Zr, Hf, Nb, Ta. The * was stand for absorbed).
Figure 4. Gibbs free energy of S8-to-Li2S reaction on VS2 and TM@VS2 (TM = Y, Zr, Hf, Nb, Ta. The * was stand for absorbed).
Nanomaterials 14 00692 g004
Figure 5. (a) Decomposition energy barriers of Li2S on VS2 and TM@VS2. Li2S decomposition pathway on (b) Nb@VS2, Hf@VS2, Pd@VS2, and VS2.
Figure 5. (a) Decomposition energy barriers of Li2S on VS2 and TM@VS2. Li2S decomposition pathway on (b) Nb@VS2, Hf@VS2, Pd@VS2, and VS2.
Nanomaterials 14 00692 g005
Figure 6. (a) COHP diagram of the TM–S and Li–S bonds in Li2S on Ta@VS2 and Y@VS2. (b) The Li2S adsorption configurations on Ta@VS2 and Y@VS2.
Figure 6. (a) COHP diagram of the TM–S and Li–S bonds in Li2S on Ta@VS2 and Y@VS2. (b) The Li2S adsorption configurations on Ta@VS2 and Y@VS2.
Nanomaterials 14 00692 g006
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

Li, Y.; Liu, Y.; Zhang, J.; Wang, D.; Xu, J. Rational Design of Non-Noble Metal Single-Atom Catalysts in Lithium–Sulfur Batteries through First Principles Calculations. Nanomaterials 2024, 14, 692. https://doi.org/10.3390/nano14080692

AMA Style

Li Y, Liu Y, Zhang J, Wang D, Xu J. Rational Design of Non-Noble Metal Single-Atom Catalysts in Lithium–Sulfur Batteries through First Principles Calculations. Nanomaterials. 2024; 14(8):692. https://doi.org/10.3390/nano14080692

Chicago/Turabian Style

Li, Yang, Yao Liu, Jinhui Zhang, Dashuai Wang, and Jing Xu. 2024. "Rational Design of Non-Noble Metal Single-Atom Catalysts in Lithium–Sulfur Batteries through First Principles Calculations" Nanomaterials 14, no. 8: 692. https://doi.org/10.3390/nano14080692

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