Next Article in Journal
Enhanced Optical Transmission through a Hybrid Bull’s Eye Structure Integrated with a Silicon Hemisphere
Next Article in Special Issue
Recent Understanding in the Chemical Vapor Deposition of Multilayer Graphene: Controlling Uniformity, Thickness, and Stacking Configuration
Previous Article in Journal
Combination of CNTs with Classical Drugs for Treatment in Human Colorectal Adenocarcinoma (HT-29) Cell Line
Previous Article in Special Issue
Optically Active Telecom Defects in MoTe2 Fewlayers at Room Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Theoretical Studies on the Quantum Capacitance of Two-Dimensional Electrode Materials for Supercapacitors

College of Physics, Changchun Normal University, Changchun 130032, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(13), 1932; https://doi.org/10.3390/nano13131932
Submission received: 30 May 2023 / Revised: 15 June 2023 / Accepted: 19 June 2023 / Published: 25 June 2023

Abstract

:
In recent years, supercapacitors have been widely used in the fields of energy, transportation, and industry. Among them, electrical double-layer capacitors (EDLCs) have attracted attention because of their dramatically high power density. With the rapid development of computational methods, theoretical studies on the physical and chemical properties of electrode materials have provided important support for the preparation of EDLCs with higher performance. Besides the widely studied double-layer capacitance (CD), quantum capacitance (CQ), which has long been ignored, is another important factor to improve the total capacitance (CT) of an electrode. In this paper, we survey the recent theoretical progress on the CQ of two-dimensional (2D) electrode materials in EDLCs and classify the electrode materials mainly into graphene-like 2D main group elements and compounds, transition metal carbides/nitrides (MXenes), and transition metal dichalcogenides (TMDs). In addition, we summarize the influence of different modification routes (including doping, metal-adsorption, vacancy, and surface functionalization) on the CQ characteristics in the voltage range of ±0.6 V. Finally, we discuss the current difficulties in the theoretical study of supercapacitor electrode materials and provide our outlook on the future development of EDLCs in the field of energy storage.

1. Introduction

In today’s world, non-renewable energy sources are decreasing. Energy supply is closely related to environmental issues and basic human needs. Improving the conversion efficiency of energy sources and developing new energy sources has become an urgent problem that needs to be solved in the midst of the energy crisis [1]. In this energy-dependent world, electrochemical devices for energy storage have played a crucial role in overcoming the depletion of fossil fuels [2]. Compared with conventional batteries, supercapacitors have the advantages of high power density, long cycle life, and fast charge and discharge rates. However, the energy density of supercapacitors is usually low, which is a major obstacle to their development [3]. Supercapacitors are usually classified into three types: (1) electrical double-layer capacitors (EDLCs) with ion adsorption through the electrode surface; (2) pseudocapacitors with surface Faraday redox reactions on the electrodes; (3) hybrid supercapacitors that are a mixture of the above two [4,5]. In this paper, we mainly focus on the electrode materials for EDLCs.
For the charging process on EDLCs, the anions and cations in the electrolyte are adsorbed to the positive and negative surfaces, respectively, forming a double-layer due to the external voltage difference. After charging, the anions on the double-layer produce a potential difference between the two plates to store energy. Discharging is the opposite process of charging. Because of the fast rate of this simple physical adsorption process, the EDLCs usually have high power densities. For the electrode materials, carbon materials with a large specific surface area and good electrical conductivity are generally the best choice for the fabrication of EDLCs [4,6,7,8].
EDLCs have high output power, fast charge and discharge rates, and long service lives but poor energy density [9,10,11]. Therefore, increasing the energy density of EDLCs has become a key research goal. The energy density of EDLCs is determined by the operating voltage and the specific capacitance of the electrode/electrolyte system [12]. The total interface capacitance (CT) of EDLCs is related to the quantum capacitance (CQ) and the double-layer capacitance (CD), with the expression of 1/CT = 1/CQ + 1/CD [13,14,15,16,17]. CQ, also known as electrode capacitance, reflects the finite quantum state process of the electron-filled system [18,19]. The theoretical prediction of increasing the total EDLC capacitance by increasing the CQ of the electrode material has been experimentally confirmed [20]. CQ is proportional to the density of electronic states. A large number of quantum states near the Fermi level can lead to a high CQ. The electronic structure of a material can be modified by changing the dopants, functional groups, defects, etc. of the structure, thus changing the specific surface area and surface morphology of the electrode material. The larger the specific surface area, the better the energy storage performance of the electrode material [21,22].
Usually, materials with a thickness of a few atomic layers are considered as two-dimensional (2D) materials [3]. Since the discovery of graphene in 2004, 2D materials have experienced rapid development. For example, 2D carbon materials demonstrate excellent properties, such as high specific surface area and high electrical conductivity [23]. Since the working mechanism of EDLCs is an electrostatic effect, the anions and cations on the electrode material surface move to the positive and negative electrodes during the charging and discharging process, forming electric double-layers at the interface. Thus, 2D electrode materials with larger specific surface areas are more suitable for EDLCs than their 3D counterparts [24,25,26,27]. However, 2D electrode materials are highly susceptible to stacking due to their high edge activity, resulting in a decreased specific surface area and capacity during reagent application. Numerous studies have been devoted to maintaining or increasing the specific surface area of the electrode materials and improving the circulation rate of anions and cations by changing the morphology of the electrode surface. In this paper, we focus on the aforementioned modification measures (defects, doping changes, adsorption of functional groups, etc.) and atomic exchange on the electrode materials. We summarize the classification of electrode materials and highlight the materials with better performance and greater potential for experimental application.

2. Theoretical Basis

Among many low-dimensional materials, the differential quantum capacitance (Cdiff) can be defined by
C d i f f = d σ d Φ G = e 2 D O S V e
in which and G represent the differential charge density and differential local potential, respectively.
Thus the magnitude of Cdiff is dependent on the density of states. The density of states is essentially the number of different states that an electron is allowed to occupy at a given energy level, i.e., the number of electron states per unit volume of energy. Due to quantum confinement effects and the limitation of the low density of states, the significant movement of Fermi levels in two-dimensional materials could accumulate a sufficient number of carriers to provide better energy density, thus improving the performance of supercapacitors.
The excess charge density can be expressed as:
Q = + D E f E f E e ϕ G d E
where D(E) represents the density of states of the system, ƒ(E) is the Fermi-Dirac distribution function, E is the electronic energy with respect to the Fermi level, and e is the fundamental charge. For 2D materials, Cdiff can be obtained by the following equation:
C d i f f = e 2 + D E F T E e ϕ G d E
where FT(E) is the thermal spreading function, which is obtained from the following:
F T E = 4 κ Β Τ 1 s e c h 2 E 2 κ Β Τ
It is also common for researchers to analyze the energy storage capacity of supercapacitors by calculating Cint. The Cint is obtained by integrating the Cdiff over the charge and discharge cycles [28,29].
C i n t ( V ) = Q V = 1 V e 0 V C d i f f ( V ) d V
In this paper, Cdiff is equally defined as CQ.

3. Research Progress of 2D Electrode Materials for EDLCs

3.1. Graphene-like 2D Main Group Elements and Compounds

Graphene with sp2 hybridization is a typical representative of 2D materials. [30] In the past decade, graphene-based electrode materials have become a popular research direction for supercapacitor electrode materials. More recently, scientists have tried to discover various analogs with six-membered ring structures that are similar to graphene, such as silicene, germanene, phosphorene, etc. On the other hand, the 2D main group compounds with graphene-like structures, such as 2D carbon nitride (CN), have also been synthesized in recent years. There are also many studies on the performance of 2D CNs as electrode materials.

3.1.1. Graphene

Graphene is prone to stacking due to its high edge activity, resulting in a decrease in specific surface area and capacity. Since graphene oxides are usually used as precursors for graphene preparation, there are many vacancies in graphene-based materials. Many studies have focused on maintaining the dispersion of graphene [31,32].
In 2009, Xia et al. measured the CQ of monolayer and bilayer graphene, and the curve of CQ-potential was symmetrical and v-shaped (Figure 1a) [33]. Many scientists have experimentally demonstrated that the capacitance of carbon electrodes can be improved by doping nitrogen (N) atoms or the functionalization of N-containing groups [34,35,36,37,38,39,40,41]. Zhang et al. demonstrated that N-doping changes the electronic structure of graphene and increases the carrier density, which changes the CQ and leads to an increase in the interfacial capacitance value (Figure 1b) [42]. Yang et al. theoretically investigated the effects of N-doping configuration, N-doping concentration, vacant concentration, and transition metal atoms (Cu, Ag, Au) adsorption on the electronic structure and CQ of graphene [14]. Their results show that N-doping, vacancy defects, and transition metal atom adsorption can significantly enhance the CQ of graphene. Among them, the maximum value of CQ increases from 32.68 to 113.1 μF/cm2 as the N-doping concentration increases from 1.4% to 12.5% (Figure 1c,d). Mousavi-Khoshdel et al. investigated the changes in CQ of functionalized graphene with monovalent functional groups (-C6H5, -C6H4NH2, -C6H4NO2, -NH2) and divalent functional groups (-C6H4, -C6H2F2, -C6H2Cl2, -C6H3CH3) [43]. Their results show that the CQ values of functionalized graphene are higher than that of pristine graphene in both cases. A schematic diagram of the structures of the three types of groups is shown in Figure 1e. Chen et al. investigated the interaction and CQ of N and S co-doped graphene [44]. The maximum CQ of pristine graphene is 14 μF/cm2 and the minimum CQ is 2.5 μF/cm2. The CQ values of pyridine-N-doped graphene and pyrrolic-N-doped graphene at the Fermi level are about 41.4 and 38.2 μF/cm2, respectively. In a subsequent study, they found that the CQ value of N/S co-doped graphene could be higher than that of single N-doped graphene, with the highest CQ value being 95.8 μF/cm2. However, the CQ does not improve more when another N or S atom is added to the co-doped system (Figure 1f).
Hirunsit et al. studied the CQ variation of Al-, B-, N-, and P-doped single-vacancy (V) and multilayer graphene. [28] They showed that Al1, V, Al3V, and N3V modification can increase the CQ by a large amount (>40 mF/cm2), and the N3V structure showed the highest CQ value, which was 82.18 mF/cm2 (0.26 V). The construction of multilayer graphene also improves the CQ (Figure 1g–i). Hu et al. investigated the effect of transition metal (Mn, Fe, Co, Ni) and N atom (TMNx, x = 1–4) co-doping on the CQ of graphene [45]. The co-doped systems showed an increase in CQ, with a maximum value of 180.50 μF/cm2 for CoN2 g at −0.3 V (Figure 1j). A similar study was carried out by Wang et al., who explored the CQ changes of transition metals after in-plane doping and out-of-plane doping on graphene [46]. Their conclusions show that the CQ of in-plane doping is larger than that of out-of-plane doping, where the charge (Q) of Sc-doped graphene could reach 85 μC/cm2 at negative bias (Figure 1k). Song et al. studied the variation of CQ of epoxy (- O -)- and hydroxy (-OH)-modified graphene oxide [47]. The results show that the modified graphene oxide also has a higher CQ than the original structure. There is a significant increase of CQ with the increasing oxidation degree on both positive and negative bias (Figure 1l,m).
Sruthi et al. found that the CQ of graphene can be significantly enhanced by doping on the pristine graphene surface with N, Cl, and P atoms [48]. Additionally, very large CQ (>600 μF/cm2) can be achieved when doping N, Cl, and P atoms near room temperature. Xu et al. investigated the CQ of graphene doped/co-doped with B, N, P, S atoms and vacancy [49]. They also obtained the CD in a classical 1 M NaCl aqueous solution by using molecular dynamics simulations. Then, the CT was calculated. Graphene that has been 3N-doped with a single vacancy is supposedly the best candidate as an EDLC electrode (Figure 1n). Zhou et al. investigated the effects of doping (B, N, Al, Si, P, S), vacancies, and Stone–Wales defects on the CQ of graphene and found that Stone–Wales defects could also improve the CQ of graphene, but not better than doping or vacancy. The maximum CQ of Si-VG is 169.76 μF/cm2 at −0.29 V. The maximum CQ is 168.90 μF/cm2 at −0.06 V, when the VG concentration is 5.9% (Figure 1o,p). Zhang et al. determined a variety of materials suitable for supercapacitor applications by systematic calculations and generalizations [50]. They explored the CQ of 56 species of transition metal atoms and vacancy-doped/co-doped graphene, named TM@G and TM@VG, respectively (Figure 1q). Sruthi et al. explored the effect of different co-doping ratios on the CQ of graphene [51]. When the dopant ratio C:O:N is 50:8:4, the CQ of the system at the Fermi energy level can reach 423.73 μF/cm2 (Figure 1r).
One of the inevitable problems in manufacturing and using graphene materials is the stacking of layers, which significantly affects the structure the electrochemical properties. Cui et al. explored the effect of stacking on multilayered graphene [52]. They assumed a two-layer ab-stacked graphene model, where the top layer is defective and the bottom layer is perfect. They showed that the CQ of the pristine bilayer graphene increases linearly with voltage, reaching a maximum value of 37.7 μF/cm2 at 1.0 V. The peak of D2_III has a maximum CQ of 56.1 μF/cm2 at a voltage of 1 V (Figure 1s,t). Zhou et al. explored co-doping with N, P, S and transition metals (Ti, V, Cr, Mn, Co, Ni) in monolayer and multilayer graphene [53]. Their study showed that doping with transition metals (TM) improves the CQ more than co-doping with N, P, and S, and the Ti/Ni and N/P/S co-doped systems exhibit excellent CQ. However, the CQ of the multilayer system decreases due to the interactions between the adjacent layers of dopants. In a study by Zeng et al., it was found that the capacitance of B (N)-doped graphene as an anode (cathode) can reach a record CQ of 4317 F/g (6150 F/g) [54].

3.1.2. Silicene

Inspired by graphene, silicene is made from 2D layered nanosheets. Silicene sheets with different structures have been successfully synthesized on various substrates. Silicene with a buckling layer structure has a high surface area [55]. It is considered as an excellent anode material for Li-ion batteries because it has enough space to adsorb Li-ions and prevents structural breakage induced by the insertion of Li-ions. Similar to graphene, it is also expected to be one of the ideal electrodes for EDLCs.
Yang et al. explored the effects of vacancy and dopants (N, P, B, and S) concentration on the CQ of silicene [56]. Their results show that the maximum CQ of silicene increases with the defect concentration from 1.91 μF/cm2 at −0.38 V to 102.65 μF/cm2 at −0.19 V. When the pyridine-N doping concentration is 5.6%, the maximum CQ is 73.28 μF/cm2 (−0.07 V). The CQ is higher than that of the pristine silicene in all modified structures (Figure 2a). Momeni et al. explored the CQ of pristine silicene, defective silicene, and XSi3-like silicene (X = Al, B, C, N, P) structures [57]. Their results show that the alternative doped XSi3-like silicene structures have higher CQ compared to pristine silicene (CQ = 1200 F/g) and graphene (CQ = 500 F/g). The AlSi3 system reaches a maximum CQ of 2573 F/g under positive bias. They also showed that the large CQ of XSi3-like silicene originates from the high electronic states at the Fermi level of 2p and/or 3p orbitals of X and Si atoms, as evidenced by projected density of state analysis (Figure 2b,c). Xu et al. explored the CQ of silicene with metal atom (Ti, Au, Ag, Cu, and Al atoms) adsorption and single-vacancy doping. [58] It was found that a single vacancy with metal adsorption can significantly increase CQ. When the Ti concentration is increased from 2% to 12.5%, the maximum value of CQ increases from 52.2 μF/cm2 at −0.12 V to 132.2 μF/cm2 at 0.12 V (Figure 2d).

3.1.3. Germanene

Silicene and Germanene are of great interest as 2D layered nanosheet materials inspired by graphene. Germanene is more prominent than silicene and graphene in terms of its spin–orbit interaction. The large spin–orbit gap (24 meV) of germanene makes it a typical alternative material with the quantum spin Hall effect [59,60,61,62,63]. Moreover, germanene is more easily to be functionalized and has been synthesized by different chemical methods [64,65,66,67]. In order to further investigate the electrochemical properties of germanene and probe for more superior performance electrode materials, numerous researchers have investigated the CQ of germanene with doping, co-doping, and vacancy defects.
Si et al. explored the effect of single vacancy (SV), adsorption of Ti, Au, Ag, Cu, Al atoms, and different doping concentrations on the CQ of germanene [61]. Similar to graphene and silicene, vacancies can increase the CQ of germanene, especially in the positive bias range. The CQ of Ti- and Cu-doped SV germanene is superior to that of Au-, Ag-, and Al-doped ones. Moreover, Ti-doping is more stable in graphene, silicene, and germanene (Figure 2e). Zhou et al. found that transition metal (Ti, Cr, Mn, and Co) doping enhanced the CQ better than B/N/Al doping [68]. The maximum CQ can reach 91.47 μF/cm2 (0.2 V) for Ti-doping near the Fermi level. The co-doped system improves CQ more than single-doping (Figure 2f). Si et al. further investigated the effects of doping/co-doping, vacant defects, and multilayer structure on the electronic structure and CQ of germanene [69]. Their results show that N-doping can significantly improve the CQ of germanene. In a study of single and multilayered germanene co-doped with NAl, NNAl, NPAl, and NSAl, it was found that the interlayered interactions contributed more to the increase in CQ (Figure 2g).

3.1.4. Stanene

Stanene is a novel material that has received increasing attention in recent years. It has been successfully realized by epitaxial growth on Bi2Te3 (111) substrates [70,71,72]. Stanene exhibits several remarkable features, including large spin–orbit gaps, topological superconductivity, quantum anomalous Hall behavior, giant magnetoresistance, and efficient thermoelectricity [73]. Additionally, there have been numerous studies showing that measures such as doping with metal atoms can have a large effect on the structural and electrochemical properties of Stanene [74,75].
Zhou et al. verified the effect of vacancies and the single-doping and co-doping of light element atoms (B, N, Al, Si, P, S) and transition metals (Ti, V, Cr, Mn, Fe, Ni) on the geometry, electronic structure, and CQ of stanine [76]. Their results show that vacancy, doping, and co-doping can improve the CQ of stanene and that co-doped defective stanene exhibits better CQ at negative potentials than at positive bias, indicating that it can be used as a good anode material. The maximum CQ of BFeVSn is 76.52 μF/cm2 (0.29 V) under positive bias conditions (Figure 2h,i). Zhou et al. also investigated the effect of N/P/S and line co-doping with heavy metals (Ti, V, Fe, Ni) on stanine [77]. The effect of line co-doping to improve the CQ of the system is more obvious, where the maximum CQ at a positive bias of SSTiSn is 77.18 μF/cm2, which could be attributed to the increased electronic states of the Ti dopant and adjacent Sn atoms (Figure 2j–l).

3.1.5. Boronene

Due to the high carrier concentration, boronene has been used in plasma devices, extending the functionality to the visible region. Boronene is predicted to be an excellent candidate for Li-ion batteries due to its high Li capacity [78,79,80].
Kolavada et al. theoretically analyzed the CQ of δ-6 boronene with different layer numbers in aqueous electrolytes (AEs) and ionic liquid electrolytes (ILEs) [81]. In both AE and ILE systems, CQ enhances as the number of layers increases from 1 monolayer (ML) to 4 ML. When the number of layers is 4 ML, CQ can reach more than 600 μF/cm2 in both systems (Figure 2m).

3.1.6. Phosphorene

Phosphorene is a relatively new member to the group of 2D materials discussed in this study. Its strong in-plane anisotropy makes phosphorene a unique material for novel electronic devices [82,83,84,85,86]. Zu et al. fabricated supercapacitors by using phosphorene as electrodes and with the discharge capacity of 3181.5 F/g in a three-electrode configuration [87].
Ramesh et al. computationally examined the effect on phosphorene when half-metal (Si)-dopants, active-nonmetal (S)-dopants, and two transition metal (Ti, Ni) dopants replace the P atom [88]. The CQ of pristine phosphorene is approximately symmetric, with a minimum value of 2.27 μF/cm2 at the Fermi level. The CQ of all substitution systems is higher than that of the pristine phosphorene, with the highest CQ value of 92.1 μF/cm2 for Ti-doping at 0.4 V (Figure 2n).

3.1.7. Main Group Compounds

In recent years, four 2D carbon nitride (CN) structures, hg-C3N4, tg-C3N4, C2N, and C3N have been experimentally synthesized to further enrich the 2D electrode materials for supercapacitors [89,90,91]. These CN structures have high specific surface areas and excellent electrochemical stability. Therefore, CNs are considered good electrode materials.
Chen et al. investigated the effects of B and O doping on the electronic properties and CQ of 2D CNs. They found that doping with B or O could convert CNs from semiconductors to metals, thus improving the electrical conductivity [92]. The CQ values of B-doped CNs are all higher than those of B-doped monolayer graphene. The increased CQ can mainly be attributed to the strong hybridization between the dopant and the adjacent C and N atoms (Figure 2o,p).
Majdi et al. investigated the electrochemical properties of a new 2D Fe-doped boron carbide monolayer (FBC3ML) [93]. The maximum CQ of FBC3ML increases to 150.09 μF/cm2 compared to the original BC3ML, and the CQ-V curve becomes symmetric (Figure 2q).

3.2. Transition Metal Carbides or Nitrides (MXenes)

MXenes are 2D-layered materials derived from transition metal carbides, nitrides, or carbonitrides [94]. MXenes can be produced by selectively removing the A-layer from MAX phases, the 3D precursors of MXenes, noted as Mn+1AXn phases (n = 1, 2 and 3). MAX phases are generally divided into three types: 211, 312, and 413 structures. M denotes early transition metal elements (such as Sc, Ti, Zr, Hf, V, Nb, Ta, Cr, Mo, etc.); A denotes elements of group 13 or 14, such as Al or Si; and X refers to C, N, or their mixtures [2]. MXenes are often used in the field of energy storage because of their special physical and chemical properties.
The first MXene, Ti3C2, was isolated from Ti3AlC2 powder by immersing it in a hydrofluoric acid solution [94]. Subsequently, many MXene family members have been synthesized using selective etching methods and many new MXene structures have been theoretically predicted [2]. When performing chemical etching methods, the metal atoms on the MXene’s surface can easily react with -H, -O, -F, and -OH groups in solution and terminate them on the MXene’s surface, thus giving rise to functionalized MXenes, Mn+1XnTx, where T is the surface termination group [95,96,97]. This surface functionalization usually has an impact on the energy storage capacity of MXenes. Since the MAX phases usually have carbon vacancies (VC) [98,99], MXenes derived from the MAX phases are considered to have the same nature of carbon vacancies. The treatment of MXene materials via doping and vacancy also cause changes in the CQ of the materials. The current research on the CQ of MXene materials after modulation is more comprehensive. In this section, the effect of various modulation means on the CQ performance of MXene electrode materials will be discussed separately according to the different M elements.

3.2.1. Tin+1CnTx

As the first successfully prepared MXene, Ti3C2 and its isomer Ti2C have received increasing attention as electrode materials for supercapacitors [100,101].
Si et al. focused on the modulation of the two Ti-C MXenes materials using doping, vacancy, and adsorption methods [102]. Their calculations show the pristine structures have higher CQ compared to the functionalized Ti3C2 and Ti2C. The CQ of the functionalized structures decrease in an order of OH > F > H > O. The maximum CQ of OH groups adsorbed on Ti3C2 and Ti2C near the Fermi level is 264.414 μF/cm2 (0 V) and 276.960 μF/cm2 (−0.12 V) (Figure 3a,b). On the other hand, the adsorption of metal atoms on the surface of Ti3C2 and Ti2C can also change their CQ considerably. It is shown that on Ti3C2, the adsorption of Al atoms significantly increases the CQ (398.193 μF/cm2 at −0.072 V) due to the increase of the low potential local electronic states. A similar trend also appears for Ti2C, where Ti2C-Al has a maximum CQ of 444.192 μF/cm2 (0.312 V) near the Fermi level (Figure 3c,d). Furthermore, they have found that the adsorption of Ca atoms on Ti3C2F2 significantly improves the energy storage performance of the system with a CQ value of 488.153 μF/cm2. Nevertheless, the CQ of the Ti2C system shows no clear change (Figure 3e,f) [102].
Bafekry et al. investigated the oxygen vacancies in the Ti2CO2 monolayer and confirmed its semi-metallic properties by calculating the density of states of the system [103]. To further investigate the effect of O vacancy concentration on the properties of Ti2CO2, Su et al. systematically explored the CQ of Ti2CO2 with different O vacancy concentrations [104]. Their results demonstrate that the O vacancy concentration has a strong effect on CQ. The pristine Ti2CO2 has a low CQ under negative bias, with a maximum value of 4204.4 μF/cm2 (0.53 V). When one oxygen vacancy (5.56%) is introduced, the maximum CQ values increase to 4263.85 μF/cm2 (−0.1 V) and 4142.0 μF/cm2 (0.43 V) for negative and positive potentials, respectively. However, the maximum CQ decreases when two or three oxygen vacancies are introduced. The DOS of Ti2CO2 near the Fermi level mainly originates from O-pz and Ti-d orbitals. They speculate that the introduction of oxygen vacancies increases the charge transfer between adjacent O and Ti atoms (Figure 3g,h) [104]. While Li et al. theoretically investigated the CQ of Ti2CO2 monolayer with carbon vacancy line (CVL) [105]. The introduction of CVL improved the CQ and Q of the system under negative and positive bias within ±0.6 V. The CVL4 system improves the maximum CQ of 469.7 μF/cm2 under negative bias compared to the pristine Ti2CO2 monolayer (Figure 3i). Li et al. calculated the CQ of pristine Ti2CO2 (PT) and C-vacant Ti2CO2 (VT) monolayers adsorbed by Li atoms. [106] From their results, it can be seen that the maximum CQ of PT-LT monolayer is 10,993 μF/cm2 (0.48 V), the maximum CQ of VT monolayer is 7592 μF/cm2 (0.36 V), and the maximum CQ of VT-LC1 is 6866 μF/cm2 (0.54 V), within the positive bias voltage (Figure 3j,k).

3.2.2. Scn+1CnTx

Sc-based MXenes have the lightest M atom. Among them, Sc2CF2 is a semiconductor with strong anisotropic carrier mobility and thermal conductivity. The electron mobility of Sc2CF2 in the zigzag direction is almost four times higher than that of phosphorene in the armchair direction, and its thermal conductivity is higher than that of most low-dimensional metals and semiconductor materials. Due to its excellent properties in electronic devices, Sc2CF2 has received much attention in recent years [107,108].
Cui et al. studied the exchange defects of Sc/F, Sc/C, and C/F atoms (Figure 4a) [109]. Their study shows that the atomic exchange has little effect on the semiconducting properties. Additionally, the maximum CQ of pristine Sc2CF2 under negative bias is 739.39 μF/cm2 (−0.48 V). The atomic exchange of C/F atoms and C/Sc atoms reduces the maximum CQ of Sc2CF2 monolayer under negative bias to 488.60 μF/cm2 (−0.44 V) and 282.57 μF/cm2 (−0.48 V), respectively. The atomic exchange between F and Sc atoms increases the maximum CQ of Sc2CF2 monolayer under negative bias to 1037.76 μF/cm2 (−0.48 V). Cui et al. also studied the CQ of Sc2CT2 (T = F, P, Cl, Se, Br, O, Si, S, OH) monolayers (Figure 4b) [110]. They found that the maximum CQ of the Sc2C monolayer in aqueous electrolyte (±0.6 V) is 1025.01 μF/cm2 and 1297.03 μF/cm2 under negative and positive bias, respectively. Compared with the original Sc2C structure, the maximum CQ under the negative bias of Sc2CT2 (T=P, Cl, Se, Si) increases, with the maximum CQ of Sc2CP2 being 3800.34 μF/cm2. The maximum CQ of the Sc2CSi2 monolayer increases to 1708.82 μF/cm2 under positive bias voltage, but the maximum CQ of all other systems decreases under the positive bias voltage. However, all of them have a larger maximum CQ than the system with Sc2CF2 for atomic exchange.
Subsequently, Cui et al. theoretically studied the CQ of Sc2CF2 with different atomic vacancies (Figure 4c) [111]. They observe that, at positive bias, pristine Sc2CF2 has almost no CQ, while the introduction of vacancies increases it. The maximum CQ of Sc2CF2-VF at positive bias is 493 μF/cm2 (0.40 V). It is noteworthy that the CQ of all systems with vacancies at the Fermi level is larger than that of the original system. Rui et al. investigated the CQ of doped Sc2CF2 with 13 transition metal atoms (Figure 4d) [112]. Fe-doped Sc2CF2 shows a symmetry CQ-V curve, with a maximum CQ of 5407.6 μF/cm2 at 0 V. All the 4d transition metal atoms-doped Sc2CF2 structures show asymmetric CQ-V curves. The maximum CQ of the Mo-doped system at negative bias is 6917.88 μF/cm2 at −0.2 V. For the Nb-doped system, the maximum CQ is 2599.72 μF/cm2 at 0.52 V under positive bias (Figure 4e). For 5d transition metal dopants, the maximum CQ is 1833.15 μF/cm2 (0.48 V) in the Re-doped system. Cui et al. also studied the geometry, electronic properties, and CQ of Sc2CF2 with intrinsic defects [113]. They found that the CQ fluctuations are more pronounced for systems with defects. Among them, the maximum CQ of Sc2CF2-dF→C and Sc2CF2-dC→Sc monolayers are 924.69 μF/cm2 (−0.48 V) and 1024.03 μF/cm2 (0.56 V), respectively (Figure 4f). In their study, the charge (Q) of the Sc2CF2-dF→C monolayer was mainly stored in the negative potential (Figure 4g).

3.2.3. Hfn+1CnTx

Most functionalized MXenes have metallic properties, while Hf2CO2 has a moderate band gap and good thermal conductivity [114,115].
Liu et al. studied the CQ of Hf2CT2 (T = -O, -F, -S, -Cl, -OH, -Se) [116]. Their results show that the maximum CQ value of Hf2C is 549 μF/cm2 (0.56 V). At positive bias, the CQ of the other functionalized systems (T = -O, -S, -Se) are smaller than the original system. The CQ of Hf2CO2 is almost zero at positive bias. The maximum CQ of Hf2CSe2 at negative bias is 564 μF/cm2 (−0.56 V). The maximum CQ of Hf2C(OH)2 is 804 μF/cm2 at 0.44 V (Figure 5a). In experiments, mixed terminations are often randomly attached to the surface of MXenes during etching; thus, there are various configurations of surface coverage and mixed terminations. The surfaces of MXenes are often bound by mixed terminations, mainly -O, -F, and -OH [117]. Therefore, Liu et al. considered MXene groups with a mixed termination of -O, -F, and -OH [116]. The results show the symmetrical characteristics of CQ-V curves for all three groups, with the highest CQ at zero potential of 778.82, 552.17, and 177.97 μF/cm2, respectively (Figure 5b).
Liu et al. also computationally investigated the effect of N doping concentration on the electronic properties and CQ of Hf2CO2 [118]. Based on the calculation results, the CQ of Hf2CO2 systems with doping concentrations of 11%, 22%, 33%, and 44% are relatively low at negative bias. The maximum CQ of the pristine Hf2CO2 system (PH) is 84.06 μF/cm2. The maximum CQ values of PH-33% and PH-100% at positive bias are 423.62 μF/cm2 and 441.16 μF/cm2, respectively. The maximum CQ of PH-78% is 1208 μF/cm2. Thus, it indicates that the N-doping concentration also changes the CQ of Hf2CO2 monolayer. (Figure 5c) Subsequently, Liu et al. further investigated the maximum CQ of PH and the doped systems at different temperatures [118]. They noted that the maximum CQ decreased with increasing temperature for all systems except PH-22%. Among all systems, the maximum CQ is 1535.2 μF/cm2 at 233 K for the PH-78% system (Figure 5d). As with their previous studies, they also considered the case of mixed terminals. Their results show that the maximum CQ is increased for all other systems after N-doping. As the doping concentration increases, the CQ-V curve becomes more symmetrical (in the range of ±0.6 V).
More interestingly, Cui et al. explored the CQ of Hf2CO2 monolayers under bi-axial strain [119]. Strain is a common strategy to modulate the properties of materials, which can tune the electronic structure of the material, thus affecting many physical properties of the material. For example, it has been experimentally demonstrated that the introduction of tensile strain can lead to a transition from direct to indirect bandgap in the MoS monolayer, which expands the light absorption range and reduces the complexation of photogenerated carriers. Cui et al. have demonstrated that strain can significantly modulate the electronic structure of Hf2CO2 monolayers, which has a very important impact on the material properties [119]. The results show that the maximum CQ values of strain-free Hf2CO2 are 1.57 μF/cm2 (0 V) and 78.99 μF/cm2 (−0.6 V) under positive and negative bias, respectively. Under positive bias, the maximum CQ of the Hf2CO2 monolayer increases at all strains except 3%. Under negative bias, the maximum CQ of the Hf2CO2 monolayer increases at all strains except −6%, −4%, and −2% (Figure 5e). Li et al. explored the effect of adsorption of NH3 on pristine Hf2CO2 and varying its biaxial stress [120]. In the range of ±0.6 V, the CQ-V curve of Hf2CO2 under strain is asymmetric. The CQ at negative potentials is significantly higher than that at positive potentials, and the maximum CQ of Hf2CO2 under free strain is 38.75 μF/cm2 at −0.57 V. The maximum CQ increases gradually with increasing tensile strain and reaches a maximum of 244.27 μF/cm2 at +5% strain (Figure 5f,g).

3.2.4. Zrn+1CnTx

Zr2CO2 is an excellent functionalized MXene with many excellent properties. Due to its excellent photovoltaic properties and high hole mobility, it can be considered as a suitable photocatalyst [121,122].
Xu et al. investigated the electronic properties and CQ of pristine, doped, and single C vacant (VC) Zr2CO2 [123]. The doped atoms were chosen as Y = Si, Ge, Sn, N, B, S, F (Figure 6a,b). The doped atoms had a significant effect on CQ and Q. The maximum CQ of pristine Zr2CO2 was 407 μF/cm2 (−0.6 V) and 32.3 μF/cm2 at the Fermi level. The introduction of C vacancies increased the CQ at positive bias. The introduction of all the considered dopant atoms increased the maximum CQ, and B doping at negative bias increased the maximum CQ to 1993 μF/cm2. The maximum CQ of S-doped structure was 3293.7 μF/cm2 (0.4 V). At 0 V, a significant increase in CQ can be observed in the systems doped with VC, F, N and S atoms. Xu et al. also explored the maximum CQ of pristine Zr2CO2, Zr2CO2-VC, and doped Zr2CO2 at different temperatures [123]. Similar to the trend of CQ with temperature for N-doped Hf2CO2, they noted that the maximum CQ of each of the studied Zr2CO2 systems decreased gradually with increasing temperature. (Figure 6c)
Li et al. investigated the effect of C, O, and Zr vacancies (VC, VO, VZr) on the CQ of Zr2CO2 monolayers [124]. Their study shows that the introduction of atomic vacancies increases the maximum CQ in the range of ±0.6 V. In particular, the maximum CQ of PZ-VZr at positive bias is 586 μF/cm2 (−0.30 V). The maximum CQ of PZ-VC MXene under positive bias is 422 μF/cm2 (0.53 V), while the maximum CQ of PZ-VO MXene is 359 μF/cm2 (0.57 V) (Figure 6d,e). Yin et al. investigated the CQ of Zr2CO2 with an atomic exchange [125]. According to the plot of CQ-V, they point out that, in the range of ±0.6 V, the CQ of the pristine Zr2CO2 tends to be zero in positive bias, and in negative bias, the highest CQ is 76.8 μF/cm2. The highest CQ is 737.1 μF/cm2 in negative bias for the C-O1 exchanged system and 814.6 μF/cm2 in negative bias for the Zr-O2 exchanged system. The atomic exchange in Zr2CO2 greatly improves the top CQ at negative bias in ±0.6 V. The Zr-O2 exchanged system has the highest CQ of 425.3 μF/cm2 at 0 V (Figure 6f,g).

3.2.5. Nbn+1AnTx

The first synthesized 2D niobium carbide was the thicker Nb4C3. The Nb2C and Nb3C2 systems have an extremely high theoretical capacitance of Li atoms. Additionally, the surface termination has a considerable effect on the energy storage performance [126].
Xin et al. explored the CQ properties of different thicknesses of Nbn+1Cn (n = 2, 3, 4) [127]. They noted that the CQ values of all intrinsic niobium carbides were higher than that of functionalized ones in the positive bias voltage range. Except for Nb5C4, the other three niobium carbides have higher CQ values at positive bias than at negative bias. The functionalized Nbn+1Cn shows a higher CQ than the intrinsic system only at potentials below −0.4 V. In the positive bias range, functionalization causes a significant decrease in the CQ of the system. To quantitatively compare the CQ, they calculated the theoretical integrated CQ of the positive and negative electrodes from 0 to 0.83 V and from −0.62 to 0 V, respectively. Their results illustrate that the CQ of the intrinsic niobium carbide gradually decreases with the increasing number of layers in the positive potential region. For different thicknesses of functionalized Nbn+1Cn MXenes, the CQ is smaller than the intrinsic state. The CQ of intrinsic Nb2C is up to 1828.4 F/g at the positive electrode and 1091.1 F/g at the negative electrode (Figure 7a–c).
Transition metal nitrides, such as vanadium nitride, titanium nitride, and tungsten nitride, have been studied as electrode materials for EDLCs. It has been demonstrated that cobalt doping can increase the capacitance of niobium nitride. Bharti et al. calculated the CQ of Nb2N and Nb4N3 and investigated the effect of Co-doping on their CQ [128]. Their calculations show that the CQ value of Nb2N is remarkably high (1196.28 μF/cm2, −1 V) and the CQ of Nb4N3 is much lower than that of Nb2N. When they increased the number of layers of Nb2N and Nb4N3, they found that the CQ kept increasing with the more layer numbers. The CQ of both Nb4N3 (Nb4N3-2Co) and Nb2N (Nb2N-2Co) increase after Co-doping at the Fermi level, with the CQ of Nb2N-2Co reaching 1052.2 μF/cm2 (Figure 7d–f).

3.2.6. Mo2C and V2C

Two-dimensional transition metal carbides (TMCs) have high melting points and good electrical conductivity and chemical stability [129,130,131]. With its excellent electrochemical properties, Mo2C has been experimentally prepared as an electrode material for capacitors. In a experiment prepared by Lu et al., the capacitor with Mo2C as the electrode material had a high specific capacitance and excellent cycling stability, and its performance was significantly better than most carbide-based asymmetric supercapacitors [132]. As early as 2015, it was shown that V2CTx could be used as the positive electrode of sodium ion capacitors [133]. The results of Ai et al.’s study show that the specific capacitance of V2C in 1 M Na2SO4 is high (223.5 F/g) and the cycling stability is good (capacitance retention could be maintained at 94.7% after 5000 cycles) when the current density is 100 mA/g [134].
Bharti et al. discussed the CQ of Mo2C and V2C [135]. They highlighted that the CQ of intrinsic Mo2C and V2C reaches 3243.99 μF/cm2 and 3465.51 μF/cm2 at the Fermi level, respectively. In the positive potential range, the CQ decreases rapidly and drops to 0. Similar to Nb2N and Nb4N3, the CQ of Mo2C and V2C enhances with the increasing number of layers. The CQ values of both Mo2C and V2C at the Fermi energy decreased after O-functionalization. Similar to the pristine system, the O-functionalized V2C and Mo2C had higher CQ at negative bias (Figure 7g,h).

3.3. Transition Metal Dichalcogenides (TMDs)

Transition metal-based materials are considered to have higher energy density than other materials [136,137,138,139]. Among them, transition metal dichalcogenides (TMDs) are a class of graphene-like structures that have been commonly used in terms of electrode materials for supercapacitors in recent years. Although TMDs usually store energy in the form of intercalation with alkali metals, they exhibit quantum effects that are reflected in their capacitive behavior.
MoS2 is a typical representative of TMDs and exists in three main phases (2H, 3R, and 1T) with unique capacitive properties. Graphene-like MoS2 materials have special structures, fast ionic conductivity, and high specific capacitance [140,141]. In addition, the electron correlation between Mo layers in the sandwich structure facilitates the carrier transport [142,143,144]. It is an excellent electrode material for supercapacitors and has attracted more and more attention.
MoS2 was the subject of a comprehensive study by Xu et al. [145]. They first investigated the relationship between the CQ and the potential of MoS2 containing different dopants (where Ti, Au, Ag, Cu, and Al replace Mo atoms), single-vacancy VMo. Their results show that the CQ of pristine MoS2 is almost zero in the region near 0 V and increases at a higher voltage. For the CQ of Al-doped MoS2 and single-vacancy VMo, the local CQ maxima near 0 V are 157.7 μF/cm2 and 156.5 μF/cm2, respectively. Subsequently, they observed the effect of Al doping on the CQ in both pristine and single-vacant (VS) MoS2 monolayers with different Al concentrations of 1.3%, 2.1%, 3.7%, and 8.3%. The CQ value increases from 44.76 μF/cm2 (0.17 V) to 227.85 μF/cm2 (0.53 V) with the increase in doping concentration (Figure 8a–e). Secondly, they investigated the doped VS-MoS2, where Ti, Au, Ag, Cu, and Al replace the S atoms. The CQ increased in all systems except when doped with Ti, which is similar to that of S substitution in pristine MoS2 (Figure 8f). Finally, they calculated the local CQ maxima of 200.89 μF/cm2, 132.77 μF/cm2, and 254.29 μF/cm2 in B-, N-, and P-substituted S atoms in the B-, N-, and P-doped MoS2 monolayer (doping concentration kept at 3.7%), respectively. At the Fermi level, the B-doped system had higher CQ and a clear advantage in terms of positive potential (Figure 8g). Therefore, they continued to investigate the effect of B-doping concentration on CQ. The CQ value increased gradually with the increase of B-doping concentration.
It is worth noting that MoS2 should be considered as a van der Waals (vdW) 2D material. Biby et al. investigated the CQ of multilayered MoS2 with embedding and co-embedding in relation to vdW forces [146]. The CQ of the three-layered 1T phases was as high as 2080 F/g, and the CQ of 3R-MoS2 was slightly higher than that of 2H-MoS2 under negative bias. Subsequently, in their investigation on the effect of vdW forces on the CQ, they pointed out that the absence of vdW forces increased the strength of the density of electric states. Thus, for 1T-MoS2, the absence of vdW leads to a higher CQ at positive bias, and in the case of 2H and 3R-MoS2, the absence of vdW shifts the Fermi level, leading to a higher CQ in the negative potential window. This finding also directly emphasizes the importance of vdW forces in the accurate calculation of 2D material properties (Figure 8h,i). Finally, they investigated the intercalation of cations (Li+, Na+, K+ and H+) in the three phases of MoS2 and mixed Li+ and Na+ intercalation. They conclude that the CQ of the 1T phase is increased near the Fermi level with Na+ intercalation. Additionally, the 2H and 3R phases have a larger improvement, mainly in the positive bias voltage. There are three different intercalation modes in the case of mixed doping, and the CQ values of the co-intercalated system are higher than 2H Li-MoS2 and close to 2H Na-MoS2. The maximum CQ of LiNa-MoS2 (HASI) can reach 3163 F/g (Figure 8j) [146].
Irham et al. showed that the introduction of defects in h-FeS increased the CQ at positive bias up to 2280 F/g, but the CQ at negative bias decreased [147]. Cr-doped FeS has a maximum CQ of 3076 F/g (0.6 V) at positive bias. P-type dopants (Co or Ni) do not significantly increase the CQ in the positive voltage range. In their study of the integrated CQ, they found that the Cint also changes nonlinearly with the doping concentration at positive bias, where the Cint can reach 1013 F/g with the Cr-doping concentration of 6.24%. They attribute the emergence of these nonlinear changes to the appearance of electronic off-domain states that hinder the increase of Cint (Figure 8k).

4. Conclusions

In this paper, we reviewed studies investigating the CQ of 2D electrode materials through using theoretical calculations. In general, there are two solutions to enhance the performance of electrode materials for supercapacitors. One is to develop new electrode materials with higher performance, and the other is to modify the already found electrode materials, mainly by means of doping, adsorption, defects, atom exchange, etc. As far as the available studies are concerned, for graphene-like main group elements and compounds, all of the modification measures mentioned above can improve the CQ of the electrode material. In contrast, in MXene materials, not all modification measures are able to improve the material performance. For example, the CQ of functionalized Ti3C2 is not higher than that of the pristine structure. However, the introduction of functional groups during the preparation process is inevitable. Thus, it is necessary to investigate the CQ of the functionalized MXenes to select the preparation precursors with the least impact on energy storage performance. There are not many studies on CQ in transition metal-based supercapacitor electrode materials. Most TM-based materials are considered as pseudocapacitance supercapacitor materials. However, both pseudocapacitance and electric double-layer processes exist in such supercapacitors. So calculating the CQ of TM-based electrode materials can lead to a more accurate prediction of the theoretical capacitance and provide more possibilities for the development of supercapacitors.
Nowadays, theoretical calculation plays an important role in scientific research. On the one hand, theoretical calculations can help interpret the results of existing experimental phenomena, and on the other hand, they facilitate the prediction and development of new materials. Most of the current studies on the theoretical calculation of electrode materials for supercapacitors focus on predicting new materials. Although the high performance of electrode materials has been calculated theoretically, there are still great difficulties in the preparation process. In addition, the current research only pertains to the electrode part, and the performance of the whole capacitor has not been completely considered. Further research should pay more attention to the feasibility and stability of material modification. Secondly, various variables that may cause influence should be fully considered so that the theoretical prediction is closer to the real situation. And thirdly, the amount of attention given to the overall performance and process of supercapacitors should be improved. With the advancement of science and technology, theoretical calculations have become more accurate and fast. This also provides us with a more thorough explanation of supercapacitor performance and greater feasibility for designing functional materials with various properties. We expect that theoretical calculations will provide a greater contribution to the development of supercapacitors.

Funding

This work was funded by the Natural Science Foundation of China under No. 12204065; the Natural Science Foundation of Jilin Province (Grant Nos. YDZJ202201ZYTS576 and 20220508020RC), the 13th and 14th Five-year Planning Project of Jilin Provincial Education Department Foundation (Grant Nos. JJKH20200828KJ and JJKH20220836KJ), the Natural Science Foundation (Grant Nos. 001010, and 003179), and the PhD Starting Scientific Research Funding Project (No. 00300200360) of Changchun Normal University.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, Y.; Zhu, M.; Huang, Y.; Pei, Z.; Li, H.; Wang, Z.; Xue, Q.; Zhi, C. Multifunctional Energy Storage and Conversion Devices. Adv. Mater. 2016, 28, 8344–8364. [Google Scholar] [CrossRef]
  2. Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y. 25th anniversary article: MXenes: A new family of two-dimensional materials. Adv. Mater. 2014, 26, 992–1005. [Google Scholar] [CrossRef]
  3. Yu, Z.; Tetard, L.; Zhai, L.; Thomas, J. Supercapacitor electrode materials: Nanostructures from 0 to 3 dimensions. Energy Environ. Sci. 2015, 8, 702–730. [Google Scholar] [CrossRef] [Green Version]
  4. Vangari, M.; Pryor, T.; Jiang, L. Supercapacitors: Review of Materials and Fabrication Methods. J. Energ. Eng. 2013, 139, 72–79. [Google Scholar] [CrossRef]
  5. Raza, W.; Ali, F.; Raza, N.; Luo, Y.; Kim, K.-H.; Yang, J.; Kumar, S.; Mehmood, A.; Kwon, E.E. Recent advancements in supercapacitor technology. Nano Energy 2018, 52, 441–473. [Google Scholar] [CrossRef]
  6. Libich, J.; Máca, J.; Vondrák, J.; Čech, O.; Sedlaříková, M. Supercapacitors: Properties and applications. J. Energy Storage 2018, 17, 224–227. [Google Scholar] [CrossRef]
  7. Yang, W.; Ni, M.; Ren, X.; Tian, Y.; Li, N.; Su, Y.; Zhang, X. Graphene in Supercapacitor Applications. Curr. Opin. Colloid. Interface Sci. 2015, 20, 416–428. [Google Scholar] [CrossRef]
  8. Najib, S.; Erdem, E. Current progress achieved in novel materials for supercapacitor electrodes: Mini review. Nanoscale Adv. 2019, 1, 2817–2827. [Google Scholar] [CrossRef] [Green Version]
  9. Yin, X.; Zhang, J.; Yang, L.; Xiao, W.; Zhou, L.; Tang, Y.; Yang, W. Carbon electrodes with ionophobic characteristics in organic electrolyte for high-performance electric double-layer capacitors. Sci. China Chem. 2021, 65, 383–390. [Google Scholar] [CrossRef]
  10. Schütter, C.; Pohlmann, S.; Balducci, A. Industrial Requirements of Materials for Electrical Double Layer Capacitors: Impact on Current and Future Applications. Adv. Energy Mater. 2019, 9, 00334. [Google Scholar] [CrossRef]
  11. Zhan, C.; Neal, J.; Wu, J.; Jiang, D.-E. Quantum Effects on the Capacitance of Graphene-Based Electrodes. J. Phys. Chem. C 2015, 119, 22297–22303. [Google Scholar] [CrossRef]
  12. Paek, E.; Pak, A.J.; Kweon, K.E.; Hwang, G.S. On the Origin of the Enhanced Supercapacitor Performance of Nitrogen-Doped Graphene. J. Phys. Chem. C 2013, 117, 5610–5616. [Google Scholar] [CrossRef]
  13. Pak, A.J.; Paek, E.; Hwang, G.S. Relative contributions of quantum and double layer capacitance to the supercapacitor performance of carbon nanotubes in an ionic liquid. Phys. Chem. Chem. Phys. 2013, 15, 19741–19747. [Google Scholar] [CrossRef]
  14. Yang, G.M.; Zhang, H.Z.; Fan, X.F.; Zheng, W.T. Density Functional Theory Calculations for the Quantum Capacitance Performance of Graphene-Based Electrode Material. J. Phys. Chem. C 2015, 119, 6464–6470. [Google Scholar] [CrossRef]
  15. Paek, E.; Pak, A.J.; Hwang, G.S. Curvature Effects on the Interfacial Capacitance of Carbon Nanotubes in an Ionic Liquid. J. Phys. Chem. C 2013, 117, 23539–23546. [Google Scholar] [CrossRef]
  16. Paek, E.; Pak, A.J.; Hwang, G.S. A Computational Study of the Interfacial Structure and Capacitance of Graphene in [BMIM][PF6] Ionic Liquid. J. Electrochem. Soc. 2012, 160, A1–A10. [Google Scholar] [CrossRef] [Green Version]
  17. Yu, G.L.; Jalil, R.; Belle, B.; Mayorov, A.S.; Blake, P.; Schedin, F.; Morozov, S.V.; Ponomarenko, L.A.; Chiappini, F.; Wiedmann, S.; et al. Interaction phenomena in graphene seen through quantum capacitance. Proc. Natl. Acad. Sci. USA 2013, 110, 3282–3286. [Google Scholar] [CrossRef] [Green Version]
  18. Luryi, S. Quantum capacitance devices. Appl. Phys. Lett. 1988, 52, 501–503. [Google Scholar] [CrossRef]
  19. John, D.L.; Castro, L.C.; Pulfrey, D.L. Quantum capacitance in nanoscale device modeling. J. Appl. Phys 2004, 96, 5180–5184. [Google Scholar] [CrossRef] [Green Version]
  20. Zhang, L.; Huang, K.; Wen, P.; Wang, J.; Yang, G.; Liu, D.; Lin, Z.; Lian, C.; Liu, H.; Zheng, S.; et al. Tailoring the defects of two-dimensional borocarbonitride nanomesh for high energy density micro-supercapacitor. Energy Stor. Mater. 2021, 42, 430–437. [Google Scholar] [CrossRef]
  21. Lu, X.; Wang, C.; Favier, F.; Pinna, N. Electrospun Nanomaterials for Supercapacitor Electrodes: Designed Architectures and Electrochemical Performance. Adv. Energy Mater. 2017, 7, 1601301. [Google Scholar] [CrossRef]
  22. Poudel, M.B.; Kim, H.J. Confinement of Zn-Mg-Al-layered double hydroxide and α-Fe2O3 nanorods on hollow porous carbon nanofibers: A free-standing electrode for solid-state symmetric supercapacitors. Chem. Eng J. 2022, 429, 132345. [Google Scholar] [CrossRef]
  23. He, P.; Chen, S. Vertically Oriented Graphene Nanosheets for Electrochemical Energy Storage. ChemElectroChem 2021, 8, 783–797. [Google Scholar] [CrossRef]
  24. Bokhari, S.W.; Siddique, A.H.; Sherrell, P.C.; Yue, X.; Karumbaiah, K.M.; Wei, S.; Ellis, A.V.; Gao, W. Advances in graphene-based supercapacitor electrodes. Energy Rep. 2020, 6, 2768–2784. [Google Scholar] [CrossRef]
  25. Frackowiak, E. Carbon materials for supercapacitor application. Phys. Chem. Chem. Phys. 2007, 9, 1774–1785. [Google Scholar] [CrossRef]
  26. Zhang, L.L.; Zhao, X.S. Carbon-based materials as supercapacitor electrodes. Chem. Soc. Rev. 2009, 38, 252031. [Google Scholar] [CrossRef]
  27. Borenstein, A.; Hanna, O.; Attias, R.; Luski, S.; Brousse, T.; Aurbach, D. Carbon-based composite materials for supercapacitor electrodes: A review. J. Mater. Chem. A 2017, 5, 12653–12672. [Google Scholar] [CrossRef]
  28. Hirunsit, P.; Liangruksa, M.; Khanchaitit, P. Electronic structures and quantum capacitance of monolayer and multilayer graphenes influenced by Al, B, N and P doping, and monovacancy: Theoretical study. Carbon 2016, 108, 7–20. [Google Scholar] [CrossRef]
  29. Kopp, T.; Mannhart, J. Calculation of the capacitances of conductors: Perspectives for the optimization of electronic devices. J. Appl. Phys. 2009, 106, 064504. [Google Scholar] [CrossRef] [Green Version]
  30. Lin, J.; Yu, T.; Han, F.; Yang, G. Computational predictions of two-dimensional anode materials of metal-ion batteries. Wiley. Interdiscip. Rev. Comput. Mol. Sci. 2020, 10, 147310. [Google Scholar] [CrossRef] [Green Version]
  31. Malko, D.; Neiss, C.; Vines, F.; Gorling, A. Competition for graphene: Graphynes with direction-dependent Dirac cones. Phys. Rev. Lett. 2012, 108, 086804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Tiwari, S.K.; Sahoo, S.; Wang, N.; Huczko, A. Graphene research and their outputs: Status and prospect. J. Sci. Adv. Mater. Devices 2020, 5, 10–29. [Google Scholar] [CrossRef]
  33. Xia, J.; Chen, F.; Li, J.; Tao, N. Measurement of the quantum capacitance of graphene. Nat. Nanotechnol. 2009, 4, 505–509. [Google Scholar] [CrossRef]
  34. Jeong, H.M.; Lee, J.W.; Shin, W.H.; Choi, Y.J.; Shin, H.J.; Kang, J.K.; Choi, J.W. Nitrogen-doped graphene for high-performance ultracapacitors and the importance of nitrogen-doped sites at basal planes. Nano Lett. 2011, 11, 2472–2477. [Google Scholar] [CrossRef] [PubMed]
  35. Kim, W.; Joo, J.B.; Kim, N.; Oh, S.; Kim, P.; Yi, J. Preparation of nitrogen-doped mesoporous carbon nanopipes for the electrochemical double layer capacitor. Carbon 2009, 47, 1407–1411. [Google Scholar] [CrossRef]
  36. Li, W.; Chen, D.; Li, Z.; Shi, Y.; Wan, Y.; Wang, G.; Jiang, Z.; Zhao, D. Nitrogen-containing carbon spheres with very large uniform mesopores: The superior electrode materials for EDLC in organic electrolyte. Carbon 2007, 45, 1757–1763. [Google Scholar] [CrossRef]
  37. Seredych, M.; Hulicova-Jurcakova, D.; Lu, G.Q.; Bandosz, T.J. Surface functional groups of carbons and the effects of their chemical character, density and accessibility to ions on electrochemical performance. Carbon 2008, 46, 1475–1488. [Google Scholar] [CrossRef]
  38. Lota, G.; Frackowiak, E. Pseudocapacitance Effects for Enhancement of Capacitor Performance. Fuel Cells 2010, 10, 848–855. [Google Scholar] [CrossRef] [Green Version]
  39. Qiu, Y.; Zhang, X.; Yang, S. High performance supercapacitors based on highly conductive nitrogen-doped graphene sheets. Phys. Chem. Chem. Phys. 2011, 13, 125548. [Google Scholar] [CrossRef]
  40. Paraknowitsch, J.P.; Zhang, J.; Su, D.; Thomas, A.; Antonietti, M. Ionic liquids as precursors for nitrogen-doped graphitic carbon. Adv. Mater. 2010, 22, 87–92. [Google Scholar] [CrossRef]
  41. Hulicova-Jurcakova, D.; Kodama, M.; Shiraishi, S.; Hatori, H.; Zhu, Z.H.; Lu, G.Q. Nitrogen-Enriched Nonporous Carbon Electrodes with Extraordinary Supercapacitance. Adv. Funct. Mater. 2009, 19, 1800–1809. [Google Scholar] [CrossRef]
  42. Zhang, L.L.; Zhao, X.; Ji, H.; Stoller, M.D.; Lai, L.; Murali, S.; McDonnell, S.; Cleveger, B.; Wallace, R.M.; Ruoff, R.S. Nitrogen doping of graphene and its effect on quantum capacitance, and a new insight on the enhanced capacitance of N-doped carbon. Energy Environ. Sci. 2012, 5, 9618–9625. [Google Scholar] [CrossRef]
  43. Mousavi-Khoshdel, S.M.; Targholi, E. Exploring the effect of functionalization of graphene on the quantum capacitance by first principle study. Carbon 2015, 89, 148–160. [Google Scholar] [CrossRef]
  44. Chen, L.; Li, X.; Ma, C.; Wang, M.; Zhou, J. Interaction and Quantum Capacitance of Nitrogen/Sulfur Co-Doped Graphene: A Theoretical Calculation. J. Phys. Chem. C 2017, 121, 18344–18350. [Google Scholar] [CrossRef]
  45. Hu, R.; Shang, J. Quantum capacitance of transition metal and nitrogen co-doped graphenes as supercapacitors electrodes: A DFT study. Appl. Surf. Sci. 2019, 496, 143659. [Google Scholar] [CrossRef]
  46. Wang, M.; Chen, L.; Zhou, J.; Xu, L.; Li, X.; Li, L.; Li, X. First-principles calculation of quantum capacitance of metals doped graphenes and nitrogen/metals co-doped graphenes: Designing strategies for supercapacitor electrodes. J. Mater. Sci. 2018, 54, 483–492. [Google Scholar] [CrossRef]
  47. Song, C.; Wang, J.; Meng, Z.; Hu, F.; Jian, X. Density Functional Theory Calculations of the Quantum Capacitance of Graphene Oxide as a Supercapacitor Electrode. Chemphyschem 2018, 19, 1579–1583. [Google Scholar] [CrossRef]
  48. Sruthi, T.; Kartick, T. Route to achieving enhanced quantum capacitance in functionalized graphene based supercapacitor electrodes. J. Phys. Condens. Matter. 2019, 31, 475502. [Google Scholar] [CrossRef] [Green Version]
  49. Xu, Q.; Yang, G.; Fan, X.; Zheng, W. Improving the Quantum Capacitance of Graphene-Based Supercapacitors by the Doping and Co-Doping: First-Principles Calculations. ACS Omega 2019, 4, 13209–13217. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, B.; Peng, Z.; Song, L.; Wu, X.; Fu, X. Computational screening toward quantum capacitance of transition-metals and vacancy doped/co-doped graphene as electrode of supercapacitors. Electrochim. Acta 2021, 385, 138432. [Google Scholar] [CrossRef]
  51. Sruthi, T.; Tarafder, K. Enhancement of quantum capacitance by chemical modification of graphene supercapacitor electrodes: A study by first principles. Bull. Mater. Sci. 2019, 42, 257–263. [Google Scholar] [CrossRef] [Green Version]
  52. Cui, G.-Y.; Yi, Z.-L.; Su, F.-Y.; Chen, C.-M.; Han, P.-D. A DFT study of the effect of stacking on the quantum capacitance of bilayer graphene materials. New Carbon Mater. 2021, 36, 1062–1070. [Google Scholar] [CrossRef]
  53. Zhou, Q.; Ju, W.; Yong, Y.; Zhang, Q.; Liu, Y.; Li, J. Effect of the N/P/S and transition-metal co-doping on the quantum capacitance of supercapacitor electrodes based on mono- and multilayer graphene. Carbon 2020, 170, 368–379. [Google Scholar] [CrossRef]
  54. Zeng, W.; Zhang, Y.; Liu, X.; Qi, L.; Kang, W.; Fang, L.; Zhou, M. B/N-doped graphdiyne as superior supercapacitor electrode with record high quantum capacitance. Appl. Surf. Sci. 2020, 523, 146468. [Google Scholar] [CrossRef]
  55. Nawaz, S.; Tahir, M. Quantum capacitance in monolayers of silicene and related buckled materials. Phys. E Low Dimens. Syst. Nanostruct. 2016, 76, 169–172. [Google Scholar] [CrossRef] [Green Version]
  56. Yang, G.M.; Xu, Q.; Fan, X.; Zheng, W.T. Quantum Capacitance of Silicene-Based Electrodes from First-Principles Calculations. J. Phys. Chem. C 2018, 122, 1903–1912. [Google Scholar] [CrossRef]
  57. Momeni, M.J.; Mousavi-Khoshdel, M.; Leisegang, T. Exploring the performance of pristine and defective silicene and silicene-like XSi3 (X = Al, B, C, N, P) sheets as supercapacitor electrodes: A density functional theory calculation of quantum capacitance. Phys. E Low Dimens. Syst. Nanostruct. 2020, 124, 114290. [Google Scholar] [CrossRef]
  58. Xu, Q.; Yang, G.M.; Fan, X.; Zheng, W.T. Adsorption of metal atoms on silicene: Stability and quantum capacitance of silicene-based electrode materials. Phys. Chem. Chem. Phys. 2019, 21, 4276–4285. [Google Scholar] [CrossRef]
  59. Liu, C.-C.; Jiang, H.; Yao, Y. Low-energy effective Hamiltonian involving spin-orbit coupling in silicene and two-dimensional germanium and tin. Phys. Rev. B 2011, 84, 195430. [Google Scholar] [CrossRef] [Green Version]
  60. Kaloni, T.P. Tuning the Structural, Electronic, and Magnetic Properties of Germanene by the Adsorption of 3d Transition Metal Atoms. J. Phys. Chem. C 2014, 118, 25200–25208. [Google Scholar] [CrossRef] [Green Version]
  61. Xu, Q.; Si, X.; She, W.H.; Yang, G.M.; Fan, X.; Zheng, W.T. First-Principles Calculation of Optimizing the Performance of Germanene-Based Supercapacitors by Vacancies and Metal Atoms. J. Phys. Chem. C 2020, 124, 12346–12358. [Google Scholar] [CrossRef]
  62. Zhang, L.; Bampoulis, P.; Rudenko, A.N.; Yao, Q.; van Houselt, A.; Poelsema, B.; Katsnelson, M.I.; Zandvliet, H.J. Structural and Electronic Properties of Germanene on MoS2. Phys. Rev. Lett. 2016, 116, 256804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Liu, N.; Bo, G.; Liu, Y.; Xu, X.; Du, Y.; Dou, S.X. Recent Progress on Germanene and Functionalized Germanene: Preparation, Characterizations, Applications, and Challenges. Small 2019, 15, 1805147. [Google Scholar] [CrossRef] [Green Version]
  64. Zhuang, J.; Gao, N.; Li, Z.; Xu, X.; Wang, J.; Zhao, J.; Dou, S.X.; Du, Y. Cooperative Electron-Phonon Coupling and Buckled Structure in Germanene on Au(111). ACS Nano 2017, 11, 3553–3559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Qin, Z.; Pan, J.; Lu, S.; Shao, Y.; Wang, Y.; Du, S.; Gao, H.J.; Cao, G. Direct Evidence of Dirac Signature in Bilayer Germanene Islands on Cu(111). Adv. Mater. 2017, 29, 1606046. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, W.; Uhrberg, R.I.G. Investigation of the atomic and electronic structures of highly ordered two-dimensional germanium on Au(111). Phys. Rev. Mater. 2017, 1, 074002. [Google Scholar] [CrossRef] [Green Version]
  67. Zhuang, J.; Liu, C.; Zhou, Z.; Casillas, G.; Feng, H.; Xu, X.; Wang, J.; Hao, W.; Wang, X.; Dou, S.X.; et al. Dirac Signature in Germanene on Semiconducting Substrate. Adv. Sci. (Weinh) 2018, 5, 180–207. [Google Scholar] [CrossRef]
  68. Zhou, Q.; Ju, W.; Yong, Y.; Liu, Y.; Li, J. Quantum capacitance of supercapacitor electrodes based on germanene influenced by vacancy and co-doping: A first-principles study. Comput. Mater. Sci. 2021, 188, 110131. [Google Scholar] [CrossRef]
  69. Si, X.; She, W.; Xu, Q.; Yang, G.; Li, Z.; Wang, S.; Luan, J. First-Principles Density Functional Theory Study of Modified Germanene-Based Electrode Materials. Materials 2021, 15, 103. [Google Scholar] [CrossRef]
  70. Abbasi, A.; Sardroodi, J.J. The adsorption of sulfur trioxide and ozone molecules on stanene nanosheets investigated by DFT: Applications to gas sensor devices. Phys. E Low Dimens. Syst. Nanostruct. 2019, 108, 382–390. [Google Scholar] [CrossRef]
  71. Zhang, A.; Yang, H.; Liu, Q.; Li, W.; Wang, Y. DFT insights into the adsorption properties of toxic gas molecules on pure and transition metal embedded stanene monolayers: Towards gas sensor devices. Synth. Met. 2020, 266, 116441. [Google Scholar] [CrossRef]
  72. Nagarajan, V.; Chandiramouli, R. Adsorption behavior of NH3 and NO2 molecules on stanene and stanane nanosheets—A density functional theory study. Chem. Phys. Lett. 2018, 695, 162–169. [Google Scholar] [CrossRef]
  73. Sahoo, S.K.; Wei, K.H. A Perspective on Recent Advances in 2D Stanene Nanosheets. Adv. Mater. Interfaces 2019, 6, 00752. [Google Scholar] [CrossRef] [Green Version]
  74. Shaidu, Y.; Akin-Ojo, O. First principles predictions of superconductivity in doped stanene. Comput. Mater. Sci. 2016, 118, 11–15. [Google Scholar] [CrossRef] [Green Version]
  75. Evazzade, I.; Roknabadi, M.R.; Morshedloo, T.; Modarresi, M.; Mogulkoc, Y.; Nemati Far, H. Semimetal behavior of bilayer stanene. Phys. E Low Dimens. Syst. Nanostruct. 2017, 89, 155–159. [Google Scholar] [CrossRef] [Green Version]
  76. Zhou, Q.; Wang, L.; Ju, W.; Yong, Y.; Wu, S.; Cai, S.; Li, P. Quantum capacitance of vacancy-defected and co-doped stanene for supercapacitor electrodes: A theoretical study. Electrochim. Acta 2022, 433, 141261. [Google Scholar] [CrossRef]
  77. Zhou, Q.; Wang, L.; Ju, W.; Miao, H.; Ye, S.; Liu, Y. Influence of the co-doping and line-doping on the quantum capacitance of stanene for supercapacitor electrodes. Chem. Phys. Lett. 2022, 808, 140123. [Google Scholar] [CrossRef]
  78. Zhong, H.-X.; Gao, S.; Shi, J.-J.; Yang, L. Quasiparticle band gaps, excitonic effects, and anisotropic optical properties of the monolayer distorted1Tdiamond-chain structuresReS2 andReSe2. Phys. Rev. B 2015, 92, 115438. [Google Scholar] [CrossRef] [Green Version]
  79. Ramasubramaniam, A. Large excitonic effects in monolayers of molybdenum and tungsten dichalcogenides. Phys. Rev. B 2012, 86, 115409. [Google Scholar] [CrossRef] [Green Version]
  80. Cheng, T.; Lang, H.; Li, Z.; Liu, Z.; Liu, Z. Anisotropic carrier mobility in two-dimensional materials with tilted Dirac cones: Theory and application. Phys. Chem. Chem. Phys. 2017, 19, 23942–23950. [Google Scholar] [CrossRef] [Green Version]
  81. Kolavada, H.; Singh, S.; Lukačević, I.; Gajjar, P.N.; Gupta, S.K. Quantum capacitance of multi-layered δ-6 borophene: A DFT study. Electrochim. Acta 2023, 439, 141589. [Google Scholar] [CrossRef]
  82. Chen, X.; Wang, L.; Wu, Y.; Gao, H.; Wu, Y.; Qin, G.; Wu, Z.; Han, Y.; Xu, S.; Han, T.; et al. Probing the electronic states and impurity effects in black phosphorus vertical heterostructures. 2D Mater. 2016, 3, 015012. [Google Scholar] [CrossRef] [Green Version]
  83. Chen, X.; Xu, G.; Ren, X.; Li, Z.; Qi, X.; Huang, K.; Zhang, H.; Huang, Z.; Zhong, J. A black/red phosphorus hybrid as an electrode material for high-performance Li-ion batteries and supercapacitors. J. Mater. Chem. A 2017, 5, 6581–6588. [Google Scholar] [CrossRef]
  84. Batmunkh, M.; Bat-Erdene, M.; Shapter, J.G. Phosphorene and Phosphorene-Based Materials—Prospects for Future Applications. Adv. Mater. 2016, 28, 8586–8617. [Google Scholar] [CrossRef]
  85. Vishnoi, P.; Rajesh, S.; Manjunatha, S.; Bandyopadhyay, A.; Barua, M.; Pati, S.K.; Rao, C.N.R. Doping Phosphorene with Holes and Electrons through Molecular Charge Transfer. Chemphyschem 2017, 18, 2985–2989. [Google Scholar] [CrossRef]
  86. Huang, Z.; Wu, Y.; Qi, X.; He, C.; Ren, X.; Zhong, J. Electronic and Magnetic Properties of Monolayer and Bilayer Phosphorene Doped with Transition-Metal Atoms. Phys. Status Solidi. B 2018, 255, 1700370. [Google Scholar] [CrossRef]
  87. Zu, L.; Gao, X.; Lian, H.; Li, C.; Liang, Q.; Liang, Y.; Cui, X.; Liu, Y.; Wang, X.; Cui, X. Electrochemical prepared phosphorene as a cathode for supercapacitors. J. Alloys Compd. 2019, 770, 26–34. [Google Scholar] [CrossRef]
  88. Ramesh, A.; Mir, A. Influence of heteroatom doping on the quantum capacitance of phosphorene supercapacitors. J. Energy Storage 2022, 56, 106013. [Google Scholar] [CrossRef]
  89. Kessler, F.K.; Zheng, Y.; Schwarz, D.; Merschjann, C.; Schnick, W.; Wang, X.; Bojdys, M.J. Functional carbon nitride materials—Design strategies for electrochemical devices. Nat. Rev. Mater. JCR 2017, 2, 17030. [Google Scholar] [CrossRef]
  90. Zhao, Z.; Sun, Y.; Dong, F. Graphitic carbon nitride based nanocomposites: A review. Nanoscale 2015, 7, 15–37. [Google Scholar] [CrossRef]
  91. Liu, A.Y.; Wentzcovitch, R.M. Stability of carbon nitride solids. Phys. Rev. B Condens. Matter. 1994, 50, 10362–10365. [Google Scholar] [CrossRef] [PubMed]
  92. Chen, X.; Xu, W.; Jin, J.; Wang, P.; Song, B.; He, P. Effect of B and O doping on the electronic structure and quantum capacitance of carbon nitride monolayers using first-principles calculations. J. Appl. Phys. 2021, 129, 174301. [Google Scholar] [CrossRef]
  93. Majdi, A.; Kadhim Wadday, A.; Sabri Abbas, Z.; Kadhim, M.M.; Mahdi Rheima, A.; Barzan, M.; Haitham Alattia, L.; Hachim, S.K.; Abdul Hadi, M. Quantum capacitance of iron metal doped boron carbide monolayer-based for supercapacitors electrodes: A DFT study. Inorg. Chem. Commun. 2023, 150, 110480. [Google Scholar] [CrossRef]
  94. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [Green Version]
  95. Zhan, X.; Si, C.; Zhou, J.; Sun, Z. MXene and MXene-based composites: Synthesis, properties and environment-related applications. Nanoscale Horiz. 2020, 5, 235–258. [Google Scholar] [CrossRef]
  96. Shuck, C.E.; Sarycheva, A.; Anayee, M.; Levitt, A.; Zhu, Y.; Uzun, S.; Balitskiy, V.; Zahorodna, V.; Gogotsi, O.; Gogotsi, Y. Scalable Synthesis of Ti3C2Tx MXene. Adv. Eng. Mater. 2020, 22, 01241. [Google Scholar] [CrossRef]
  97. Li, X.; Wang, C.; Cao, Y.; Wang, G. Functional MXene Materials: Progress of Their Applications. Chem. Asian J. 2018, 13, 2742–2757. [Google Scholar] [CrossRef]
  98. Hu, T.; Yang, J.; Wang, X. Carbon vacancies in Ti2CT2 MXenes: Defects or a new opportunity? Phys. Chem. Chem. Phys. 2017, 19, 31773–31780. [Google Scholar] [CrossRef]
  99. Cheng, Y.-W.; Dai, J.-H.; Zhang, Y.-M.; Song, Y. Transition metal modification and carbon vacancy promoted Cr2CO2(MXenes): A new opportunity for a highly active catalyst for the hydrogen evolution reaction. J. Mater. Chem. A 2018, 6, 20956–20965. [Google Scholar] [CrossRef]
  100. Er, D.; Li, J.; Naguib, M.; Gogotsi, Y.; Shenoy, V.B. Ti3C2 MXene as a high capacity electrode material for metal (Li, Na, K, Ca) lon batteries. ACS Appl. Mater. Interfaces 2014, 6, 11173–11179. [Google Scholar] [CrossRef]
  101. Lipatov, A.; Alhabeb, M.; Lukatskaya, M.R.; Boson, A.; Gogotsi, Y.; Sinitskii, A. Effect of Synthesis on Quality, Electronic Properties and Environmental Stability of Individual Monolayer Ti3C2MXene Flakes. Adv. Electron. Mater. 2016, 2, 1600255. [Google Scholar] [CrossRef] [Green Version]
  102. Si, X.; Xu, Q.; Lin, J.; Yang, G. Quantum capacitance modulation of MXenes by metal atoms adsorption. Appl. Surf. Sci. 2023, 618, 156586. [Google Scholar] [CrossRef]
  103. Bafekry, A.; Nguyen, C.V.; Stampfl, C.; Akgenc, B.; Ghergherehchi, M. Oxygen Vacancies in the Single Layer of Ti2CO2 MXene: Effects of Gating Voltage, Mechanical Strain, and Atomic Impurities. Phys. Status Solidi B 2020, 257, 2000343. [Google Scholar] [CrossRef]
  104. Su, X.; Guo, R.-G.; Xu, S.; Wang, S.-J.; Li, X.-H.; Cui, H.-L. Influence of O-vacancy concentration on the structural, electronic properties and quantum capacitance of monolayer Ti2CO2: A first-principles study. Vacuum 2022, 196, 110740. [Google Scholar] [CrossRef]
  105. Li, X.-H.; Liu, M.-Z.; Tan, X.-M.; Su, X.; Guo, R.-G.; Zhang, R.-Z.; Cui, H.-L. Influence of C-vacancy-line defect on electronic and optical properties and quantum capacitance of Ti2CO2 MXene: A first-principles study. J. Phys. Chem. Solids 2023, 176, 111254. [Google Scholar] [CrossRef]
  106. Li, X.-H.; Li, S.-S.; Cui, X.-H.; Zhang, R.-Z.; Cui, H.-L. First-principle study of electronic properties and quantum capacitance of lithium adsorption on pristine and vacancy-defected O-functionalized Ti2C MXene. Appl. Surf. Sci. 2021, 563, 150264. [Google Scholar] [CrossRef]
  107. Kumar, S.; Schwingenschlögl, U. Thermoelectric performance of functionalized Sc2C MXenes. Phys. Rev. B 2016, 94, 035405. [Google Scholar] [CrossRef] [Green Version]
  108. Ma, Z.; Hu, Z.; Zhao, X.; Tang, Q.; Wu, D.; Zhou, Z.; Zhang, L. Tunable Band Structures of Heterostructured Bilayers with Transition-Metal Dichalcogenide and MXene Monolayer. J. Phys. Chem. C 2014, 118, 5593–5599. [Google Scholar] [CrossRef]
  109. Cui, X.-H.; Li, X.-H.; Jin, X.-J.; Zhang, R.-Z.; Cui, H.-L.; Yan, H.-T. First-principles study on the effect of atomic swap on the electronic properties and quantum capacitance of Sc2CF2 monolayer. Vacuum 2022, 204, 111371. [Google Scholar] [CrossRef]
  110. Cui, X.-H.; Li, X.-H.; Zhang, R.-Z.; Cui, H.-L.; Yan, H.-T. Theoretical insight into the electronic, optical, and photocatalytic properties and quantum capacitance of Sc2CT2 (T = F, P, Cl, Se, Br, O, Si, S, OH) MXenes. Vacuum 2023, 207, 111615. [Google Scholar] [CrossRef]
  111. Cui, X.-H.; Li, X.-H.; Li, S.-S.; Cui, H.-L. The effect of vacancy defect on quantum capacitance, electronic and magnetic properties of Sc2CF2 monolayer. Int. J. Quantum Chem. 2021, 121, e26666. [Google Scholar] [CrossRef]
  112. Zhang, R.-Z.; Cui, X.-H.; Li, S.-S.; Li, X.-H.; Cui, H.-L. DFT computation of quantum capacitance of transition-metals and vacancy doped Sc2CF2 MXene for supercapacitor applications. J. Mol. Liq. 2022, 345, 118263. [Google Scholar] [CrossRef]
  113. Cui, X.-H.; Li, X.-H.; Zhang, R.-Z.; Cui, H.-L.; Yan, H.-T. Computational study of intrinsic defects on the electronic properties and quantum capacitance of Sc2CF2 monolayer. Surf. Interfaces 2022, 33, 102193. [Google Scholar] [CrossRef]
  114. Zha, X.H.; Huang, Q.; He, J.; He, H.; Zhai, J.; Francisco, J.S.; Du, S. The thermal and electrical properties of the promising semiconductor MXene Hf2CO2. Sci. Rep. 2016, 6, 27971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wang, Y.; Ma, S.; Wang, L.; Jiao, Z. A novel highly selective and sensitive NH3 gas sensor based on monolayer Hf2CO2. Appl. Surf. Sci. 2019, 492, 116–124. [Google Scholar] [CrossRef]
  116. Liu, M.-Z.; Li, X.-H.; Cui, X.-H.; Yan, H.-T.; Zhang, R.-Z.; Cui, H.-L. The influence of different functional groups on quantum capacitance, electronic and optical properties of Hf2C MXene. Appl. Surf. Sci. 2022, 605, 154830. [Google Scholar] [CrossRef]
  117. Lei, J.-C.; Zhang, X.; Zhou, Z. Recent advances in MXene: Preparation, properties, and applications. Front. Phys. 2015, 10, 276–286. [Google Scholar] [CrossRef]
  118. Liu, M.-Z.; Li, X.-H.; Yan, H.-T.; Zhang, R.-Z.; Cui, H.-L. Influence of N-doped concentration on the electronic properties and quantum capacitance of Hf2CO2 MXene. Vacuum 2023, 210, 111826. [Google Scholar] [CrossRef]
  119. Cui, X.-H.; Li, X.-H.; Jin, X.-J.; Yan, H.-T.; Zhang, R.-Z.; Cui, H.-L. Biaxial strain tunable quantum capacitance and photocatalytic properties of Hf2CO2 monolayer. Appl. Surf. Sci. 2023, 616, 156579. [Google Scholar] [CrossRef]
  120. Li, S.-S.; Li, X.-H.; Cui, X.-H.; Zhang, R.-Z.; Cui, H.-L. Effect of the biaxial strain on the electronic structure, quantum capacitance of NH3 adsorption on pristine Hf2CO2 MXene using first-principles calculations. Appl. Surf. Sci. 2022, 575, 151659. [Google Scholar] [CrossRef]
  121. Li, X.-H.; Cui, X.-H.; Xing, C.-H.; Cui, H.-L.; Zhang, R.-Z. Strain-tunable electronic and optical properties of Zr2CO2 MXene and MoSe2 van der Waals heterojunction: A first principles calculation. Appl. Surf. Sci. 2021, 548, 149249. [Google Scholar] [CrossRef]
  122. Mostafaei, A.; Faizabadi, E.; Semiromi, E.H. Theoretical studies and tuning the electronic and optical properties of Zr2CO2 monolayer using biaxial strain effect: Modified Becke–Johnson calculation. Physica. E Low Dimens. Syst. Nanostruct. 2019, 114, 113559. [Google Scholar] [CrossRef]
  123. Xu, S.; Wang, S.J.; Sun, W.Q.; Li, X.H.; Cui, H.L. Density functional theory study of the electronic properties and quantum capacitance of pure and doped Zr2CO2 as electrode of supercapacitors. Int. J. Quantum. Chem. 2021, 122, e26844. [Google Scholar] [CrossRef]
  124. Li, X.-H.; Jia, H.-Y.; Li, S.-S.; Cui, H.-L. Effect of atomic vacancy on the electronic and optical properties, quantum capacitance of Zr2CO2-based electrodes. Vacuum 2021, 190, 110284. [Google Scholar] [CrossRef]
  125. Yin, S.H.; Cui, X.H.; Li, X.H.; Wang, S.J.; Zhang, R.Z.; Cui, H.L.; Yan, H.T. Theoretical study on the electronic properties and quantum capacitance of Zr2CO2 MXene with atomic swap. Int. J. Quantum. Chem. 2023, 125, e27135. [Google Scholar] [CrossRef]
  126. Bai, L.; Yin, H.; Wu, L.; Zhang, X. First-principle study of the Nbn+1CnT2 systems as electrode materials for supercapacitors. Comput. Mater. Sci. 2018, 143, 225–231. [Google Scholar] [CrossRef]
  127. Xin, Y.; Yu, Y.-X. Possibility of bare and functionalized niobium carbide MXenes for electrode materials of supercapacitors and field emitters. Mater. Des. 2017, 130, 512–520. [Google Scholar] [CrossRef]
  128. Bharti; Ahmed, G.; Kumar, Y.; Bocchetta, P.; Sharma, S. Determination of Quantum Capacitance of Niobium Nitrides Nb2N and Nb4N3 for Supercapacitor Applications. J. Compos. Sci. 2021, 5, 85. [Google Scholar] [CrossRef]
  129. Hussain, S.; Rabani, I.; Vikraman, D.; Feroze, A.; Ali, M.; Seo, Y.-S.; Song, W.; An, K.-S.; Kim, H.-S.; Chun, S.-H.; et al. MoS2@X2C (X = Mo or W) hybrids for enhanced supercapacitor and hydrogen evolution performances. Chem. Eng. J. 2021, 421, 127843. [Google Scholar] [CrossRef]
  130. Hussain, S.; Faizan, M.; Vikraman, D.; Rabani, I.; Ali, B.; Kim, H.-S.; Jung, J.; Nam, K.-W. Eutectoid WXC embedded WS2 nanosheets as a hybrid composite anode for lithium-ion batteries. Ceram. Int. 2021, 47, 18646–18655. [Google Scholar] [CrossRef]
  131. Poudel, M.B.; Logeshwaran, N.; Kim, A.R.; Karthikeyan, S.C.; Vijayapradeep, S.; Yoo, D.J. Integrated core-shell assembly of Ni3S2 nanowires and CoMoP nanosheets as highly efficient bifunctional electrocatalysts for overall water splitting. J. Alloys Compd. 2023, 960, 170678. [Google Scholar] [CrossRef]
  132. Lu, C.; Zhang, Y.-S.; Hu, Y.-X.; Li, J.; Zhang, B.-M.; Liu, M.-C. Ultrafine Mo2C nanoparticles supported on three-dimensional hierarchical porous carbon architecture toward electrochemical energy storage applications. J. Energy Storage 2021, 33, 101855. [Google Scholar] [CrossRef]
  133. Dall’Agnese, Y.; Taberna, P.L.; Gogotsi, Y.; Simon, P. Two-Dimensional Vanadium Carbide (MXene) as Positive Electrode for Sodium-Ion Capacitors. J. Phys. Chem. Lett. 2015, 6, 2305–2309. [Google Scholar] [CrossRef] [Green Version]
  134. Ai, W.; Zhang, C.; Xia, L.; Miao, H.; Yuan, J. Synthesis of High-Quality Two-Dimensional V2C MXene for Supercapacitor Application. Energies 2022, 15, 3696. [Google Scholar] [CrossRef]
  135. Bharti; Kumar, Y.; Gupta, M.; Sharma, S. Study of pristine and functionalized V2C and Mo2C MXenes as novel electrode material for supercapacitors. J. Mol. Graph. Model. 2023, 118, 108366. [Google Scholar] [CrossRef]
  136. Karade, S.S.; Dwivedi, P.; Majumder, S.; Pandit, B.; Sankapal, B.R. First report on a FeS-based 2 V operating flexible solid-state symmetric supercapacitor device. Sustain. Energy Fuels 2017, 1, 1366–1375. [Google Scholar] [CrossRef]
  137. Yu, F.; Chang, Z.; Yuan, X.; Wang, F.; Zhu, Y.; Fu, L.; Chen, Y.; Wang, H.; Wu, Y.; Li, W. Ultrathin NiCo2S4@graphene with a core–shell structure as a high performance positive electrode for hybrid supercapacitors. J. Mater. Chem. A 2018, 6, 5856–5861. [Google Scholar] [CrossRef]
  138. Rui, X.; Tan, H.; Yan, Q. Nanostructured metal sulfides for energy storage. Nanoscale 2014, 6, 9889–9924. [Google Scholar] [CrossRef] [PubMed]
  139. Septianto, R.D.; Liu, L.; Iskandar, F.; Matsushita, N.; Iwasa, Y.; Bisri, S.Z. On-demand tuning of charge accumulation and carrier mobility in quantum dot solids for electron transport and energy storage devices. NPG Asia Mater. 2020, 12, 2–14. [Google Scholar] [CrossRef]
  140. Lin, Y.C.; Dumcenco, D.O.; Huang, Y.S.; Suenaga, K. Atomic mechanism of the semiconducting-to-metallic phase transition in single-layered MoS2. Nat. Nanotechnol. 2014, 9, 391–396. [Google Scholar] [CrossRef]
  141. Ali, B.A.; Metwalli, O.I.; Khalil, A.S.G.; Allam, N.K. Unveiling the Effect of the Structure of Carbon Material on the Charge Storage Mechanism in MoS2-Based Supercapacitors. ACS Omega 2018, 3, 16301–16308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Yin, X.; Wang, Q.; Cao, L.; Tang, C.S.; Luo, X.; Zheng, Y.; Wong, L.M.; Wang, S.J.; Quek, S.Y.; Zhang, W.; et al. Tunable inverted gap in monolayer quasi-metallic MoS2 induced by strong charge-lattice coupling. Nat. Commun. 2017, 8, 006402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Singh, N.; Schwingenschlogl, U. A Route to Permanent Valley Polarization in Monolayer MoS2. Adv. Mater. 2017, 29, 1600970. [Google Scholar] [CrossRef] [PubMed]
  144. Nie, Z.; Long, R.; Teguh, J.S.; Huang, C.-C.; Hewak, D.W.; Yeow, E.K.L.; Shen, Z.; Prezhdo, O.V.; Loh, Z.-H. Ultrafast Electron and Hole Relaxation Pathways in Few-Layer MoS2. J. Phys. Chem. C 2015, 119, 20698–20708. [Google Scholar] [CrossRef]
  145. Xu, Q.; Yang, G.M.; Zheng, W.T. DFT calculation for stability and quantum capacitance of MoS2 monolayer-based electrode materials. Mater. Today Commun. 2020, 22, 100772. [Google Scholar] [CrossRef]
  146. Biby, A.H.; Ali, B.A.; Allam, N.K. Interplay of quantum capacitance with Van der Waals forces, intercalation, co-intercalation, and the number of MoS2 layers. Mater. Today Energy 2021, 20, 100677. [Google Scholar] [CrossRef]
  147. Irham, M.A.; Muttaqien, F.; Bisri, S.Z.; Iskandar, F. Enhancing quantum capacitance of iron sulfide supercapacitor through defect-engineering: A first-principles calculation. Electrochim. Acta 2023, 449, 142235. [Google Scholar] [CrossRef]
Figure 1. (a) CQ versus potential curves for single- and double-layer graphene. (b) Dependence of CQ on N-dopant concentration as a function of graphene potential (Vch). (c) The structures of six defective graphene models. (d) CQ maps of the graphene model with different N-doping concentrations. (e) Schematic structures of monovalent, divalent, and trivalent functional groups. (f) CQ-V curves of pristine, cavity, pyridinic-N-doped and pyrrolic-N-doped graphene. (g) Structure of 3N-doped and single-vacant graphene (N3V). (h,i) CQ versus potential plot for X1 and X3V monolithic graphene structures. (j) CQ versus potential plot for CoNx (x = 1, 2, 3, 4) co-doped graphene. (k) CQ versus potential plot for graphene doping within the transition metal plane. (l) The maximum CQ in the potential range of −1.5 V~1.5 V for graphene with different oxygen-containing group concentrations at the negative and positive electrodes. (m) Optimized structures of pristine, epoxy graphene and graphene oxide containing hydroxyl groups. (n) Change trend chart of the maximum value of CQ for the B(N, P, S)-doped graphene with different doping models (model-a, model-b, model-c, and model-d) and the N/S, N/P-co-doped the supercell 4 × 4 graphene with different models (model-e, model-f, and model-g). (o) Structures of vacancy-defected (VG) and Stone–Wales defected graphenes (SWG). (p) CQ of VG and SWG at different concentrations. (q) The stability of TM@G and TM@VG models. (r) PDOS and CQ versus potential curves for N, O co-doped graphene. (s) CQ versus potential for monolayered and bilayered graphene with different defects. (t) Configuration of defective bilayer graphene containing D2_11 type (555–777) point defects.
Figure 1. (a) CQ versus potential curves for single- and double-layer graphene. (b) Dependence of CQ on N-dopant concentration as a function of graphene potential (Vch). (c) The structures of six defective graphene models. (d) CQ maps of the graphene model with different N-doping concentrations. (e) Schematic structures of monovalent, divalent, and trivalent functional groups. (f) CQ-V curves of pristine, cavity, pyridinic-N-doped and pyrrolic-N-doped graphene. (g) Structure of 3N-doped and single-vacant graphene (N3V). (h,i) CQ versus potential plot for X1 and X3V monolithic graphene structures. (j) CQ versus potential plot for CoNx (x = 1, 2, 3, 4) co-doped graphene. (k) CQ versus potential plot for graphene doping within the transition metal plane. (l) The maximum CQ in the potential range of −1.5 V~1.5 V for graphene with different oxygen-containing group concentrations at the negative and positive electrodes. (m) Optimized structures of pristine, epoxy graphene and graphene oxide containing hydroxyl groups. (n) Change trend chart of the maximum value of CQ for the B(N, P, S)-doped graphene with different doping models (model-a, model-b, model-c, and model-d) and the N/S, N/P-co-doped the supercell 4 × 4 graphene with different models (model-e, model-f, and model-g). (o) Structures of vacancy-defected (VG) and Stone–Wales defected graphenes (SWG). (p) CQ of VG and SWG at different concentrations. (q) The stability of TM@G and TM@VG models. (r) PDOS and CQ versus potential curves for N, O co-doped graphene. (s) CQ versus potential for monolayered and bilayered graphene with different defects. (t) Configuration of defective bilayer graphene containing D2_11 type (555–777) point defects.
Nanomaterials 13 01932 g001
Figure 2. (a) CQ versus potential plot for single vacant and pyridine-N-doped silicene. (b) Schematic diagram of the atomic structure of 2D AlSi3. (c) CQ versus potential curves for pristine and X-doped silicene (X = Al, B, C, P, N). (d) CQ versus potential plot for single vacant silicene adsorbed with different Ti concentrations. (e) Adsorption energy of metal atoms on defective germanene, silicene, and graphene. (f) The plot of CQ versus potential for defective germanene co-doped with Ti, Cr, Mn, Co, and Al. (g) CQ versus potential curves for N/S co-doped single vacant germanene. (h) Structure of B and transition metal atoms co-doped with defective stanene. (i) CQ versus potential plot for BTMVSn. (j,k) Structures of S and Ti atoms co-doped and line-doped stanene. (l) TDOS diagram of SSTiSn. (m) CQ diagrams of multilayered boronene. (n) Structure of Ni-doped phosphorene. (o,p) DOS and CQ diagrams of B-doped tg-C3N4. (q) CQ versus potential plots for iron-doped BC3 monolayer.
Figure 2. (a) CQ versus potential plot for single vacant and pyridine-N-doped silicene. (b) Schematic diagram of the atomic structure of 2D AlSi3. (c) CQ versus potential curves for pristine and X-doped silicene (X = Al, B, C, P, N). (d) CQ versus potential plot for single vacant silicene adsorbed with different Ti concentrations. (e) Adsorption energy of metal atoms on defective germanene, silicene, and graphene. (f) The plot of CQ versus potential for defective germanene co-doped with Ti, Cr, Mn, Co, and Al. (g) CQ versus potential curves for N/S co-doped single vacant germanene. (h) Structure of B and transition metal atoms co-doped with defective stanene. (i) CQ versus potential plot for BTMVSn. (j,k) Structures of S and Ti atoms co-doped and line-doped stanene. (l) TDOS diagram of SSTiSn. (m) CQ diagrams of multilayered boronene. (n) Structure of Ni-doped phosphorene. (o,p) DOS and CQ diagrams of B-doped tg-C3N4. (q) CQ versus potential plots for iron-doped BC3 monolayer.
Nanomaterials 13 01932 g002
Figure 3. (a,b) CQ diagrams of functionalized Ti3C2 and Ti2C. (c,d) CQ diagrams of metal atoms-adsorbed Ti3C2 and Ti2C. (e) CQ diagram of metal atoms-adsorbed Ti3C2F2. (f) Structure of Ti3C2F2 adsorbed by metal atoms. (g) PDOS diagram of Ti2CO2 with an oxygen vacancy concentration of 5.56%. (h) CQ diagram of Ti2CO2 with different oxygen vacancy concentrations. (i) Schematic diagram of the vacant defected Ti2CO2 structure by removing atoms with atomic numbers (2), (3, 4), (1, 5, 6) and (3, 5, 7), denoted as CVL1, CVL2, CVL3, and CVL4, respectively. (j) The optimized structures of PT-LT monolayer. The red, light blue, brown, and light green balls denote the oxygen, titanium, carbon, and lithium atoms, respectively. (k) CQ diagrams of Li atoms adsorbed on Ti2CO2 (PT) and C-vacant Ti2CO2 (VT) monolayers.
Figure 3. (a,b) CQ diagrams of functionalized Ti3C2 and Ti2C. (c,d) CQ diagrams of metal atoms-adsorbed Ti3C2 and Ti2C. (e) CQ diagram of metal atoms-adsorbed Ti3C2F2. (f) Structure of Ti3C2F2 adsorbed by metal atoms. (g) PDOS diagram of Ti2CO2 with an oxygen vacancy concentration of 5.56%. (h) CQ diagram of Ti2CO2 with different oxygen vacancy concentrations. (i) Schematic diagram of the vacant defected Ti2CO2 structure by removing atoms with atomic numbers (2), (3, 4), (1, 5, 6) and (3, 5, 7), denoted as CVL1, CVL2, CVL3, and CVL4, respectively. (j) The optimized structures of PT-LT monolayer. The red, light blue, brown, and light green balls denote the oxygen, titanium, carbon, and lithium atoms, respectively. (k) CQ diagrams of Li atoms adsorbed on Ti2CO2 (PT) and C-vacant Ti2CO2 (VT) monolayers.
Nanomaterials 13 01932 g003
Figure 4. (a) TDOS and CQ of Sc2CF2-SC↔F, Sc2CF2-SC↔Sc, and Sc2CF2-SF↔Sc atomic exchange monolayers. (b) Three possible configurations of Sc2CT2, namely Sc-top, C-top, and hybrid configurations. (c) CQ of three vacancy-defective systems formed by removing a C, F, or Sc atom from Sc2CF2. (d) Schematic diagram of Sc2CF2 replacing one Sc atom for doping. (e) CQ versus potential plot for Sc2CF2 doped with 4d TM atoms (including Y, Zr, Nb, and Mo). (f) The plot of CQ versus potential for atomic exchanged Sc2CF2. (g) Charge versus potential for atomic exchanged Sc2CF2.
Figure 4. (a) TDOS and CQ of Sc2CF2-SC↔F, Sc2CF2-SC↔Sc, and Sc2CF2-SF↔Sc atomic exchange monolayers. (b) Three possible configurations of Sc2CT2, namely Sc-top, C-top, and hybrid configurations. (c) CQ of three vacancy-defective systems formed by removing a C, F, or Sc atom from Sc2CF2. (d) Schematic diagram of Sc2CF2 replacing one Sc atom for doping. (e) CQ versus potential plot for Sc2CF2 doped with 4d TM atoms (including Y, Zr, Nb, and Mo). (f) The plot of CQ versus potential for atomic exchanged Sc2CF2. (g) Charge versus potential for atomic exchanged Sc2CF2.
Nanomaterials 13 01932 g004
Figure 5. (a) The plot of CQ versus potential for Hf2CT2 (T = -O, -F, -S, -Cl, -OH, -Se). (b) The plots of CQ and surface charge versus potential for the mixed terminal of -O, -F, and -OH groups in Hf2C. (c) CQ versus potential plot of Hf2CO2 with different N-doping concentrations. (d) The maximum CQ of Hf2CO2 with different N-doping concentrations at 233 K, 300 K, and 353 K (e) Top and side views of the Hf2CO2 monolayer, where the applied biaxial strain is along the a and b directions. (f) Schematic diagram of the site of the NH3 molecule adsorption. (g) CQ and surface charge versus potential plots for NH3-adsorbed Hf2CO2 under different strain magnitudes.
Figure 5. (a) The plot of CQ versus potential for Hf2CT2 (T = -O, -F, -S, -Cl, -OH, -Se). (b) The plots of CQ and surface charge versus potential for the mixed terminal of -O, -F, and -OH groups in Hf2C. (c) CQ versus potential plot of Hf2CO2 with different N-doping concentrations. (d) The maximum CQ of Hf2CO2 with different N-doping concentrations at 233 K, 300 K, and 353 K (e) Top and side views of the Hf2CO2 monolayer, where the applied biaxial strain is along the a and b directions. (f) Schematic diagram of the site of the NH3 molecule adsorption. (g) CQ and surface charge versus potential plots for NH3-adsorbed Hf2CO2 under different strain magnitudes.
Nanomaterials 13 01932 g005
Figure 6. (a,b) CQ and surface charge versus potential curves for the doped or vacant Zr2CO2. (c) The maximum CQ of pristine Zr2CO2, Zr2CO2-VC, and doped Zr2CO2 at different temperatures. (d) DOS of a C-vacant Zr2CO2 monolayer (PZ-VC). (e) shows the CQ versus potential of PZ-VC, PZ-VO, and PZ-VZr. (f) Schematic diagram of Zr2CO2 with different atomic exchange modifications. (g) CQ versus potential plot for atomic exchanged Zr2CO2.
Figure 6. (a,b) CQ and surface charge versus potential curves for the doped or vacant Zr2CO2. (c) The maximum CQ of pristine Zr2CO2, Zr2CO2-VC, and doped Zr2CO2 at different temperatures. (d) DOS of a C-vacant Zr2CO2 monolayer (PZ-VC). (e) shows the CQ versus potential of PZ-VC, PZ-VO, and PZ-VZr. (f) Schematic diagram of Zr2CO2 with different atomic exchange modifications. (g) CQ versus potential plot for atomic exchanged Zr2CO2.
Nanomaterials 13 01932 g006
Figure 7. (a) Top view of Nb2C structure. (b) Side view of Nb2C(OCH3)2. (c) CQ versus potential curves of Nb5C4 adsorbed with different functional groups. (d) Variation in CQ for pristine unpolarized niobium nitride structures under bias voltage. (e,f) CQ of multilayered Nb2N and Nb4N3. (g,h) CQ versus potential curves for functionalized Mo2C and V2C.
Figure 7. (a) Top view of Nb2C structure. (b) Side view of Nb2C(OCH3)2. (c) CQ versus potential curves of Nb5C4 adsorbed with different functional groups. (d) Variation in CQ for pristine unpolarized niobium nitride structures under bias voltage. (e,f) CQ of multilayered Nb2N and Nb4N3. (g,h) CQ versus potential curves for functionalized Mo2C and V2C.
Nanomaterials 13 01932 g007
Figure 8. Structures of (a) single-vacant (VS) MoS2 monolayer; (b) double-vacant (V2S) MoS2 monolayer; (c) B, N, P, Ti, Au, Ag, Cu, and Al-doped MoS2 monolayer; (d) Ti, Au, Ag, Cu, and Al-doped VS-MoS2 monolayer. (e) The plot of CQ versus potential for VS-MoS2 monolayers with Al substituted S atoms at different doping concentrations. (f) CQ versus potential plot for VS-MoS2 monolayers with Ti-, Au-, Ag-, Cu-, and Al-substituted S atoms. (g) The plot of CQ versus potential for pristine MoS2 monolayers with B-, N-, and P-substituted S atoms. (h) Plots of CQ versus potential with and without vdW for three-layered 1T-phase MoS2. (i) Plots of differential charge density for 1T, 2H, and 3R-phases MoS2. (j) CQ versus potential curves for Li+ and Na+ co-doped 2H phase MoS2. (k) Cint of h-FeS for anode-like and cathode-like supercapacitors at different doping concentrations.
Figure 8. Structures of (a) single-vacant (VS) MoS2 monolayer; (b) double-vacant (V2S) MoS2 monolayer; (c) B, N, P, Ti, Au, Ag, Cu, and Al-doped MoS2 monolayer; (d) Ti, Au, Ag, Cu, and Al-doped VS-MoS2 monolayer. (e) The plot of CQ versus potential for VS-MoS2 monolayers with Al substituted S atoms at different doping concentrations. (f) CQ versus potential plot for VS-MoS2 monolayers with Ti-, Au-, Ag-, Cu-, and Al-substituted S atoms. (g) The plot of CQ versus potential for pristine MoS2 monolayers with B-, N-, and P-substituted S atoms. (h) Plots of CQ versus potential with and without vdW for three-layered 1T-phase MoS2. (i) Plots of differential charge density for 1T, 2H, and 3R-phases MoS2. (j) CQ versus potential curves for Li+ and Na+ co-doped 2H phase MoS2. (k) Cint of h-FeS for anode-like and cathode-like supercapacitors at different doping concentrations.
Nanomaterials 13 01932 g008
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

Lin, J.; Yuan, Y.; Wang, M.; Yang, X.; Yang, G. Theoretical Studies on the Quantum Capacitance of Two-Dimensional Electrode Materials for Supercapacitors. Nanomaterials 2023, 13, 1932. https://doi.org/10.3390/nano13131932

AMA Style

Lin J, Yuan Y, Wang M, Yang X, Yang G. Theoretical Studies on the Quantum Capacitance of Two-Dimensional Electrode Materials for Supercapacitors. Nanomaterials. 2023; 13(13):1932. https://doi.org/10.3390/nano13131932

Chicago/Turabian Style

Lin, Jianyan, Yuan Yuan, Min Wang, Xinlin Yang, and Guangmin Yang. 2023. "Theoretical Studies on the Quantum Capacitance of Two-Dimensional Electrode Materials for Supercapacitors" Nanomaterials 13, no. 13: 1932. https://doi.org/10.3390/nano13131932

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