Thi Nhan Trana,
Nguyen Vo Anh Duyb,
Nguyen Hoang Hieuc,
Truc Anh Nguyend,
Nguyen To Vane,
Thi Viet Bac Phungf,
Yohandys A. Zuluetag,
Minh Tho Nguyenhi,
Peter Schallj and
Minh Triet Dang*c
aFaculty of Fundamental Sciences, Hanoi University of Industry, 298 Cau Dien, Bac Tu Liem, Hanoi, Vietnam
bFPT University, 600 Nguyen Van Cu, Ninh Kieu, Can Tho, Vietnam
cDepartment of Physics Education, Can Tho University, 3/2 Street, Ninh Kieu, Can Tho, Vietnam. E-mail: dmtriet@ctu.edu.vn
dFaculty of Mechanics, Can Tho University of Technology, 256 Nguyen Van Cu Street, Ninh Kieu, Can Tho, Vietnam
eFaculty of Chemico-Physical Engineering, Le Quy Don Technical University, Ha Noi, Vietnam
fCenter for Environmental Intelligence and College of Engineering & Computer Science, VinUniversity, Hanoi, Vietnam
gDepartamento de Física, Facultad de Ciencias Naturales y Exactas, Universidad de Oriente, Santiago de Cuba, CP 90500, Cuba
hLaboratory for Chemical Computation and Modeling, Institute for Computational Science and Artificial Intelligence, Van Lang University, Ho Chi Minh City, Vietnam
iFaculty of Applied Technology, School of Technology, Van Lang University, Ho Chi Minh City, Vietnam
jVan der Waals-Zeeman Institute, University of Amsterdam, Science Park 904, Amsterdam, The Netherlands
First published on 22nd October 2024
Enhancement of the ionic conductivity and reduction of diffusion barriers of lithium-ion batteries are crucial for improving the performance of the fast-growing energy storage devices. Recently, the fast-charging capability of commercial-like lithium-ion anodes with the smallest modification of the current manufacturing technology has been of great interest. We used first principles methods computations with density functional theory and the climbing image-nudged elastic band method to evaluate the impact of an external electric field on the stability, electronic band gap, ionic conductivity, and lithium-ion diffusion coefficient of penta-graphene nanoribbons upon lithium adsorption. By adsorbing a lithium atom, these semiconductor nanoribbons become metal with a formation energy of −0.22 eV, and an applied electric field perpendicular to the surface of these nanoribbons further stabilizes the structure of these lithium-ion systems. Using the Nernst–Einstein relation, in the absence of an electric field, the ionic conductivity of these penta-graphene nanoribbons amounts to 1.24 × 10−4 S cm−1. In the presence of an electric field, this conductivity can reach a maximum value of 8.89 × 10−2 S cm−1, emphasizing the promising role of an electric field for supporting fast-charging capability. Our results highlight the role of an external electric field as a novel switch to improve the efficiency of lithium-ion batteries with penta-graphene nanoribbon electrodes and open a new horizon for the use of pentagonal materials as anode materials in the lithium-ion battery industry.
Typical LIBs consist of both anode and cathode electrodes, a separator, and all immersed electrolyte solution. To fulfil the aforementioned requirements, the critical task is to improve the performance of cathode and/or anode electrode materials. Several commercial types of cathode electrode materials for LIBs are LiCoO2 (LCO),3,4 LiMnO2 (LMO),5,6 LiFePO4 (LPO)7–9 and Li[NixMnyCoz]O2 (NMC).10 For long lifetime and large energy storage LIBs, the use of nickel-rich NMC layered cathodes can be considered as an excellent approach to reconcile the requirement of high specific discharge capacity, reasonable durability and working voltages, and affordable production cost.11 In the case of anode materials, the most common electrode material for commercial lithium-ion rechargeable batteries is graphite.12 The advantages of graphite electrodes are their low price, large reserves, and high conductivity. However, a disadvantage is its low maximum storage capacity (372 mA h g−1) due to the very flat surface and the fact that the charge/discharge rate tends to decrease rapidly after each charge/discharge cycle because of the formation of solid–electrolyte interface (SEI).13 Therefore, the search for new electrode materials that possess the superior properties of graphite electrodes but can overcome their inherent disadvantages (storage capacity and ionic diffusion ability) is a major challenge in the lithium-ion rechargeable battery industry.
In 2015, Zhang et al. theoretically proposed two-dimensional T12-carbon penta-graphene (PG) as a completely new type of hypothetical carbon material assembled from carbon atoms with pentagonal patterns.14 These PG sheets have intrinsic indirect bandgap of 3.25 eV, predicted by computations using the Heyd–Scuseria–Ernzerhof (HSE06) exchange–correlation functional of density functional theory, and are thermally stable at temperatures up to 1000 K. In 2016, using density functional theory and molecular dynamics simulations, Xiao et al. theoretically showed that albeit penta-graphene sheets are mechanically less stable than graphite monolayers, their maximum storage capacity and average open circuit voltage (OCV) and electron diffusion barrier amount to 1489 mA h g−1, 0.55 V and 0.17 eV, respectively, making them outstanding candidates for lithium-ion battery anodes.15 By cutting the PG sheets along different crystallographic directions, we can obtain penta-graphene nanoribbons (PGNR).16 These nanoribbons exhibit a high carrier mobility ranging from ∼101 to ∼104 cm2 V−1 s−1 at room temperature depending on the edge-terminated atoms.17 With hydrogen termination at the edges, the PGNR behaves as a semiconductor with a bandgap of ∼2.4 eV17 and a low energy diffusion barrier of 0.40 eV.18 Furthermore, within this pentagonal world of materials, penta-silicene and penta-germanene have been synthesized successfully using the in situ molecular beam epitaxy on gold (110)19–22 and (111)23 surfaces, respectively. This demonstrates the potential synthesis success and application of penta-graphene materials on advanced electronic materials.
Recently, substantial progress has been made to meet the fast-charging requirement by applying external magnetic24,25 or electric fields.26 Recent experimental reports demonstrate that the charge rates of LIBs can significantly be improved by imposing an external magnetic field parallel to the material surface to reduce the Li-ion transmission paths and increase the diffusion coefficients.25 For instance, by applying an external electric field on lithium hexafluorophosphate (LiPF6) in ethylene carbonate, Kumar and Seminario27 showed a strong impact of the electric field on the mobility of lithium ions by improving the drift velocities in ethylene carbonate electrolyte. Specifically, while the diffusion coefficient is increased linearly under a small applied electric field with magnitude below 2 V nm−1, it turns out to increase exponentially when an electric field strength is above that threshold.27 Besides the common use of the electric fields on the electrolytes, more interestingly, by constructing heterojunctions assembled by nanomaterials with different bandgaps, recent experimental and theoretical reports28–31 demonstrated the ability for building built-in electric field across the material interfaces to boost the electron and/or ion mobility. For instance, by creating a built-in electric field across the p–n junction of P-Co3O4/TiO2 anode, Kong et al. obtained a remarkable rate capability of 801 mA h g−1 even after 1600 cycles at 2 A g−1.28 Fast-charging anodes with high ionic conductivity can be controlled effectively under an external electrical field.
In this context, we set out to employ first-principles methods computations with density functional theory and the climbing image nudged elastic band (CI-NEB) method to investigate the Li-ion adsorption, electronic band structure modification, and ionic diffusion and conductivity for PGNR anodes in lithium-ion batteries. The CI-NEB32 has been known as an effective method to find the minimum reaction paths and determine the corresponding energy barriers between the given initial and final configurations. During optimization, this method creates a set of images to connect the initial and final states by performing a linear interpolation between the neighboring images to find the highest energy image of the transition state structure. In this paper, we report on the greatly improved electrochemical properties of penta-graphene anodes resulting from the change in structural stability, density of states, ionic diffusion and transition energy following application of electric fields. Such an improvement points out the potential application of penta-graphene nanoribbons as anode materials for the next generation of lithium-ion rechargeable batteries.
To determine the most stable of PGNR with a single Li atom, we place the Li atom 3 Å above the top layer of the PGNR at all possible adsorption sites as shown by letters A–I in Fig. 1(a). The stability of the adsorption configurations is evaluated by the formation energy as follows40
For further analyses of orbital hybridization, we perform the periodic energy decomposition analyses (PEDA) using the BP86-D3 exchange–correlation potential with the same basic sets in the AMS package. Since the HSE06 functional is not yet implemented in the AMS package for decomposing the interaction energies, we use the BP86-D3 functional, which is identical to the exchange–correlation optB86b-vdW functional in VASP package.
To investigate the adsorption of Li atoms, we evaluate the structural stability upon adsorption of a single Li atom on nine possible adsorption sites of the most stable PGNR configuration (Fig. 1(a)). Table 1 presents the formation energy Eads and the corresponding adsorption distances of all possible adsorption sites. The adsorption distance is the center-to-center distance from the adsorbed Li atom to the nearest C atoms of the PGNR. As shown in Table 1, the lowest adsorption energy of all possible adsorption sites is −0.220 eV at the D and E sites. Furthermore, after equilibration, irrespective of the initial configurations at D or E sites, the adsorbed Li atom relaxes on top of the sp3 hybridized C atom, illustrating that the sp3 configuration is the most energetical stable one. The fundamental phonon dispersion characteristic remains almost unchanged for these nanoribbons upon lithium adsorption (Fig. S1b†). Thus, for the further calculations, we associate the PGNR with a Li atom on top of the sp3 C atom as the most stable configuration.
Site | A | B | C | D | E | F | G | H | I |
---|---|---|---|---|---|---|---|---|---|
Eads (eV) | −0.219 | −0.219 | −0.216 | −0.220 | −0.220 | −0.219 | −0.218 | −0.218 | −0.219 |
a (Å) | 1.345 | 1.444 | 1.943 | 1.498 | 1.491 | 1.014 | 1.026 | 1.031 | 1.148 |
We now apply an electric field perpendicular to the surface of the penta-graphene nanoribbons. The field is applied to all atoms in the adsorbed systems via the Kohn–Sham equations. This artificial field is to mimic the built-in electric field, i.e., generated in between the p–n junction of a heterostructure anode as reported in ref. 28. Note that since this electric field is perpendicular to the surface of the ribbons, this force will not perform work to the moving carriers parallel to the surface of the ribbons. The strength of the applied field varies from −2 V nm−1 to 2 V nm−1 with a stepsize of 1 V nm−1, where positive (negative) values indicate a field in the positive (negative) z-direction. This field is strong enough to disturb the electron diffusion process but is weak enough not to destroy the adsorption system. Under a positive electric field, the adsorbed Li atom is relaxed on top of the sp2 hybridized carbon atom as indicated in Fig. 2. However, under a negative electric field, the adsorbed Li atom anchors close to the top of the sp3 hybridized carbon atom. To evaluate the stability of these configurations, we calculate the relative energy difference which is the energy difference between the systems with the electric field and the one without it. As shown in Table 2, the adsorption systems are slightly more stable with applied electric field than without, as indicated by the negative values of the relative energy difference. Apart from that, the lattice constant in the x-direction and the buckling distance (being the furthest distance between two carbon atoms in the z direction of the PGNR) of the systems show minor changes with the largest buckling of 9.75%.
Electric field strength (V nm−1) | −2 | −1 | 0 | 1 | 2 |
---|---|---|---|---|---|
Relative energy difference (eV) | −0.018 | −0.162 | 0 | −0.084 | −0.169 |
Lattice constant in x-direction (Å) | 10.854 | 10.854 | 10.857 | 10.854 | 10.855 |
Buckling distance (Å) | 0.642 | 0.743 | 0.677 | 0.712 | 0.667 |
Although there is a presence of a negative dispersion band when these nanoribbons are placed under a positive electric field (Fig. S1c and d†), in general, these ribbons are mechanically stable under these small electric fields. The most interesting observation is that the Li atom is adsorbed more closely to the PGNR when a positive electric field is applied. As shown in Fig. 2, the adsorption distance from the Li atom to the PGNR reduces by about 30% when the field is applied in the z-direction, while for fields in the opposite direction, this quantity extends by about 30% with respect to the field-less case. This implies significant changes of the ionic diffusion and conductivity of the adsorbed systems under an external electric field, as discussed in the next section.
Fig. 3 Band structure and partial density of states of penta-graphene nanoribbons calculated in the AMS package. The Fermi level indicated by the dashed lines is set to zero. |
Upon lithium adsorption, the semiconductor PGNR becomes metal as indicated in Fig. 4, in line with previous reports.14,48 This semiconductor-to-metal transition demonstrates that PGNR is a good anode material for lithium-ion batteries due to creation of free charges from the interaction between the adsorbed Li atoms and C atoms of the PGNR. The energy states around the Fermi level are driven by the π electrons of carbon atoms as displayed in Fig. S4 (ESI file†), demonstrating a charge transfer from the adsorbed Li atom to the PGNR as pointed out in ref. 48 and a high possibility of hybridization between p-orbital of the C and s-orbital of the Li atom. The significant contribution of the Li-atom's s-orbital to the density of states close to the Fermi level cause an increase in the conductivity of PGNR upon lithium adsorption and lead to the semiconductor-to-metal transition. Additionally, the interaction with the Li atom can enhance the flexibility of p-orbital electrons of the C atoms, leading to a strong shift of the highest occupied molecular orbital (HOMO) towards the bottom of the conduction band (cf. Fig. 4 and S4, ESI file†). The electrons transferred from the Li atom, and the flexible p-orbital electrons of the C atoms occupying the HOMO may play an important role in increasing the electronic conductivity of the PGNR + Li systems.
A significant band modification of the PGNR + Li under an external electric field is clearly visible in Fig. 4. Under a 2 V nm−1 electric field, we observe a considerable down-shift of the band dispersion curves in the conduction band towards the Fermi level while this trend is less pronounced in the case of a −2 V nm−1 electric field. Similar behavior is also observed in Fig. S4 (ESI file†) obtained from the VASP package. Furthermore, under a positive applied electric field, a new state appears in the conduction band just above the Fermi level. This state is occupied by s-orbitals of the adsorbed Li atoms, increasing its pDOS compared to nearby states (Fig. S4†) further increasing the conductivity of PGNR + Li under positive electric fields. We would stress that these modifications of the band structure and pDOS are different from those under negative electric field; hence, the applied field direction exerts a strong impact on the electronic characteristics of the PGNR + Li systems.
The electrons of frontier orbitals play an important role, in part governing the electronic properties of the material. To get further insights into the change in electronic properties of PGNR caused from the Li adsorption, and the effect of the external electric field, we probe the modification of the local electron distribution of HOMO and LUMO states of the material. Table 3 displays the shapes of frontier orbitals of pristine PGNR, PGNR + Li without and with an external electric field. For pristine PGNR (without applied external electric field), electrons of the HOMO and LUMO mainly localize around sp2 C atoms and distribute over the areas inside the PGNR like that observed for PG.14 The density of electrons of these two states around C atoms at the edge of the nanoribbon is inconsiderable, possibly, due to effect of the H passivation. Upon lithium adsorption, the distribution of electrons of the HOMO and LUMO exhibits a higher localization. They are mainly concentrated on the top of C atoms near the adsorbed Li atom with relative high density and could act as free electrons, leading to an increase of the electronic conductivity of PGNR upon adsorption. Such a redistribution of electrons of the frontier orbitals due to the Li adsorption also results the transition from the conductor-to-metal behavior of PGNR. As shown in Table 3, under an electric field, the LUMO states are more locally placed around the adsorbed Li atoms. Such a change of the LUMO orbital shapes implies a strong increase of the conductivity of PGNR + Li under the external electric field. Interestingly, Table 3 also shows that under a negative electric field, the electron depletion areas of frontier orbitals are replaced by the accumulation areas, whereas this phenomenon is not observed when applying a positive field, demonstrating the dependence of electronic properties of the PGNR + Li systems on the direction of the applied fields.
To further understand the influence of external electric fields on the electronic properties of the adsorbed systems, we perform periodic energy decomposition analyses (PEDA) to calculate the intrinsic bond energy ΔEint consisting of the electrostatic energy ΔEeslast, the Pauli repulsion energy ΔEPauli, and the orbital relaxation energy between the adsorbed lithium atom (the first fragment) and all other carbon atoms (the second fragment including 54 carbon and 12 hydrogen atoms) according to eqn (1):
ΔEint = ΔEeslast + ΔEPauli + ΔEorb, | (1) |
The computed intrinsic bond and partial energies are given in Table 4. Note that the relatively high energy values in Table 4 are summed over each fragment. These indicate that a weak external electric field can modify the intrinsic bond energies and their parts, by modifying the interaction between PGNR and Li atoms. Looking in more detail, the electrostatic energy emerges as the most dominant factor in the absence or presence of external electric fields, illustrating that the interaction between PGNR and Li is mainly driven by electrostatic effects, and the adsorbed Li atoms yield considerable electronic contributions to the PGNR. Looking at the orbital relaxation energies, which cover the orbital hybridization effect in the adsorption system, with respect to the zero electric field, the magnitudes of this quantity tend to decrease under a negative electric field and increase significantly under the positive field, again indicating the sensitivity of these adsorbed systems under an external electric field. The increase in electrostatic energy under positive electric field results from a charge redistribution of the frontier orbital and the nearby states as shown in Table 3 and Fig. S4 (ESI file†). Hence, the enhanced charge distribution of the adsorbent and adsorbate originates from charge transfer due to the reduced distance between the Li and PGNR as pointed out above. As a result, the diffusion coefficient of Li ions on PGNR under positive electric field is significantly higher than that without and under a negative field. On the other hand, the smaller orbital relaxation energy under negative electric fields, implies a weakening of the orbital hybridization between the adsorbed Li and the PGNR. Therefore, positive electric fields enhance the flexibility of the p-orbital electrons of C atoms; and these results anticipate an increase of the conductivity as well as the diffusion coefficient of Li ions on the adsorbed systems.
PEDA types | E = – 2 V nm−1 | E = 0 V nm−1 | E = 2 V nm−1 |
---|---|---|---|
ΔEint | −41.9 | −64.9 | −62.2 |
ΔEeslast | −318.5 | −418.3 | −670.2 |
ΔEorb | −183.0 | −249.4 | −367.6 |
Fig. 5 Diffusion barriers of lithium-atoms on penta-graphene nanoribbons under an external electric field. |
In the absence of an electric field, the diffusion barrier calculated by the CI-NEB method along the diffusion path highlighted in Fig. 5a is 0.272 eV, which is in quantitative agreement with that of two-dimensional penta-graphene (≤0.33 eV),15 comparable to the case of penta-hexagonal graphene (in between 0.21 eV and 0.32 eV)49 and significantly lower than those of commercial graphite anodes (0.4–0.6 eV)49 and LiFePO4 (1.02 eV) electrodes.50
In the presence of an external electric field, this diffusion barrier significantly reduces to the lowest value of ∼0.1 eV at 2 V nm−1 electric field as shown in Table 5 and Fig. 5. Such low diffusion barrier, which is comparable with that of Ti3C2 (ref. 49) and the recent proposed MoS2/penta-graphene heterostructures,51 strongly demonstrates the remarkable fast ion diffusion of penta-graphene nanoribbons under an applied electric field.
Electric field strength (V m−1) | −2 | −1 | 0 | 1 | 2 |
---|---|---|---|---|---|
Diffusion barrier (eV) | 0.144 | 0.136 | 0.272 | 0.119 | 0.102 |
Diffusion coefficient (cm2 s−1) | 4.5 × 10−4 | 6.2 × 10−4 | 3.2 × 10−6 | 1.2 × 10−3 | 2.3 × 10−3 |
The lithium diffusion coefficient ratio between penta-graphene nanoribbons and graphitic carbon55 | 102.78 | 139.99 | 0.73 | 269.98 | 520.71 |
Conductivity (S cm−1) | 1.74 × 10−2 | 2.40 × 10−2 | 1.24 × 10−4 | 4.64 × 10−2 | 8.89 × 10−2 |
The diffusion coefficient of lithium-ions can be calculated via the relation (2)
(2) |
σ(T) = HVNq2D(T)/kBT | (3) |
As shown in Table 5 and in the absence of an electric field, the lithium-ion diffusion coefficient and conductivity of penta-graphene nanoribbons are 3.2 × 10−6 cm2 s−1 and 1.24 × 10−4 S cm−1, respectively, while under an external electric field, these quantities increase significantly. Specifically, under a positive electric field of 1 V nm−1 and 2 V nm−1, it reaches to 1.2 × 10−3 cm2 s−1 and 2.3 × 10−3 cm2 s−1, which is approximately about ∼375 and ∼719 times faster than that of the case of zero electric field, respectively. Under negative electric fields, though we also observe an increase of diffusion coefficients, these quantities are relatively smaller than in the case of positive electric fields. These observations are in line with the additional electronic states formed in the conduction bands close to the Fermi level (Fig. 4).
A similar behavior is also observed in the case of ionic conductivity. To demonstrate the outperformance of PGNR as a promising material for lithium-ion anodes, we compare our calculated diffusion coefficients with those of carbon graphite layers, a common electrode material of commercial batteries.55 Table 5 illustrates that the diffusion coefficient of PGNR under a zero electric field is slightly lower that of graphitic carbon layers but reaches ∼521 times higher under a 2 V nm−1 electric field. Therefore, the external electric field can be used as a novel switch to remarkably improve the charge/discharge rate of penta-graphene lithium-ion anodes.
Besides the diffusion barrier, diffusion coefficient and conductivity, another essential factor which greatly contributes to the charge/discharge rates of lithium-ion anodes is the electronic conductivity, which can be inferred by carrier effective mass.56
The effective masses of electrons and holes are calculated from eqn (4)
(4) |
While under the negative electric fields, these ratios increase a few times, in the opposite electric field directions, we observe remarkable increases up to ∼26 times higher (see Table 6). Obviously, this sudden increase of the ratios highlights the role of an external electric field as a novel switch to improve the conductivity of penta-graphene lithium-ion anodes.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra05464d |
This journal is © The Royal Society of Chemistry 2024 |