Next Article in Journal
Osteoimmune Properties of Mesoporous Bioactive Nanospheres: A Study on T Helper Lymphocytes
Previous Article in Journal
Progress in Carbon Nanostructures: From Synthesis to Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Heteroatom-Doped Molybdenum Disulfide Nanomaterials for Gas Sensors, Alkali Metal-Ion Batteries and Supercapacitors

by
Lyubov G. Bulusheva
*,
Galina I. Semushkina
and
Anastasiya D. Fedorenko
Nikolaev Institute of Inorganic Chemistry SB RAS, 3 Acad. Lavrentiev Ave., 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(15), 2182; https://doi.org/10.3390/nano13152182
Submission received: 15 June 2023 / Revised: 11 July 2023 / Accepted: 23 July 2023 / Published: 26 July 2023
(This article belongs to the Section 2D and Carbon Nanomaterials)

Abstract

:
Molybdenum disulfide (MoS2) is the second two-dimensional material after graphene that received a lot of attention from the research community. Strong S–Mo–S bonds make the sandwich-like layer mechanically and chemically stable, while the abundance of precursors and several developed synthesis methods allow obtaining various MoS2 architectures, including those in combinations with a carbon component. Doping of MoS2 with heteroatom substituents can occur by replacing Mo and S with other cations and anions. This creates active sites on the basal plane, which is important for the adsorption of reactive species. Adsorption is a key step in the gas detection and electrochemical energy storage processes discussed in this review. The literature data were analyzed in the light of the influence of a substitutional heteroatom on the interaction of MoS2 with gas molecules and electrolyte ions. Theory predicts that the binding energy of molecules to a MoS2 surface increases in the presence of heteroatoms, and experiments showed that such surfaces are more sensitive to certain gases. The best electrochemical performance of MoS2-based nanomaterials is usually achieved by including foreign metals. Heteroatoms improve the electrical conductivity of MoS2, which is a semiconductor in a thermodynamically stable hexagonal form, increase the distance between layers, and cause lattice deformation and electronic density redistribution. An analysis of literature data showed that co-doping with various elements is most attractive for improving the performance of MoS2 in sensor and electrochemical applications. This is the first comprehensive review on the influence of foreign elements inserted into MoS2 lattice on the performance of a nanomaterial in chemiresistive gas sensors, lithium-, sodium-, and potassium-ion batteries, and supercapacitors. The collected data can serve as a guide to determine which elements and combinations of elements can be used to obtain a MoS2-based nanomaterial with the properties required for a particular application.

1. Introduction

Molybdenum disulfide (MoS2) is a layered compound, where the layers are three atoms thick due to the sandwich-like S–Mo–S structure [1]. Within the layer, molybdenum forms strong bonds with six sulfur atoms, while sulfur atoms from neighboring layers interact through weak van der Waals (vdW) forces. Depending on the coordination of sulfur atoms to the molybdenum atom and the vertical stacking of layers, MoS2 can form several crystal phases [2]. The hexagonal 2H phase is the most thermodynamically stable and constitutes about 80% of the natural mineral molybdenite [3]. Owing to the easy sliding of layers under applied shear, MoS2 finds industrial use as a dry lubricant [4]. Another practical application of MoS2 is oil hydrodesulfurization [5]. The catalytic properties of MoS2 are associated with chemically active edge states. The large interlayer distance of 0.62 nm makes MoS2 attractive for intercalation reactions [6]. Such reactions proceed most easily with lithium-containing compounds. The electrochemical intercalation of lithium ions into MoS2 was studied at the end of the last century [7,8] in the search for a suitable cathode material for rechargeable lithium-ion batteries [9]. However, the lithiation voltage of 1.1–2.0 V for MoS2 [10] is too low to provide a high energy density of a full cell.
Interestingly, the delamination of 2H-MoS2 crystals to monolayers by the peeling technique was carried out in 1966 [11], forty years before graphene was obtained from graphite in a similar way. After it was found that graphene has unique properties as compare to the bulk parent material [12], interest in layered materials increased significantly [13], and MoS2 has become the second most intensively studied material from this family [14]. In contrast to graphene, MoS2 is a semiconductor with a varied band gap depending on the number of adjacent layers [15]. This property makes MoS2 a candidate for use in electronic and optoelectronic circuits, memory elements, etc. [16,17]. The sulfur-rich surface is capable of adsorbing various substances; therefore, MoS2 is promising for water purification [18,19] and detection of chemicals in liquid and gaseous environments [20,21]. Sulfur has a strong affinity for heavy metals, and depending on the nature of the metal, adsorption occurs through complexation, electrostatic interaction, or redox reaction mechanisms, which were discussed in the critical review [19]. Detailed data on the development of MoS2 nanostructures for real-time detection of biomarkers, drugs, and food ingredients were given in the review [22]. Electrochemical MoS2-based sensing elements can be used in portable and wearable devices for monitoring blood glucose levels, on-site detection of inorganic ions, and undesirable additives in water and food [22], as well as for recording human motion, since they show a high response over a wide operation pressure range [23]. The appropriate band gap and significant change in conductivity due to charge transfer from/to adsorbed molecules make MoS2 useful for gas detection. Reviews in this field of application emphasize the need to find suitable modifiers for MoS2 to achieve ultra-high selectivity in gas sensing [24,25]. Energy applications of MoS2 can be classified as energy storage and energy generation [26] The first is provided by batteries and electrochemical capacitors (or supercapacitors), the most studied process of generating energy using MoS2 as a catalyst is the hydrogen evolution reaction (HER) [27,28].
Most applications of MoS2 require tuning the band gap and/or creating the necessary active sites. These characteristics can be purposefully changed by replacing Mo and S with other chemical elements. It was shown that doping with heteroatoms improves the tribological properties of MoS2 [29], enhances the efficiency of water purification [30], electrocatalytic water splitting [31], and detection of NO2 gas [32]. Ni and Co located at the edges of MoS2 clusters were revealed to promote the desulfurization of organic compounds [33]. Since 2018, more than two thousand papers in the field of MoS2-based nanomaterials have been published annually, and so, the description and critical analysis of the current state of research is constantly required. Strategies for inserting transition metals into MoS2 nanomaterials were described in [34]. The review considered the change in the electrical conductivity of MoS2 upon substitutional doping and catalytic properties for HER, hydrodesulfurization reactions, and reduction of carbon dioxide to methanol. An overview of developments on substitutional doping of layered transition metal dichalcogenides, including MoS2, is presented in [35]. It was concluded that many foreign elements could be incorporated in the lattice of these materials, which opens up wide opportunities for the engineering of electronic structure. Changes in the structure, electrical, optical, and magnetic properties of MoS2 crystals due to various single atom dopants were discussed in [36]. However, there is still no comprehensive review considering the effect of heteroatoms on the properties of MoS2 nanomaterials for gas sensors and electrochemical energy storage devices, the efficiency of which strongly depends on the conductivity of the nanomaterial and its interaction with the adsorbate.
This gap is filled by the present review paper devoted to the performance of heteroatom-doped MoS2 nanomaterials in chemiresistive gas sensors, alkali metal-ion batteries, and supercapacitors. Modern research cannot be imagined without theoretical calculations evaluating the thermodynamic possibility of the formation of a particular structure, changes in the band gap of a semiconductor during doping, the binding energy with an adsorbate, and the charge transfer accompanying this process. Therefore, we first analyzed theoretical studies on the change in the electronic structure of MoS2, as a result of substitutional doping in Section 2.1. Section 2.2 describes the experimental methods, which allow identifying the incorporation of heteroatoms into the MoS2 lattice. Section 3 is devoted to chemiresistive MoS2 sensors, it describes and analyzes the data obtained for undoped MoS2, compares them with data for heteroatom-doped MoS2, and presents the adsorption energies and charge transfer values in models calculated using density functional theory (DFT). Section 4 provides electrochemical characteristics of heteroatom-doped MoS2 nanomaterials in lithium-ion batteries (LIBs), sodium-ion batteries (SIBs), potassium-ion batteries (PIBs), and supercapacitors. The main results of the review and the challenges are given in Section 5.

2. Foreign Elements in MoS2

Sulfur belongs to the chalcogen family and in compounds with metals has an oxidation state of −2. Since there are two sulfur atoms per molybdenum atom in MoS2, molybdenum oxidation state is +4. The incorporation of foreign atoms into the MoS2 structure can occur along the cationic and anionic sublattices. The elements considered in theoretical studies as possible substituents for Mo and S are shown in Figure 1 in shaded cells of the periodic table of chemical elements.
Substitution by an element belonging to the same group is isovalent and basically causes a shift in the energy levels of the resulting compound. The introduction of elements with a different number of valence electrons than those of Mo and S atoms can additionally locally distort the lattice and affect the character of the chemical bonding. In this section, we will first consider these changes revealed by DFT calculations and then briefly describe the experimental data on the ability of MoS2 to include foreign elements and the methods for identifying doping.

2.1. Theoretical Aspects

DFT calculations are used to estimate the formation energy of alloys and heteroatoms, the thermodynamically preferable configurations of foreign atoms, and the changes in the electronic structure of the compounds that accompany the replacement. Most of the calculations are carried out for monolayers.
The formation energy of an alloy Mo1−xMxS2 or MoS2−xYx-containing metal (M) or nonmetal (Y) heteroatom is defined as:
Ef = Ealloy − (1 − x)EMoS2 − xEMS2(MoY2),
where the terms are the total energies of the optimized alloy structure and the optimized hexagonal constituents MoS2 and metal sulfide MS2 or molybdenum compound MoY2 per formula unit.
The formation energy (formation enthalpy) of a substitutional dopant can be calculated as the difference between the total energies of a monolayer with a heteroatom (EH) and an ideal MoS2 monolayer, taking into account the energy (or chemical potential) of an individual native atom (EN) and an individual inserting atom (EA):
Ef = EH − EMoS2 + EN − EA
Some authors consider the filling of an atomic vacancy present in the MoS2 layer. In this case, the formation (or binding) energy is calculated relative to the layer with the vacancy:
Ef = EV + EA − EH,
where EV is the total energy of optimized supercell with an atomic vacancy.
For a uniform distribution of a foreign atom in the crystal, the central Mo or S atom in the supercell is replaced by another element and the portion of the dopant depends on the size of the supercell. Since this size and the calculation method used affect the energy of the system, we will only consider the trends obtained in a particular work.

2.1.1. Molybdenum Replacement

In the hexagonal layer, the molybdenum atom has six molybdenum neighbors (top projection in Figure 1) located at a distance of 0.316 nm, and DFT calculations show these atoms weakly interact with each other [37]. The top of the valence band of MoS2 is mainly formed by the Mo 4d states, while the low conduction band consists of the Mo 4d and S 2p orbitals [38].
Metals from s- and p-blocks. Alkali metals (Li, Na, and K), alkaline earth metals (Mg, Ca, and Sr), and boron group metals (Al, Ca, and In) were considered in [39] to assess the stability of doped MoS2 monolayers. The impurity concentration was 6.25%. Calculations showed that the foreign metal–sulfur bond is longer than the Mo–S bond. The greatest energy gain, determined by Equation (3), was obtained when the molybdenum vacancy was filled with Al, and among the considered metals, the filling with K was energetically less favorable. All introduced metals created impurity states near the top of the valence band. The study of p-block metals Ga, Sn, and Sb revealed that they prefer to be located on the edges rather in the MoS2 basal plane [40]. Sulfur-rich conditions were shown to be favorable for replacing Mo with Sb [41]. The heteroatom acts as the predominant p-type dopant and its incorporation in an amount of 4% into the MoS2 monolayer leads to the appearance of five states inside the band gap. These states are the result of re-hybridization of the levels induced by a single Mo vacancy and the Sb 5s orbital.
Metals from d-block. Cr and W belong to the same metal group of the periodic table of the periodic table of the elements (Figure 1) and the substitution of Mo by these metals is isovalent. The introduction of isolated Cr atoms into a bulk hexagonal MoS2 crystal with a concentration of 6.25 and 12.5% only slightly changes the band gap (by 0.05–0.07 eV), retaining the semiconducting nature of the compound [42]. The lattice mismatch of hexagonal MoS2 and CrS2 monolayers is ~4%, and the formation of Mo1−xCrxS2 alloys was studied at x = 0.2, 0.4, 0.6, and 0.8 [43]. The formation of alloys was found to be endothermic, and so, they would prefer to segregate at 0 K. However, very low energies calculated by Equation (1) suggest their formation under ambient conditions. Due to desirable Cr–Cr interactions, these atoms tend to form ordered lines in the MoS2 lattice. The strong hybridization of the Cr 3d and Mo 4d orbitals at the edge of the valence band and the dominant contribution of the Cr 3d orbitals to the edge of the conduction band cause a decrease in the band gap of the alloy with increasing concentration of Cr.
Hexagonal monolayers of MoS2 and WS2 have close lattice constants (the difference is ~0.2%), and so, the structural deformation due to the substitution of one metal for another is negligible, and the band gap depends on the charge exchange [44]. The charge transfer from weak Mo−S bonds to the nearest strong W−S bonds [45] causes negative formation enthalpies of hexagonal Mo1−xWxS2 for the entire range 0 ≤ x ≤ 1 [46]. At T = 0 K, solid solutions tend to chemical ordering, rather than clustering. However, random solid solutions can stabilize at elevated temperature due to the increasingly strong contribution of configuration entropy. The most stable alloys are formed in the x range from 0.33 to 0.66 [45,46,47,48]. In this range, the alloy formation energy is practically independent of the dopant concentration, and a large degree of disorder can be realized in the course of synthesis [49]. Since there are many energetically degenerate structures for a given composition, the alloys are disordered in the long range, while the position of foreign atoms in the lattice has little effect on the band gap [45]. The order–disorder transition temperatures for the Mo1−xWxS2 monolayers with 0.33 < x < 0.66 are estimated at approximately 88–102 K [50]. The W impurity in MoS2 causes a decrease in the band gap to a minimum value at x = 0.33 [47]. With a further increase in the content of tungsten, a parabolic rise in the band gap occurs due to the dominant contribution of W orbitals to the conduction band edge [51]. The band gap of the ordered phase is smaller than the band gap of the disordered phase for each concentration W [52]. The valence band edge only correlates with the x-composition, while the conduction band edge is very sensitive to the degree of disorder.
The screening of 26 metals with a concentration of 6.25% in a MoS2 monolayer was performed in [53]. The dopant formation energy was calculated using Equation (3); the results are shown in Figure 2. The higher the energy, the energetically easier it is to fill the molybdenum vacancy with an element. The gain in self-substitution of Mo is 13.5 eV. The most preferred metals for insertion are Ta, W, and Nb, and the introduction of transition metals Hg, Cd, Zn, Ag, Au, and Cu is difficult.
The introduction of metals into the hexagonal MoS2 lattice having fewer and more valence electrons than the molybdenum atom leads to the formation of acceptor levels (p-type doping) and donor levels (n-type doping), respectively.
Six metals, V, Nb, Ta, Mn, Fe, and Co, were used to obtain 4% impurities in MoS2 monolayers [54]. The formation energy was calculated using Equation (2). The negative formation energies found for Nb and Ta indicate that these elements are desired for incorporation into the MoS2 lattice. For Nb, it was shown that the value became more negative as the number of dopants increased [55]. Among the studied magnetic transition metals, the lowest formation energy was determined for Mn-doped MoS2 [54]. The energetically favorable formation of less than 5% Mn defects in MoS2 monolayers was also shown in [56]. The thermodynamic driving force of clustering of Mn dopants was predicted in [57]. Replacing Mo with metals having one less valence electron (V, Nb, and Ta) produced the states near the valence band, i.e., induced p-type doping [54]. The transition metals Mn, Fe, and Co have one, two, and three valence electrons in excess with respect to Mo, and these electrons occupied the impurity states appeared in the band gap [58]. The Mn dopant created impurity states close to the edge of the conduction band, while the corresponding states induced by Fe and Co were below. The small separation between the Fermi level and the conduction band of 0.48 eV makes Mn the most attractive for n-type doping.
The heteroatoms Y, Zr, Nb, Re, Rh, Ru, Pd, Ag, Cd were introduced into an MoS2 monolayer at a concentration of 4% and the formation energy was calculated according to Equation (2) [59]. It was shown that Nb and Zr are the most preferred substitutional elements. Re has one extra electron as compared to Mo, and its introduction into MoS2 creates a donor level 0.2 eV below the conduction band. This level is occupied by rhenium electron with the dz2 orbital character [60]. Rhenium can form complexes with a sulfur vacancy that quench n-type doping induced by isolated heteroatoms.
An increase in the occupation of the valence d-orbitals in the Re, Ru, Rh, Pd, Ag, Cd set results in a progressive increase in the states in the band gap [59]. Nb, Zr, and Y have, respectively, one, two, and three electrons less than Mo, resulting in p-type doping. In the case of Nb substitution, the newly created impurity states arise mainly from the hybridized d orbitals of Nb and Mo (Figure 3a) and the excess charge propagates up to the third nearest neighbor Mo atoms [61]. The donor states have a similar character for the Zr and Y heteroatoms, but they are less hybridized [59]. It was shown that, in contrast to Re, the Nb dopant and the S vacancy do not tend to be neighbors [60]. Moreover, the doping properties of Nb are not sensitive to the presence of sulfur vacancies. Since these vacancies are formed during synthesis, the introduction of Nb into the MoS2 layers is promising for p-type doping of the material.
Consideration of a MoS2 supercell with more than one Nb showed a tendency for heteroatoms to stay close to each other. The preferred formation of a triatomic cluster centered on a hole of a hexagon (Figure 3b) was established for the heteroatoms Nb, Ta, Tc, and Re [62]. This behavior was related to easily accommodate the optimal atomic spacing for inter-dopant bonding. The formation energy calculated by Equation (2) was fairly insensitive to the concentration of heteroatoms in the range from 4 to 33%. It was shown that grouped heteroatoms change the local chemistry of the material and, therefore, are not efficient for n- or p-type doping as compared to isolated heteroatoms.
A screening of several earth-abundant metals Ti, V, Cr, Mn, Fe, Co, Ni, Zn, Zr, Nb, Hf, Ta, W, and Bi showed most metals prefer to be located at the S-edge, and only Nb, Zr, Hf, and Ta were better stabilized at the Mo-edge [40]. The tendency to turn on at the edges creates a problem for the synthesis of basal-plane doped MoS2. The examination of five configurations of the Fe atom in the MoS2 monolayer showed that this could be achieved under S-rich conditions [63].
The introduction of a foreign metal into the MoS2 layer is necessary to create a larger number of active sites per host unit and improve the electrical conductivity [40]. DFT calculations of Ni heteroatoms in bulk 2H-MoS2 revealed an increased conductivity due to band gap narrowing [64]. Interestingly, Ni can also easily replace sulfur. This means that Ni can fill existing S vacancies, which could be a way to synthesize Ni-doped MoS2. The results on the formation energy of the Ni heteroatom in the MoS2 monolayer obtained by Equation (3) showed that the filling of the S vacancy requires energy, while the filling of the Mo vacancy releases energy. The Ni heteroatom in the position of the Mo atom is energetically attractive for the formation of a sulfur vacancy in close proximity [65]. Easy movement of a heteroatom adsorbed above the Mo atom in MoS2 to fill a sulfur vacancy was also shown for Co [66].

2.1.2. Sulfur Replacement

Molybdenum dichalcogenides MoCh2 (Ch = S, Se, Te) have a similar crystal structure and are, therefore, attractive for mixing in one compound. DFT relaxation of the structure of MoS2(1−x)Se2x and MoS2(1−x)Te2x alloys showed a linear behavior of the lattice constant with x (the Vegard’s law) [67]. The negative formation energy obtained for MoS2(1−x)Se2x over the entire range 0 ≤ x ≤ 1 indicates complete miscibility of MoS2 and MoSe2 at low temperatures. The most thermodynamically stable alloys are those with x = 0.33, 0.5, and 0.66. In these structures, clustering of S or Se atoms is unfavorable, and two different chalcogens alternate around the Mo atom to maximize the number of dissimilar atom pairs (S–Se) [68]. The charge transfer from Se to the nearest S atoms ensures the energy gain from mixing MoS2 and MoSe2 [45]. Such charge transfer always exists whether the alloy is ordered or disordered, but it can be maximized by local S–Mo–Se ordering. MoS2(1−x)Se2x monolayers retain a direct bandgap, the value of which decreases almost linearly with increasing x [67]. Mixing MoS2 with MoSe2 is more attractive than MoS2 with WS2 since the band gap in MoS2(1−x)Se2x can vary in the range of about 0.3 eV, which is larger than the value of about 0.15 eV for Mo(1−x)WxS2 [69].
The formation energy of MoS2(1−x)Te2x monolayers is positive, which indicates the absence of a stable configuration for these alloys at 0 K and phase separation in the alloys [67]. However, low values for the formation energy of a random configuration show that such alloy structures can be obtained at experimentally achievable temperatures. The calculated phase diagram showed an asymmetry that increased with temperature. The asymmetry means that, at a giving temperature, the introduction of Te atoms into MoS2 occurs more easily than S atoms into MoTe2. Full miscibility of MoS2 and MoTe2 is expected at about 493 K. The bandgap of random MoS2(1−x)Te2x showed bowling behavior with deviation from linearity decreasing at x = 0.5. The change in the bandgap was about 0.4 eV in the range 0 ≤ x ≤ 1, which is greater than that calculated for MoS2(1−x)Se2x. In addition, substitutional Te atoms led to a larger localization of electrons near the valence band edge than Se atoms [70]. Replacing the entire S layer with Se or Te layers is possible under Mo-rich conditions, and replacing S with Se is preferred over Te due to its larger ionic size (2.1 Å vs. 1.9 Å for Se).
The calculation of MoS2(1−x)Te2x monolayers with different Te distributions showed that configurations with ordered Te lines are thermodynamically stable at 0 K and compete with random alloys [71]. Configurations where the Te lines are far apart (to avoid clustering) have the lowest energy at each probed concentration x = 0.1, 0.3, 0.5, 0.7, and 0.9. The band gap depends on the composition of the alloy and covers the range of the visible spectrum.
The formation of line-ordered MoS2-based alloys was also predicted when sulfur was replaced by oxygen [72]. Oxygen belongs to the same group of the periodic table as the chalcogens and has the smallest atomic radii. It was shown that at large x in MoS2−xOx, line-ordered O alloys are more stable than random and cluster alloys, while at small x all configurations compete. The location of an equal number of O atoms in both layers sandwiching a Mo layer is energetically favorable for random alloys. All considered MoS2−xOx structures retain the semiconducting features of MoS2 and the band gap decreases with increasing x. At the same O content, line-ordered alloys have a smaller bad gap than alloys with a random and cluster distribution of O.
Halogens F, Cl, Br, and I have one additional valence p-electron with respect to S and should act as an n-type doping source for MoS2 [59]. In the example of a single Cl substituent, it was shown that the impurity states that appear at the edge of the conduction band arise as a result of hybridization between the Cl 3p and Mo 4d orbitals. Group V nonmetals of the periodic table (N, P, and As) have one less valence electron than sulfur. Replacing S with these elements in the concentration of 2% leads in the formation of impurity states above the valence band. In the case of As, these states merge with the valence band; therefore, this element is most promising for p-type doping [54,59]. An analysis of the density of states of the N-MoS2 monolayer indicated that the impurity states mainly arise as a result of hybridization between the N 2p orbitals and the 4d orbitals of the neighboring Mo [73]. The band gap narrows with an increase in the amount of N dopants and the monolayer becomes metal at a doping level of 12.5% [74].
C, Si, and Ge from group IV of the periodic table have a deficit of two valence electrons as compared to S. The introduction of 2% these heteroatoms into the MoS2 lattice leads to the appearance of impurity states far from the valence band, which makes them unsuitable for effective p-type doping [54]. The number of electrons is further reduced for group III elements, such as B, Al, and Ga. Dopant B induces impurity states closest to the valence band than any other elements from groups III and IV. It was concluded that elements P, As, and B are best suited for introducing gap states near the edge of the MoS2 valence band. Among these elements, the formation energy of B-MoS2 calculated by Equation (2) is smallest. However, for all elements of groups III–V, the formation energy has a positive value, which indicates they are unstable dopants at thermal equilibrium. The value decreases for the Mo-rich system [59]. Such conditions can occur at the Mo-terminated edges [40,75] or when sulfur vacancies are created in the layer.
S-vacancies can be created by irradiating MoS2 with 80 eV electrons, as shown theoretically and confirmed experimentally [76]. The possibility of filling a monoatomic S-vacancy with isovalent atoms (O, S, Se, and Te), donors (F, Cl, Br, and I), acceptors (N, P, As, and Sb), hydrogen and group IV elements (C and Si) was checked by calculating the formation energy of a heteroatom in a MoS2 monolayer according to Equation (3). It was found that all substitutions are energetically favored with respect to isolated atoms. Calculations for C, Si, and N with respect to the chemical potential of CH4, SiH4, and NH3 molecules gave positive values. However, substitution by these elements can be achieved under an electron beam that destroys the molecule.
For phosphorus, it was obtained that an increase in the dopant concentration from 2% to 6% reduces the formation energy of the P-MoS2 monolayer [77]. The charge density is accumulated around the P dopant and the unoccupied P-3pz orbital provides an active site for adsorbates, for example, hydrogen. A similar situation was observed for the N dopant due to charge transfer from less electronegative atoms Mo and S to its vicinity [73]. Among the calculated N-doped MoS2 monolayers, the lowest formation energy was obtained when N dopant was located near the S vacancy. The combination of a p-type dopant (substituent N) and impurity states induced by the S-vacancy near the conduction band led to a significant narrowing of the band gap of the MoS2 monolayer with such a defect complex (Figure 4) [78].
Easy filling of S-vacancies present in the MoS2 monolayer was also predicted for Cu and Ag [79] and Ge [80]. Cu and Ag have a much lower electronegativity (1.90 and 1.93) than sulfur (2.58), and, so their introduction in the layer causes a redistribution of charge on neighboring atoms [79]. As a result, polarized regions appear, activating the MoS2 surface for interaction with various chemical species. Unlike Cu and Ag, which have the lowest formation energy of a single heteroatom, Ge prefers to form a triatomic cluster [80]. The Ge dopants introduce new states in the band gap, thus enhancing the catalytic activity of MoS2 in oxygen reduction reaction.

2.1.3. Dual Replacement in MoS2

Simultaneous replacement of a part of molybdenum and sulfur atoms in MoS2 can be a way to fine tune the band gap, improve the thermodynamic stability of the compound due to an increase in entropy, and a stronger electron density redistribution along the layer. The isostructural nature of hexagonal MoS2, WS2, MoSe2, and WSe2, as well as the negative formation enthalpies of ternary alloys with isovalent cations and isovalent anions, make these disulfides and diselenides the most suitable for creating quaternary mixed compounds [45]. The DFT calculation of 152 random Mo1−xWxSe2−ySy structures with x = y showed a continues band gap variation from 1.60 to 2.03 eV [81]. On the contrary, for the ternary compounds MoSe2−xSx and Mo1−xWxS2, the band gap varies in narrow ranges 1.62–1.86 eV and 1.87–2.03 eV, respectively [49]. The largest change in band gap was found in Mo1−xWxSe2(1−x)S2x at x = 0.5 [81]. Quarterly alloys demonstrated a great sensitivity of the electronic structure to the distribution of elements in the lattice.
Quarterly monolayer structures were created by replacing the Mo atom in MoS2 with Nb and the S atoms surrounding the substitutional Nb with one to six Se or Te atoms [70]. The Nb concentration was 2.1%, the concentration of chalcogen heteroatoms varied from 2.1 to 12.5%. It was shown that the band gap of MoS2 containing Nb substituents with six Se neighbors is larger than that of the ternary analogue without Nb. The opposite behavior (band gap reduction) was found in the case of Te. The substitutional Nb atom generates acceptor levels, resulting in p-type doping of quarterly monolayers. These levels are more localized in cases with substitutional Te than in those with substitutional Se. However, calculations revealed difficulties with the formation of these quarterly compounds because Nb incorporation is more favorable under S-rich conditions, while chalcogens require Mo-rich conditions.

2.2. Experimental Detection

Data on experimentally realized heteroatom doping of MoS2 were summarized in recent reviews [35,36]. Most of the elements considered in theoretical studies were successfully incorporated into the MoS2 structure experimentally (highlighted in red in the table in Figure 1). However, s-block metals tend to intercalate between MoS2 layers instead of replacing Mo, which is considered theoretically [39], while F and Cl atoms are more likely to be chemically adsorbed on the surface of MoS2 than to replace sulfur.
Direct methods for visualizing foreign atoms in the MoS2 lattice are high-resolution transmission electron microscopy (HRTEM) and high-angle annular-dark field/scanning transmission electron microscopy (HAADF/STEM). A clear demonstration of the possibilities of these methods for studying substitution-induced structural changes of MoS2 is given in [36]. However, it is not always possible to distinguish elements with similar contrast level. In this case, electron energy loss spectroscopy (EELS) at the single-atom level can help identify the element and its local atomic configurations [82,83]. Studies showed that some elements tend to form clusters, among them Co [82], Te [84], Re with a concentration below 1 at% [85,86], Se [87], W [88,89], Mn [90], Mg [91], Sb [92], and Ru [93] are distributed randomly regardless of concentration. Figure 5a shows an example of the identification of two different Co defects in a MoS2 monolayer by the HAADF/STEM method. Co and S have similar contrast levels and the incorporation of Co was confirmed by the Co L2,3 edge detection in the EELS spectrum (Figure 5b). The images simulated for models with a triatomic Co cluster and an isolated single Co substituent well agree with the experimental images (Figure 5a).
The concentration and oxidation state of elements are usually determined using X-ray photoelectron spectroscopy (XPS). Some papers did not report binding energy calibration, making it difficult to compare their results with others. Calibration using the Au 4f5/2 peak and the Fermi level gives the position of the Mo 3d5/2 component at about 229.7 eV and the S 2p3/2 component at about 162.6 eV for 2H-MoS2 [94]. Baker et al., showed that the binding energy of the Mo 3d5/2 component decreases with sulfur depletion in the compound, and the difference between the Mo 3d5/2 and S 2p3/2 positions can be used to determine the stoichiometry of the MoSx sample [95]. Doping causes a shift of the Fermi level of the compound, resulting in a change in the binding energy of the core-level electrons as compared to the undoped compound [96]. For example, a decrease in the binding energy of the Mo 3d5/2 and S 2p3/2 components was observed for MoS2 doped with Mg [91], Ru [93], N [97], and Nb [98,99], while the introduction of Re [86] or Se [87] substituents resulted in an increase in the binding energies.
One of the most widely used methods for characterizing MoS2 nanomaterials is Raman spectroscopy. In-plane and out-of-plane S–Mo–S vibrations in the hexagonal MoS2 layer produce an E12g peak at ~382 cm−1 and an A1g peak at ~407 cm−1, respectively [100]. The difference in the ionic radii of molybdenum, sulfur, and the substitutional elements can lead to the appearance of disorder, compressive stresses in the lattice, and changes in Mo–S distances. The E12g mode is more sensitive to these lattice modifications than the A1g mode, which causes an increase in the separation of the corresponding Raman peaks of heteroatom-doped MoS2, as was observed for Se [101,102] and W [103,104]. Modifications of the MoS2 lattice can also be tracked by changes in the position and intensity of reflections on X-ray diffraction (XRD) patterns [99,105].
Distances between atoms in MoS2 containing heteroatoms can be determined using extended X-ray absorption fine structure (EXAFS) spectroscopy. EXAFS also defines a few coordination spheres of the specified element and types of neighbors [106]. The data of this method allow determining that foreign atoms are included in the host lattice and do not form separate clusters on its surface. For example, the study of Ni-doped MoS2 detected Ni–S bonds and no Ni–Ni bonds, indicating the atomic dispersion of nickel in the nanomaterial [107]. The same results were obtained for Nb-doped [108] and Co-doped [66] MoS2 in consistence with the HRTEM or HAADF/STEM data, which provide very local information about the sample structure.

3. Gas Sensors

Layered inorganic nanomaterials are attractive for the detection of chemical species due to their large specific surface area, semiconducting properties with an appropriated band gap, and strong surface activity. An overview of the application of these nanomaterials, including MoS2, in various types of gas-sensing devices can be found in [24,25]. A report on the development of MoS2-based sensors for detection of gaseous nitrogen dioxide NO2 is presented in [32]. The dependence of the charge transfer process on MoS2 morphology, phase composition, the presence of atomic vacancies and heteroatom impurities, and the formation of heterostructures with graphene derivatives and metal oxides was considered and deeply analyzed. The NO2 sensing mechanism using MoS2-based heterostructures and composites was discussed in detail in a recent review [109].
In this section, we consider chemiresistive MoS2 sensors whose response is determined by the change in resistance due to the adsorption of the molecule. First, we briefly present the influence of morphology and edge states on the detection limits of various gases with undoped MoS2 and theoretical explanations on the observed effects. Then, we concentrate on the effect of heteroatoms on the interaction of MoS2 with molecules, studied experimentally and theoretically.

3.1. Undoped MoS2

The first sensor device was fabricated in 2012 from mono- and few-layered MoS2 films mechanically detached from a single crystal [110]. These exfoliated films were n-type doped and their resistance increased when exposed to NO gas, indicating charge transfer from MoS2 to the adsorbate. Films consisting of two to four MoS2 layers exhibited high sensitivity to NO with a detection limit of 0.8 ppm (parts per million) at room temperature, while the response of the MoS2 monolayer was fast but unstable. The devices showed a slow complete desorption of adsorbates, which was related to the strong interaction between MoS2 and NO. DFT calculations of NO adsorption on the surface of monolayer, bilayer, and trilayer MoS2 showed week physical adsorption in all cases with an energy gain of less than 215 meV [111]. NO adsorption induces impurity states near the bottom of the conduction band, which increases the conductivity of the MoS2 layers. This improvement is greater for a monolayer; however, few-layered MoS2 exhibited a higher electron mobility and this may be the determining factor for better performance of a sensor made of few layers of MoS2.
The following year, four papers on MoS2 sensors were published at once. Late et al., confirmed that few layers of mechanically delaminated MoS2 have a more stable response to NO2 and NH3 as compared to a monolayer [112]. The resistance of the devices decreased when exposed to NH3, which is an electron donor. Yao et al., fabricated a sensor device from few-layered nanosheets obtained by exfoliating bulk MoS2 crystal in an ethanol/water mixture [113]. The sensor was able to detect 5 ppm NH3 in ambient conditions. Significantly higher sensitivity to gaseous NH3, below 300 ppb (parts per billion), was achieved using the MoS2 films obtained by chemical vapor deposition (CVD) method via sulfurization of sputtered Mo thin layers [114]. Perkins et al., demonstrated the high sensitivity of a mechanically detached MoS2 monolayer to trimethylamine N(CH2CH3)3 and acetone(CH3)2CO vapors [115]. These molecules are electron donors and electron acceptors, respectively, and their adsorption on the MoS2 surface led to the opposite change in the device conductivity. The detection limit of the MoS2 monolayer device was ~1 ppm for N(CH2CH3)3 and ~500 ppm for (CH3)2CO at room temperature.
The use of a CVD-grown MoS2 monolayer as the sensing element allowed reducing the detection limit of NO2 gas to 20 ppb and NH3 gas to 1 ppm at room temperature [116]. This study confirmed the higher sensitivity of the MoS2 basal plane to NO2 than to NH3 previously predicted based on adsorption energies obtained using DFT calculations [112]. The extremely large sensitivity of the sensor was associated with the large surface area of the monolayer, and it was proposed that the detection limit can be further reduced by optimizing the device configuration [116]. Approximation of the experimental data on the dependence of the sensor response to the concentration of the analyte determined the NO2 detection limit of 1.4 ppb for the MoS2 monolayer, which was synthesized by sulfurization of the MoO3 layer [117]. The sensor response was several times greater for 100 ppm NO2 as compared to the same concentration of H2S, methanol, SO2, and CO.
The results of studying the interaction of NO2 and NH3 molecules with a defect-free MoS2 monolayer by DFT methods are summarized in [118]. All studies agree that the interaction occurs due to vdW forces, and there are several different configurations of molecules with nearly identical adsorption energy and charge-transfer characteristics. Such behavior may be critical for sensor functionality because the response is not highly dependent on the specific location or configuration of the adsorbate.
MoS2 edges are more reactive that the basal plane [119,120]. This fact was used by Cho et al., to detect NO2 and ethanol [121]. Films with different orientation of the MoS2 layers were synthesized by sulfurization of pre-deposited Mo seed layers. The MoS2 layer growth direction changed from horizontal to vertical with respect to the substrate surface as the thickness of the Mo layer increased from 1 to 15 nm. The vertically aligned MoS2 layers exhibited responses enhanced by four and five times as compared to the horizontally oriented MoS2 layers for electron donor ethanol and electron acceptor NO2 molecules, respectively. The electrical response of the device correlated directly to the density of the exposed edged S. Detection limit determined for NO2 was 0.1 ppm. DFT calculations confirmed stronger binding of NO2 to the edge sites (Mo or S-terminated) as compared to the MoS2 basal plane sites. Edge-enriched MoS2 films were obtained by CVD method via sulfurization of MoO3 powder [122]. The films exhibited stable and reproducible responses to 2 ppm NH3 at room temperature and 20 ppb NO2 at a sensor operation temperature of 100 °C. Networks of MoS2 nanowires with abundant edges showed a low NO2 detection limit of 4.6 ppb at 60 °C [123]. Annealing the device in vacuum desorbed the oxygen-containing molecules from the MoS2 surface, thereby increasing the availability of the active edge sites for NO2 adsorption.
Mechanically detached MoS2 monolayers were used to study the change in electrical conductivity upon adsorption of H2 at 200 °C [124]. It was shown that H2 molecules donate elections to MoS2, and the device is capable of detecting H2 concentration down to 0.1%. Based on the DFT calculations, it was suggested that the cause of electron donation is the dissociative adsorption of H2 on sulfur vacancies [125]. A sensor element made of vertically edge-exposed MoS2 flakes was able to detect 1% H2 at 28 °C [126]. DFT calculations revealed the highest probability of H2 adsorption at the Mo atom sites of the MoS2 edges. The Mo-terminated edges predominate in sulfur-deficient MoS2 [127].
High-performance humidity sensor was prepared from sonication-based exfoliated MoS2 flakes [128]. In wet air, the electrical conductivity of the sensor increased and changed by six orders of magnitude in the range of the relative humidity 10–95%. O2 gas was shown to act as an electron acceptor relative to the MoS2 sensor made from mechanically detached few-layer flakes [129]. The adsorption/desorption of O2 enhanced with applying a positive/negative voltage. DFT calculations showed a weak adsorption energy of 79 meV for O2 and 110 meV for H2O on the surface of a MoS2 monolayer [130]. The adsorbed molecules take 0.04e and 0.01e, respectively, from MoS2. Adsorption on sulfur-vacancy sites increases the charge transfer per molecule by a factor of five.
Nanospheres composed of thin MoS2 sheets grown in a hydrothermal process was used a sensor for the electron acceptor CO gas and showed a detection limit of 50 ppm at 175 °C [131]. The high response was attributed to point defects at the sheet edges. Indeed, according to DFT calculations, CO molecules strongly bind to unsaturated Mo atoms [132].
A comparative study of the adsorption properties of the MoS2 monolayer with respect to CO, CO2, NH3, NO, NO2, CH4, H2O, N2, O2, and SO2 with the DFT revealed a stronger binding for NO, NO2, and SO2 molecules [133]. Higher binding energy is usually associated with higher sensor sensitivity, since the MoS2 detection mechanism is based on charge transfer [134]. In addition, the adsorption of NO, NO2, and SO2 induces additional states near the edge of the conduction band of MoS2, while the other considered molecules have little effect on the band gap of MoS2 [133]. Thus, the basal plane of undoped MoS2 is most suitable for the detection of NO, NO2, and SO2 gases. The calculation of current-voltage curves revealed higher current for a bilayer MoS2 with NO2 as compared to a monolayer sensor [135].

3.2. Heteroatom-Doped MoS2

The significantly higher response observed experimentally for NO2 as compared to other gases [117,122,123,136,137] may indicate that the MoS2 sensor is selective to this toxic gas. Above, we demonstrated that sensor performance can be improved at elevated operation temperature [116,122,124,126] and by creating active sites such as edges [121] and atomic defects [126,131]. Heteroatom substituents are potential candidates not only for increasing the sensitivity of the sensor, but also for making it selective.

3.2.1. Experimental Data

In a review on molybdenum-based gas sensors published in 2021, the replacement of sulfur or molybdenum with foreign atoms was considered as one of the strategies to improve the ability of MoS2 to detect gases [138]. However, the review mentioned only three experimental works on this topic published up to that time. Shao et al., synthesized ultrathin Zn-doped MoS2 nanosheets using a hydrothermal process with a zinc content ranging from 1 to 15% [139]. At room temperature, the chemiresistive sensor with 5% Zn showed a significantly higher response to O3 and NO2 as compared to undoped MoS2 synthesized by the same method. The detection limit of the 5% Zn-MoS2 sensor was 17 ppb for O3 and 215 ppb for NO2 at room temperature. Taufik et al., irradiated solvothermally prepared MoS2 nanoparticles with O2 plasma to replace sulfur by oxygen [140]. This treatment made the MoS2-based nanomaterial more sensitive to humid air in the range of 50–95%. The response to 95% humidity was much higher than the response to 100 ppm of interfering gases (acetone, ethanol, and oxygen), suggesting the applicability of the developed nanomaterial for a breath sensor. Zhang et al., used a hydrothermal route to introduce Ni, Fe, or Co into layers of MoS2 flower-like nanoparticles [141]. The doped sensors performed better for SO2 gas from 0.25 to 4000 ppm than undoped MoS2 (Figure 6a). The highest response was observed with the Ni-doped MoS2 sensor with a detection limit of 250 ppb at room temperature. The DFT calculations performed in this work confirmed that the SO2 molecule interacts more strongly with the nickel atom inserted into the MoS2 monolayer than with the inserted iron and cobalt atoms. The Ni-doped sensor exposed to 500 ppm of SO2, NO2, NH3, CO, CO2, and H2 at room temperature demonstrated a significantly higher response to SO2 (Figure 6b). Using a similar hydrothermal synthesis but with different precursors, Bharathi et al., synthesized flower-like nanoparticles composed of MoS2 layers with a nickel content ranging from 0 to 7 at% [142]. At room temperature, a sensor made from 7% Ni-MoS2 showed a higher response to NO2 gas than undoped MoS2. In contrast to a previous work, where the Ni-doped MoS2 sensor was selective to SO2 (Figure 6b), the 7% Ni-MoS2 showed NO2 selectivity among other tested gases H2S, NH3, CO, N2O, CO2, and SO2. Bharathi et al., suggested that in their nanomaterial, Ni atoms are located above the MoS2 layer, rather than replacing Mo atoms in the lattice. Unfortunately, in both compared works, the arrangement of Ni atoms was not proven by the HRTEM and/or EXAFS methods.
Taufik et al., used hydrothermally synthesized MoS2−xSex (x = 0.2, 1.0, and 1.8) alloy nanoparticles to fabricate a room-temperature NO gas sensor [143]. The EXAFS study revealed the coexistence of Mo–S and Mo–Se bonds in the alloys. The MoSSe sensor showed a higher response as compared to other alloys and MoS2 and MoSe2 (Figure 7a) with a detection limit of 270 ppb. DFT calculations identified that NO tilted by the nitrogen atom for the surface of MoS2, MoSe2, and MoSSe monolayers, and in the latter case, the distance is shorter. In addition, the highest charge transfer was found for the NO + MoSSe system, which may be the reason for better detection performance of the MoSSe sensor. The response of this sensor to NO was stable under various humidity conditions and significantly higher than to ethanol, toluene and methanol vapors, and gaseous H2 (Figure 7b). Wu et al., synthesized N-doped MoS2 nanosheets with nitrogen contents of about 7, 10, and 13 at % using a modified solvothermal route [144]. The formation of Mo–N bonding was observed according to XPS data. Nanomaterials were used to detect NO2 gas at room temperature. All N-doped sensors demonstrated faster response and recovery as compared to undoped MoS2. The 10% N-MoS2 sensor, which showed the highest response among the examined devices, was taken to systematically evaluate overall sensing performance. This sensor was able to detect 125 ppb NO2 and at 10% relative humidity and it was highly selective for this gas among other CH4, NH3, H2S, and CO analytes. DFT calculations revealed that the NO2 molecule is located closer to the MoS2 monolayer with substitutional nitrogen than to the ideal monolayer, and the adsorption energy is larger in the former case.
Ramaraj et al., compared the performance of Nb-doped MoS2 and MoS2 monolayers synthesized using a physical vapor deposition (PVD) technique, then transferred onto flexible polyethylene terephthalate substrates, and tested for NO2 gas detection at various operation temperatures [145]. Compared to undoped MoS2, the Nb-doped sensor showed improved response over the NO2 concentration range of 5 to 16 ppm and more stable recovery. This behavior was associated with the higher adsorption energy of the NO2 molecule on the embedded Nb, obtained using DFT calculations. The response to 5 ppm NO2 was significantly higher than to the same concentrations of toluene, acetone, methane, and CO, demonstrating sensor selectivity at 50–200 °C. Zhang et al., synthesized large-areas undoped MoS2 and Nb-doped MoS2 monolayer films by the metal organic CVD method [146]. The XPS data determined about 6 at% Nb and an upshift of the valence band maximum for the latter film due to p-type doping. Since the type of doping of the films was different, they showed an increase (undoped MoS2) and a decrease (Nb-doped MoS2) in conductivity when exposed to trimethylamine. The Nb-doped sensor exhibited a 50-times-higher signal-to-noise ratio as compared to undoped MoS2. The signal was well detected by the former sensor at a trimethylamine concentration of 15 ppb. A more efficient charge transfer between the analyte and the Nb-doped sensor was associated with the proximity of the impurity states induced by Nb heteroatoms to the highly occupied molecular orbital of the adsorbed molecule.
Table 1 summarizes the data on heteroatom-doped MoS2 gas sensors together with best undoped MoS2 sensors. The parameter being compared is the detection limit of the analyzed gas at a given operation temperature. The most tested gas is NO2 since the conductivity of MoS2 changes strongly after adsorption of NO2 molecules. The lowest detection limit of 1.4 ppb NO2 was achieved using a CVD-grown MoS2 monolayer at room temperature. Note that many sensors fabricated from undoped MoS2 nanomaterials have sub-ppm levels of NO2 detection, which is ensured by the introduction of edge and defect states into the lattice and/or an increase in the operation temperature of the device. The developed heteroatom-doped MoS2 sensors also detect sub-ppm concentrations of NO2, and N-doped MoS2 demonstrated the lowest detection limit. Heteroatom doping made it possible to significantly improve the response of the MoS2-based sensor to NO and N(CH2CH3)3 gases.

3.2.2. DFT Calculation of the Adsorption Energy

The role of DFT calculations in unlocking the potential of two-dimensional nanomaterials for gas sensing applications was discussed in review [147]. Calculations are needed to explain the experimental results, to deeply understand the mechanism of the gas detection, and to provide guidance on the design of nanomaterials that are sensitive and selective for a particular analyte. The interaction of a molecule with the sensor surface is explored by analyzing the adsorption energy, charge transfer, density of states, band gap structure, and geometric parameters for optimal adsorption configurations. The density of states and the band gap structure characterize the change in the electronic properties of sensor after gas adsorption. The main calculation parameter is the adsorption energy, which is, in the case of a MoS2 sensor, defined as:
Eads = EMoS2+mol − (EMoS2 + Emol),
where the terms are the total energies of the optimized MoS2 structure with an adsorbed molecule, an adsorbate-free structure and a free molecule. A negative Eads indicates that the adsorption process is exothermic and can occur at room temperature. The greater the energy value, the stronger the interaction between the components. Charge analysis is performed using the Mulliken, Hirshfield, Löwdin, or Bader schemes, and the analyzed value is the charge transfer (ΔQ) between the molecule and the substrate. A positive (negative) charge that appears on an adsorbed molecule means it donates (accepts) electron density to (from) the sensor. The sensor–adsorbate distance is an indicator of the nature of adsorption. If the distance is larger (less) than the sum of the radii of closely contacting atoms from the surface and the molecule, adsorption is physical (chemical).
To characterize the interaction of heteroatom-doped MoS2 and a gas molecule, the three calculated parameters Eads, ΔQ, and the shortest sensor–adsorbate distance (d) are compared in Table 2. Only the results obtained using the generalized gradient approximation (GGA) for the exchange-correlation Perdew–Burke–Ernzerhof (PBE) functional with allowance for long-range interactions were considered. These interactions are especially significant in the case of physical adsorption and they were taken into account by the Grimme dispersion corrections D, D2, D3, or the vdW correction to the density functional [148]. In all cases, a MoS2 monolayer was considered, and a foreign atom was introduced into a sulfur vacancy.
DFT studies showed that the perfect MoS2 is not suitable for detecting gas molecules, while doping with a heteroatom usually has a positive effect. A systematic study of the effect of various transition metals (Fe, Co, Ni, Cu, Ag, Au, Rh, Pd, Pt, and Ir) embedded in a MoS2 monolayer on the adsorption of CO, NO, O2, NO2, and NH3 was presented in [149]. Foreign metals significantly modulate the electronic structure of MoS2, which enhances the adsorption of gas molecules. An important role in this is played by impurity states arising in the MoS2 band gap. Among the considered metals, Fe and Co are the best candidates for improving the sensor performance of MoS2. Szary investigated the sensitivity of MoS2 doped with transition metals Ti, Ni, Pd, and p-block elements Si, Ge, P, and Cl toward ethylene oxide (C2H4O) [150]. Ti, Ni, Pd, and Si atoms have been shown to facilitate the interaction of MoS2 with C2H4O, and a higher sensor response is expected for Si-doped MoS2. Deng et al., showed that undoped MoS2 is insensitive to the formaldehyde (CH2O) molecule, while its doping with Ni, Pt, Ti, and Pd leads to an increase in the adsorption energy and the largest gain is provided in the case of Ti [151]. Song and Lou considered the effect of Ag-doping on the interaction of a MoS2 monolayer with NO2, NH3, H2S, SO2, CO, and CH2O gases [153]. All molecules are chemically adsorbed on the Ag-doped surface by bonding with the foreign metal. Stronger adsorption was found for NO2. The molecule is oriented by oxygen atoms to the Ag substituent (Figure 8a) and in this case, the charge transfer to the adsorbate is six times higher than in the case of undoped MoS2 [159] (Figure 8b). In both cases, most of the transferred electron density was accumulated between substrate and NO2. According to [154], MoS2 doped with Ag is inert to NO gas, while the Au-doped MoS2 is sensitive. Co-doping of Au and Ag promises to create a highly sensitive MoS2-based sensor for NO2. MoS2 doped with V, Nb, or Ta was proposed to be suitable for detection of CO, H2O, and NH3 molecules [155], while Si-doped MoS2 could be a good sensor for SF6 [156].
The theory shows that doping of MoS2 with metals is, in general, a more efficient way to fabricate sensitive and selective sensors than doping with non-metals. Heteroatom-doped MoS2 donates electron density to CO, NO, NO2, and SO2 molecules and accepts it from NH3, C2H4O, SOF2, and H2O. The behavior, with respect to H2S and O2, depends on the type of heteroatom (Table 2). The highest binding energy is usually obtained for the NO2 molecule, which agrees with the experimental data.

4. Electrochemical Energy Storage

The ever-increasing demand for renewable energy is driving the search for new, highly efficient energy storage materials. Today, LIBs are widely used in various portable devices, but the vector is shifting towards the development of SIBs and PIBs, which are particularly required in stationary energy storage installations and electric vehicles [160]. Supercapacitors complement rechargeable batteries, as they provide high powder density and have long life. MoS2 is attractive for energy storage applications due to the weak vdW coupling of the layers and the large distance between them. This facilitates the metal ion intercalation and diffusion and is especially important for large ions like Na+ and K+ [161]. There are many simple methods for the synthesis of MoS2, which allow varying the architecture and the number of adjacent layers in a nanomaterial [162], thereby adjusting its electrical conductivity and the density of active sites for ion adsorption, which is important for high electrode capacity and its stability. The main disadvantages of MoS2 is its low conductivity, which prevents rapid electron transfer that occurs in batteries and supercapacitors, and the tendency of layers to aggregate, which reduces ion diffusion and structure stability [163,164]. In addition, the electrochemical activity of MoS2 decreases due to the sluggish reaction kinetics [165].
Rational structural engineering of MoS2 nanomaterials by controlling the shape and phase, chemical doping, interlayer spacing, type and defect density, etc., leads to an increase in capacity, power capability, and cycle life of electrodes [165,166]. The role of defects in MoS2 performance in LIBs, SIBs, and PIBs is analyzed in [167]. The review emphasized that structural defects modify the chemical and electronic properties of nanomaterials, create more storage spaces for ions (thus enhancing capacity), reduce the stress between adjacent layers (which affects the insertion/extraction of ions), and promote charge transfer in electrochemical processes.
We focused on the effect of substitutional doping MoS2 with heteroatoms on the performance of nanomaterials in LIBs, SIBs, PIBs, and supercapacitors.

4.1. Lithium-Ion Batteries

The electrochemical interaction of MoS2 with lithium ions includes intercalation and conversion reactions. The intercalation reaction proceeds at potentials above 1.1 V vs. Li/Li+ and provides a capacity of 167 mAh∙g−1 in the formation of the LiMoS2 compound [10]. With further insertion of lithium ions into the interlayer space, the MoS2 lattice is destroyed with the formation of Mo clusters and lithium sulfide Li2S [168]. The reversibility of this reaction, i.e., the recovery of MoS2 after the extraction of lithium from the electrode material, is still the subject of discussion. The most common point of view is that the reaction is irreversible [169,170]; however, the studies of the reaction products sometimes reveal the retention of MoS2 layers even after several repeated intercalation/de-intercalation cycles [127,171,172]. Synchrotron X-ray absorption spectroscopy and Raman spectroscopy studies on lithiated and delithiated MoS2 electrodes concluded that the reaction may be partially reversible [173], and nanostructuring of MoS2 [127] and/or its hybridization with graphene [174] can contribute to its reversibility.
The theoretical capacity of MoS2 provided by the intercalation reaction and the conversion reaction is 669 mAh∙g−1, due to the transfer of four electrons [10]. This value is almost two times higher than that for graphite, used as an anode material for commercial LIBs. After converting LixMoS2 to Mo and LixS2, the following electrochemical processes involve reversible reactions between sulfur and lithium ions similar to those that occur in lithium–sulfur batteries [175]. The theoretical capacity of sulfur for lithium ions is 1672 mAh∙g−1, and such values were observed for MoS2-based electrodes [176,177,178]. Note that the electrode material contains the heavy element molybdenum, and if we recalculate the specific capacity of the material only by the mass of sulfur, the experimental values of the capacity exceed the theoretical value. The enhancement may be related to the reversible interaction of Mo clusters with sulfur [179].
The main problem of MoS2 electrodes is the rapid capacity fading during LIB operation [161,180]. The reasons for this are the poor electrical conductivity of MoS2, the expansion of the lattice in the volume during intercalation, and the dissolution of lithium polysulfides Li2Sn (2 ≤ n ≤ 8) in organic electrolytes, which leads to a loss of electrochemically active sulfur. These disadvantages can be eliminated by hybridizing MoS2 with a carbon component, creating composites and heterostructures with other compounds, and inserting foreign elements into the MoS2 lattice.
The first demonstration of the performance of the MoS2 material as a LIB anode is dated 2009 [181]; therefore, there are not so many works on the heteroatom-doped MoS2 anodes. Most of the research in this area is devoted to N-doped MoS2. Liu et al., synthesized MoS2 nanoparticles by a hydrothermal route and then heated them in a flow of Ar or NH3 at 700 °C for 2 h [182]. According to XPS data, the nitrogen content in the latter sample was ~5 at%. N-doped MoS2 showed an initial specific capacity of 1130.8 mAh∙g−1 at a current density of 0.2 C (134 mA∙g−1), which was ~1.2 times higher than that of undoped MoS2, and was able to reversibly deliver ~800 mAh∙g−1. This improvement was preserved at continues cycling, while both nanomaterials constantly lost their capacity. Based on electrochemical impedance spectroscopy (EIS) data, the authors concluded that doping of MoS2 with nitrogen significantly reduces the charge resistance in the electrode. DFT calculations for MoS2 and N-doped MoS2 monolayers revealed a narrowing of the band gap by about 0.2 eV in the latter case, which could provide faster electron and ion transport. Li et al., used a similar procedure to introduce nitrogen into the MoS2 layers, but the NH3 treatment temperature was 300 °C [183]. N-doped MoS2 electrode had an initial specific capacity of 1186 mAh∙g−1 at 0.1 C, which slowly but gradually decreased over next 100 operation cycles of the LIB. Jiao et al., synthesized flower-like nanoparticles from graphene oxide (GO) and MoS2 by a hydrothermal method and they then treated the composite with N plasma at room temperature to reduce GO (rGO) and introduce nitrogen into both components of the nanomaterial [184]. Plasma treatment greatly increased the capacity and improved the stability of the electrode as compared to undoped one. The initial specific capacity of N-rGO/MoS2 was 726.0 mAh∙g−1 at 1 C and the capacity retention after 100 cycles was 81.5%. Later, researchers from this group modified the synthesis procedure by adding L-cysteine as a spacer between the MoS2 layers [185]. This structural modification increased the initial capacity of N-rGO/MoS2 to 1725.6 mAh∙g−1 at 0.1 A∙g−1, which is 1.6 times larger than that of the composite nanomaterial without N plasma treatment. The N-rGO/MoS2 electrode showed stable operation in LIB for 400 cycles at a current density of 0.5 A∙g–1. Based on experimental and theoretical studies of N-doped graphene/MoS2 composites, Cho et al., found that both of these components provide high chemisorption energy for lithium polysulfides, which can increase life of the electrode [186].
Fayed et al., showed that the use of carbon-containing compounds in the hydrothermal synthesis of MoS2 nanomaterials actually leads to co-doping with nitrogen and carbon [187]. The initial specific capacity of the electrode was 1280 mAh∙g−1 at 0.1 A∙g−1; after 60 cycles, it decreased to about 400 mAh∙g−1.
Two works were devoted to the study of the effect of heteroatoms on the intercalation reaction of MoS2 with lithium ions. Kotsun et al., synthesized N-doped MoS2 plates by rapid thermolysis of ammonium tetrathiomolybdate (NH4)2MoS4 in an ammonia atmosphere and tested them in LIBs at potentials above 1.1 V [78]. The nanomaterial with a nitrogen content of about 4% had a specific capacity of 189 mAh∙g−1 at 0.05 A∙g−1, which is higher than the theoretical value for ideal MoS2. DFT calculations showed that in the interlayer space of the N-MoS2 bilayer, lithium prefers to bind with two nitrogen atoms from opposite layers. Even when lithium is far from the incorporated N, the adsorption energy is higher than that for an ideal bilayer. This behavior provides a higher capacity for N-doped MoS2. Gong et al., used a hydrothermal method to decorate a porous graphene aerogel with oxygen-incorporated MoS2 and investigated a self-supported electrode for lithium intercalation/de-intercalation in the potential range from 3 to 1 V vs Li/Li+ [188]. The electrode showed high rate performance and outstanding durability, retaining about 91% capacity at a current density of 2 A∙g−1 after 3000 cycles.
Lu et al., synthesized Sn-doped MoS2 flowers (Figure 9a) using SnO2 as the starting compound in a solvothermal synthesis [189]. Nanomaterials with a tin content of 4.3, 7.9, and 14.9 at% showed higher stability over 100 cycles as compared to undoped material (Figure 9b). The 7.9% Sn-doped MoS2 exhibited an excellent initial capacity of 1087 mAh∙g−1 at a current density of 0.2 A∙g−1, which remained constant after 100 operation cycles of the LIB.
Lei et al., introduced 17.6% vanadium in MoS2 using vanadocene in a solid-state reaction [190]. The initial electrode capacity of 814 mAh∙g−1 at 1 A∙g−1 decreased to 350 mAh∙g−1 by the 300th cycle. The study of few-layer MoS2−xSex nanosheets synthesized using MoO3 nanowires as a skeleton structure, and a source of sulfur/selenium under hydrothermal conditions showed that x = 0.25 is optimal to provide a high reversible specific capacity of ~1077 mAh∙g−1 at 0.1 C in LIBs [191]. The incorporation of selenium not only improved the electrode capacity, but also accelerated the movement of lithium ions, which led to stable operation of the battery at 10 C.
The most impressive performances were obtained when heteroatom-doped MoS2 was combined with a carbon component. Wang et al., grew mesoporous Se-doped MoS2 layers vertically to the surface of rGO and showed high rate capability and good cycling stability for the composite with MoS1.12Se0.88 [192]. From a comparison of the performance of this composite with and without the carbon component, the authors concluded that graphene improves the conductivity and flexibility of the electrode material and reduces its volumetric expansion during the discharge/charge of LIBs. Francis et al., grew P-doped MoS2 nanoparticles on carbon cloth using sodium hypophosphite as the phosphorus source in hydrothermal synthesis [193]. The nanomaterial was used as a binder-free anode in LIB and exhibited an initial capacity of 2700 mAh∙g−1 at a current density 0.1 mAh∙g−1, which decreased to 900 mAh∙g−1 after 50 operation cycles. The authors supposed that P–O bonds reduce the agglomeration of MoS2 nanoparticles. Wang et al., used the same synthesis method to attach Mn-doped MoS2 nanosheets on a hierarchical carbon cloth [194]. The Mn-doped electrode showed a higher specific capacity as compared to the undoped one, especially at high current densities (Figure 10a), which, according to DFT calculations, is explained by increased conductivity and a lower diffusion barrier for lithium along the Mn-MoS2 layer (Figure 10b).
Qi et al., obtained ultra-thin Fe-doped MoS2 nanosheets on rGO using iron sulfite in hydrothermal synthesis [195]. The role of iron in providing a high reversible specific capacity of 946 mAh∙g−1 at 0.1 A∙g−1 after 1000 operation cycles of LIB was associated to an increase in the density of lattice defects and a decrease in electrode resistance. The Co-doped MoS2/rGO nanomaterial, in which a Co atom replaced every fifth Mo atom, was able to deliver a specific capacity of 894 mAh∙g−1 at a current density of 1 mAh∙g−1 [196]. An increase in the cobalt content in MoS2 provided excellent rate performance of the Co1/3Mo2/3S2/graphene nanocomposite, which showed stable operation at a current density of 50 A∙g−1 [197]. Han et al., synthesized nanoparticles from N-doped graphene and layers of MoS2 doped with Fe, Co, or Fe, Co by thermolysis of a mixture of compounds containing these elements [198]. The inclusion of both metals (Fe and Co) in the composition of the nanomaterial provided the highest specific capacities in the range from 0.1 to 20 C and high electrode stability at a current density of 5 C for 3000 operation cycles. The FeCo-MoS2/carbon electrode was assembled with commercial LiFePO4 into a full LIB cell, which, after 200 cycles, delivered 127 mAh∙g−1 at 1 C, demonstrating a capacity retention of 91.8%. The magnetic atoms in the composite electrode ensured efficient separation of the MoS2 and carbon layers, high electrical conductivity, and fast reaction with lithium ions.
Data on the initial specific capacity provided by heteroatom-doped MoS2 anodes in LIBs and the reversible specific capacity after several operation cycles are collected in Table 3. Among the dopants probed, nitrogen substitution gave the lowest values of specific capacity. Sn, Mn, and Co are very promising substituents, and the excellent performance obtained for FeCo-doped MoS2/carbon encourages the search for optimal combinations of dopants to ensure high rate capability, capacity retention, and life of the electrode material.
DFT calculations of lithium interaction with MoS2 monolayers containing Fe, Co, Ni, Cu, and Zn instead of Mo or N, P, As, F, Cl, and I instead of S revealed higher adsorption energies near the substituents as compared to undoped MoS2 [199]. Doping with heteroatoms does not significantly affect the barriers for lithium diffusion through the layer, except for positions near the substituent.

4.2. Sodium-Ion and Potassium-Ion Batteries

Sodium and potassium belong to the same group of the periodic table of chemical elements as lithium and, therefore, have similar (electro)chemical properties. Thus, the methods of synthesis and characterization, as well as the design of electrochemical cells developed for LIBs, can be transferred to sodium and potassium analogues. Moreover, these chemical elements, due to their low cost, are alternative candidates for rechargeable batteries [160]. MoS2 possessing a large interlayer distance is a promising host for large ions and is, therefore, actively studied as anode materials for SIBs and PIBs [161,200]. The current understanding of the basic principles of charge storage in SIBs and PIBs can be found in [161]. In contrast to the lithiation process, in which the destruction of the MoS2 lattice occurs after the formation of the LiMoS2 compound, sodium intercalation is possible up to 1.75 Na per MoS2 structural unit [201]. Electrochemical intercalation of potassium ions can proceed until 0 V vs K/K+ [202], which is different from the cases in LIBs (~1.1 V vs. Li/Li+) and SIBs (about 0.8 V vs. Na/Na+).
Heteroatom-doped MoS2 nanomaterials are usually combined with a carbon component (Table 4) to improve the electrical conductivity and structural stability of the electrode material in the SIB. This component is introduced intentionally by adding graphene oxide [197,203] or a carbon source [198,204,205,206,207,208,209,210] in the synthesis or is formed by the decomposition of a carbon-containing solvent [206]. Wang et al., used a nitrogen plasma treatment of rGO/MoS2 for the doping [203]. Nitrogen is most often used as a substitutional element for MoS2 doping (Table 4), as it easily incorporates into the MoS2 lattice under hydrothermal conditions [204,205,208] or thermal decomposition of thioacetamide [206,207]. As a rule, the formation of the Mo–N bond is confirmed by the component located around 396.4–396.9 eV in the XPS spectrum. Sometimes, the choice of this component is not obvious, since the N 1s and Mo 2p3/2 regions overlap. In all cases, the authors observed a higher specific capacity of N-doped electrode as compared to undoped electrode prepared under similar conditions. This improvement is associated with an increase in the interlayer distance and conductivity of MoS2, which facilitates the transport of sodium ions [203,206,207]. For example, the diffusion coefficient of the ions in N-rGO/MoS2 was 1.64 times higher than that in the undoped counterpart [203]. Regardless of the synthesis protocol, the N-doped MoS2/carbon composites exhibited similar values of the reversible capacity after several hundred cycles of SIB operation (Table 4). Note that the carbon component increases the capacity, and its contribution depends on the ability to accumulate sodium ions, fractions in the composite, and interaction with MoS2.
The substitution of sulfur for antimony induced the transformation of semiconducting 2H phase of MoS2 to metallic 1T phase [209]. The obtained Sb-doped MoS2/N-carbon electrode was able to deliver 183 mAh∙g−1 at 5 A∙g−1 and showed a superior long-term operation at 5 A∙g−1 (Table 4). The material was used to assemble a full SIB with Na3V2(PO4)3 cathode maintained a capacity of 242 mAh∙g−1 at 0.5 A∙g−1 after 100 cycles. After the same number of cycles, the full cell with the N-doped MoS2 foam anode showed only 143.6 mAh∙g−1 at 0.33 A∙g−1 [205]. Zong et al., prepared Te-doped MoS2 nanosheets on the surface of tubular carbon and then coated the composite with carbon [211]. The nanomaterial showed a high rate performance in SIB, delivering 255.1 mAh∙g−1 at a current density of 10 A∙g−1. The replacement of sulfur by tellurium was shown to increase the interlayer distance in MoS2 to 0.95 nm, cause the formation of the 1T phase, and promote the adsorption of Na+ ions.
Woo et al., synthesized the carbon-free undoped and Re-doped fullerene-like MoS2 nanoparticles by treating MoO3 and RexMo1−xO3 (x < 0.01) with H2S in a reducing atmosphere at 800–870 °C [212]. Tests in the range between 0.7 and 2.7 V vs. Na/Na+, when intercalation occurs, showed better cycling performance of the doped electrode. Crystal defects and dislocations served as channels for the insertion of Na+ ions and the number of these channels was greater in Re-doped MoS2. After full desodiation of the electrode, the crystal structure of MoS2 was not restored, according to XRD data.
Similar to N substituent, the effect from the incorporation of foreign metals in the MoS2 lattice is also considered as an enlarged interlayer spacing and improved conductivity of the electrode material. In the case of Mn-doped MoS2 nanotubes, this caused an increase in the Na+ diffusion coefficient by 2.9 times as compared to undoped nanotubes [213]. DFT calculations revealed that the substitution of every fourth molybdenum by a vanadium atom almost halves the energy barrier for the migration of sodium ions in the interlayer space [214]. This result was used by the authors to explain the excellent rate performance of VMoS2 electrode, which was stable at a current density of 20 A∙g−1. Furthermore, these authors modified the synthesis procedure to obtain the orderly layered VMoS2 [215]. The edges of stacked nanosheets provided additional storage sites for Na+ ions, and the anode showed a specific capacity of 497.5 mAh∙g−1 at 5 A∙g−1 with a Coulombic efficiency of up to 100% at the 490th cycle. A similar capacity value of 491 mAh∙g−1 at 5 A∙g−1 was obtained for the Co1/3Mo2/3S2/r-GO composite (Figure 11) [197]. The Ni-MoS2@porous carbon anode provided a reversible capacity of 420.8 mAh∙g−1 at 5 A∙g−1 and 386.6 mAh∙g−1 at 5 A∙g−1 [210]. The DFT calculations revealed that Ni substituent can significantly improve the adsorption of Na+. The alternation of Fe and Co-doped MoS2 layers and N and O-doped graphene layers in the nanomaterial provided high capacities ranging from 735.7 to 340.8 mAh∙g−1 at current densities from 0.1 to 20 C [198]. DFT calculations gave energy barriers of 0.26 and 0.14 eV for sodium migration in the space between graphene and MoS2 layers and graphene and metal-doped MoS2, respectively. Therefore, the charge redistribution in MoS2 caused by the Fe and Co co-doping accelerates the diffusion of ions. Moreover, the incorporation of foreign metals makes the bilayer metallic, which should facilitate the transport of electrons in the electrode.
Potassium ion has a larger radii (0.138 nm) than sodium (0.102 nm); however, its low reduction potential (−2.93 V vs. the standard hydrogen electrode) indicates possible high energy density for PIBs [216]. Insertion of heteroatoms in the MoS2 lattice increases the interlayer distance [209,217,218], narrows the bandgap [216,219,220], and increases the diffusion coefficients of potassium ions [217,218,219,220,221]. The values of the initial specific capacity and reversible capacity during long-term cycling of heteroatom-doped MoS2 anodes are presented in Table 5. The performance depends on the nature and content of foreign element, the presence of carbon component, morphology, and phase composition.
Liu et al., sulfurized a mixture of chitosan and phosphormolybdic acid at 600 °C to obtain ultrathin MoS2 nanosheets evenly embedded in a nitrogen-doped carbon matrix [209]. The addition of SbCl5 in the reaction mixture led to the formation of the 1T phase of MoS2 doped with Sb. The doped electrode showed outstanding potassium storage performance: 150 mAh∙g−1 at 1 A∙g−1 for 1000 cycles. The chitosan-derived carbon reduced the variation in the MoS2 volume upon the insertion/extraction of K+ ions.
A number of works were devoted to the synthesis of Se-doped MoS2 anodes. Kang et al., made doped nanosheets using Sb powder in hydrothermal synthesis [217]. According to the HRTEM analysis, the sheets were separated by 0.96 nm, which ensured an increase in the reversible capacity of Sb-doped MoS2 at 2 A∙g−1 by 1.34 times as compared to the undoped electrode. Gao et al., prepared MoS2−xSex nanotubes coated with a thin carbon layer doped with nitrogen and phosphorus [218]. The best performance in PIBs was found for MoS1.3Se0.7, and an excess amount of selenium substituents caused a decrease in cell cycling stability. An ex-situ XRD study of the reaction products showed that the structure of MoS1.3Se0.7/N,P-carbon nanotubes is reversible during discharge/charge processes. Fan et al., synthesized spherical N-doped carbon-coated MoS1.5Se0.5 structures, which had a narrower band gap as compared to their undoped counterpart [216]. The MoS1.5Se0.5/NC anode showed long-term stability at a high current density of 5 Ag−1 over 500 cycles due to reversible formation of KMo3Se crystals. Fast reaction kinetics and high cycling stability was also demonstrated for the MoS1.6Se0.4/N-carbon anode [221]. The reversibility of the conversion reaction was associated with the inhibition of molybdenum agglomeration due to the presence of selenium. He et al., fabricated vacancy-enriched MoS2−xSex (x = 0.75, 0.5, 0.25) nanoplates by replacing selenium with sulfur at 700 °C [219]. The EXAFS study confirmed the substitution of S for Se at the atomic level (Figure 12a) The MoSSe alloy showed much better performance in PIB than MoS2 and MoSe2 (Figure 12b).
Only two works were related to the synthesis of anodes for PIBs by replacing molybdenum in the MoS2 lattice. Kang et al., reported doping MoS2 with tellurium using Te powder in a hydrothermal process [220]. HRTEM analysis revealed that Te atoms occupy Mo positions with the formation of Te–S bonds, which increases the electrical conductivity of nanomaterial and causes the formation of 1T–1H in-plane heterojunctions. The nanomaterial demonstrated remarkable rate capability, delivering 395 mAh∙g−1 at 5 A∙g−1 with 88% capacity retention over 1000 cycles. A full cell with this anode and K2Fe[Fe(CN)6] cathode had a capacity of 61.7 mAh∙g−1 at a current density of 0.05 A∙g−1 after 50 cycles, demonstrating the practical application of Te-doped MoS2 in PIB. MoS2 co-doped with Fe and Co and inter-overlapped with N,O-doped graphene layers showed the best cycling performance at high current density (Table 5). This anode material was stable at a current density of 20 C and the fast transport of K+ ions was ascribed to a strong surface capacitive effect induced by the incorporation of magnetic atoms in the MoS2 lattice [198].

4.3. Supercapacitors

Unlike batteries, supercapacitors have a long lifespan and high power density, but insufficient energy density [222]. Therefore, efforts were focused on obtaining electrodes with good rate capability. The attractiveness of MoS2 as an electrode material for supercapacitors lies in the large interlayer space for the intercalation and diffusion of electrolyte ions and the variable oxidation states of Mo, which can participate in redox reactions [223]. Charge storage in MoS2 occurs due to the capacitance of the electric double-layer and the pseudocapacitive contribution [224,225]. Reviews on approaches to the synthesis of MoS2 nanomaterials, including their compositions with carbon derivatives, conducting polymers, metal sulfides, etc., for use in supercapacitors are given in [226,227]. The main disadvantage of MoS2 is its low electrical conductivity, which can be enhanced by combining the thermodynamically stable 2H phase and the metallic 1T phase [223]. Such a structural modification improves the performance of MoS2-based electrodes in supercapacitors; however, an analysis of the literature showed that the doping with heteroatoms is the best strategy for this [223,225].
Table 6 presents the highest specific capacitance values recorded for heteroatom-doped MoS2 electrodes using three-electrode systems, as well as power and energy densities obtained during two-electrode cell tests. All works mentioned an increase in the specific capacitance for doped MoS2 as compared to undoped MoS2 synthesized using the same synthetic approach. The reasons for this are an increase in the interlayer space due to the introduction of foreign elements, enhanced conductivity of the electrode material, the appearance of additional adsorption centers, and the contribution from redox-active substituents. An additional gain in the capacitance is achieved when the 1T phase is formed. The positive effect of this structural modification was observed in the case of N-doping [228,229], Mn-doping [230,231], Ni-doping [231,232], and Co-doping [231,233,234]. Simultaneous doping with two foreign elements is also beneficial. For example, the introduction of C, N [187] or Mn, Se [235] into the MoS2 lattice provided high values of specific capacitance, good rate capability, and capacitance retention during long-term cycling.
An analysis of the literature showed that doping with Se, Ni, and Co allows to obtain a very high specific capacitance, approaching [239,246] or even exceeding [231] the theoretically expected value of 1500 F∙g−1 for MoS2. When designing an electrode material, the optimal ratio of Mo to heteroatom is often determined, and Table 6 provides data for the best achieved cases. The optimal content of Ni or Co in the MoS2 is 6–7 at% [231,232,247]. A comparative study of MoS2 nanomaterials doped with Ni, Fe, or Cu in a 6 M KOH electrolyte revealed the pseudocapacitive behavior of the Fe-doped electrode and predominance of electric double-layer capacitance for electrodes doped with Cu and Ni [244]. A study of kinetics of the electrochemical reactions occurring with Ni-doped MoS2 revealed a significant contribution of intercalation process, which is controlled by diffusion [243]. DFT calculations showed that strong interaction between Ni and S results in a distortion of MoS2 lattice and a redistribution of electron density. This interaction also causes a decrease in the number of adjacent layers in the nanomaterial, which increases the rate of insertion and extraction of electrolyte ions.
Heteroatom-doped MoS2 nanomaterials were used to make binder-free electrodes by depositing a sample on a nickel foam [241] or glassy carbon [245]. The procedure of forming a working electrode using electrodeposition looks promising, since it consist of a single step and a deposited film tightly bonded to the electrode surface, which ensures high operation stability of the cell. Flexible supercapacitors were fabricated from carbon cloth directly during hydrothermal synthesis [229,235,248] or by applying a nanomaterial dispersion [233]. Cyclic voltammetry (CV) curves of flexible solid-state supercapacitors showed no significant changes after bending and twisting tests (Figure 13), indicating that they can be used in wearable and foldable electrical devices.
The ideal energy devices should possess a high energy density at high power density. For heteroatom-doped MoS2, the best characteristics are obtained with the substitution of a half of sulfur atoms by selenium [239], and the incorporation of Mn [230,235] and Ni [231,232]. A symmetrical supercapacitor with Mn-doped MoS2 electrodes and a hybrid supercapacitor with a Ni-doped MoS2 positive electrode demonstrated practical applicability by lighting LED indicators [230,231].

5. Conclusions

Substitutional heteroatom doping is a common way to modify the structural, chemical, and electronic properties of a compound. Theory predicts that most elements of the periodic table can be introduced into the MoS2 lattice. The thermodynamically stable hexagonal form MoS2 is a semiconductor, and the introduction of an element from the same group as Mo or S, and, therefore, isovalent to the substituent reduces the band gap. Elements with a smaller (larger) number of the valence electrons induce impurity states near the valence (conductance) band. These states change the type of electrical conductivity of MoS2 and create reaction centers on its surface. Many elements that are theoretically suitable as substituents in MoS2 have been experimentally introduced into its structure. The rest of the elements present a challenge for experimentalists to find the necessary synthetic procedure. The electron density redistribution in MoS2 doped with heteroatoms makes the surface more active for adsorbates, which is an important factor, in particular, for sensor and electrochemical applications. In addition, the use of sources of foreign elements in the synthesis of MoS2 usually leads to the formation of few-layer nanomaterials with an increased interlayer space. This also increases the surface area of MoS2, which is also important for many applications.
This review was devoted to experimental and theoretical studies of MoS2 nanomaterials in chemiresistive sensors and energy storage devices, such as LIBs, SIBs, PIBs, and supercapacitors, and, for the first time, focused exclusively on the effect of substitutional heteroatoms on these properties. Studies of thin layers of MoS2 showed their ability to detect some gases at the ppb level. In DFT calculations, sensitivity is defined as the strength of the sensor–analyte interaction, and calculations showed that this strength increases when MoS2 is doped with heteroatoms. Experimental and theoretical data agree that MoS2 sensors are most sensitive to nitrogen dioxide NO2, and substituents can provide selectivity to one or another gas. Increased interlayer spacing, better electrical conductivity, and surface polarization are key factors for the improved electrochemical performance typically seen with heteroatom-doped MoS2 electrodes. The best characteristics are obtained by doping with Se or magnetic metals Fe, Co, Ni. DFT calculations showed that the presence of these elements in the MoS2 lattice lowers the diffusion barrier for alkali metal ions, resulting in high rate capability of the battery. The rare examples currently available showed that co-doping with different elements is a fruitful approach to improve the performance of MoS2 in sensors and energy storage applications, and developments in this direction could be the next step towards creating novel amazing MoS2-based nanomaterials.
To realize the full potential of heteroatom-doped MoS2 nanomaterials, several challenges need to be solved. The available studies showed that many foreign metals occupy sulfur vacancies, which are often realized in the MoS2 lattice during synthesis. Thus, additional efforts are required to find conditions for replacing molybdenum. It is extremely important to establish clear relationships between the property and the content and position of the substituent in the MoS2 lattice. At present, the composition of such nanomaterials is usually determined by EDS and XPS methods. The former method gives a significant error in the values and, in addition, the signals from some elements overlap. The latter method is surface-sensitive and can give distorted information for nanoparticles thicker than ~5 nm. Therefore, appropriate analytical chemical procedures should be developed and used. Direct visualization of foreign atoms in a crystal is only possible using the HRTEM and HAADF/STEM methods. These methods are local, and close contrasts of some elements make their identification difficult. Moreover, these methods are more suitable for thin films, like monolayers. The environment of the element under study, including the number of neighbors and the distance to them, is determined by the EXAFS method. However, in most cases, models are needed to determine the actual structure. Finding the right models can require extensive DFT calculations.
The performance of heteroatom-doped MoS2 in batteries and supercapacitors is significantly improved by the addition of carbon component. This component also contributes to the accumulation processes, and this factor should be clearly defined for further successful design of nanomaterials. MoS2-based nanomaterials have great prospects as chemiresistive gas sensors and electrodes in electrochemical energy storage devices. Due to their excellent mechanical and chemical stability, they can be used in wearable electronics. It has been shown that supercapacitors with MoS2 electrodes retain their characteristics when bent and twisted. This property is also desirable for gas sensors that can be used in e-skin for on-site environmental monitoring. Substitutional doping with heteroatoms often improves the characteristics of MoS2, and this direction will undoubtedly be actively developed in the near future.

Author Contributions

Conceptualization, L.G.B.; writing—original draft preparation, L.G.B., G.I.S. and A.D.F.; writing—review and editing, project administration, L.G.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation (grant No. 23-73-00048).

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Dickinson, R.G.; Pauling, L. The crystal structure of molybdenite. J. Am. Chem. Soc. 1923, 45, 1466–1471. [Google Scholar] [CrossRef]
  2. Song, I.; Park, C.; Choi, H.C. Synthesis and properties of molybdenum disulphide: From bulk to atomic layers. RSC Adv. 2015, 5, 7495–7514. [Google Scholar] [CrossRef] [Green Version]
  3. Frondel, J.; Wickman, F. Molybdenite polytypes in theory and occurrence. II. Some naturally-occurring polytypes of molybdenite. Am. Miner. 1970, 55, 1857–1875. [Google Scholar]
  4. Winer, W.O. Molybdenum disulfide as a lubricant: A review of the fundamental knowledge. Wear 1967, 10, 422–452. [Google Scholar] [CrossRef] [Green Version]
  5. Furimsky, E. Role of MoS2 and WS2 in Hydrodesulfuritation. Catal. Rev. 1980, 22, 371–400. [Google Scholar] [CrossRef]
  6. Benavente, E. Intercalation chemistry of molybdenum disulfide. Coord. Chem. Rev. 2002, 224, 87–109. [Google Scholar] [CrossRef]
  7. Laman, F.C.; Stiles, J.A.R.; Shank, R.J.; Brandt, K. Rate limiting mechanisms in lithium—Molybdenum disulfide batteries. J. Power Sources 1985, 14, 201–207. [Google Scholar] [CrossRef]
  8. Imanishi, N.; Kanamura, K.; Takehara, Z. Synthesis of MoS2 Thin Film by Chemical Vapor Deposition Method and Discharge Characteristics as a Cathode of the Lithium Secondary Battery. J. Electrochem. Soc. 1992, 139, 2082–2087. [Google Scholar] [CrossRef]
  9. Reddy, M.V.; Mauger, A.; Julien, C.M.; Paolella, A.; Zaghib, K. Brief History of Early Lithium-Battery Development. Materials 2020, 13, 1884. [Google Scholar] [CrossRef] [Green Version]
  10. Stephenson, T.; Li, Z.; Olsen, B.; Mitlin, D. Lithium ion battery applications of molybdenum disulfide (MoS2) nanocomposites. Energy Environ. Sci. 2014, 7, 209–231. [Google Scholar] [CrossRef]
  11. Frindt, R.F. Single Crystals of MoS2 Several Molecular Layers Thick. J. Appl. Phys. 1966, 37, 1928–1929. [Google Scholar] [CrossRef]
  12. Bodenmann, A.K.; MacDonald, A.H. Graphene: Exploring carbon flatland. Phys. Today 2007, 60, 35–41. [Google Scholar] [CrossRef]
  13. Novoselov, K.S.; Jiang, D.; Schedin, F.; Booth, T.J.; Khotkevich, V.V.; Morozov, S.V.; Geim, A.K. Two-dimensional atomic crystals. Proc. Natl. Acad. Sci. USA 2005, 102, 10451–10453. [Google Scholar] [CrossRef] [PubMed]
  14. Rao, C.N.R.; Maitra, U.; Waghmare, U.V. Extraordinary attributes of 2-dimensional MoS2 nanosheets. Chem. Phys. Lett. 2014, 609, 172–183. [Google Scholar] [CrossRef]
  15. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically ThinMoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [Green Version]
  16. Yazyev, O.V.; Kis, A. MoS2 and semiconductors in the flatland. Mater. Today 2015, 18, 20–30. [Google Scholar] [CrossRef]
  17. Ganatra, R.; Zhang, Q. Few-Layer MoS2: A Promising Layered Semiconductor. ACS Nano 2014, 8, 4074–4099. [Google Scholar] [CrossRef]
  18. Amaral, L.O.; Daniel-da-Silva, A.L. MoS2 and MoS2 Nanocomposites for Adsorption and Photodegradation of Water Pollutants: A Review. Molecules 2022, 27, 6782. [Google Scholar] [CrossRef]
  19. Liu, C.; Wang, Q.; Jia, F.; Song, S. Adsorption of heavy metals on molybdenum disulfide in water: A critical review. J. Mol. Liq. 2019, 292, 111390. [Google Scholar] [CrossRef]
  20. Singh Rana, D.; Thakur, N.; Singh, D.; Sonia, P. Molybdenum and tungsten disulfide based nanocomposites as chemical sensor: A review. Mater. Today Proc. 2022, 62, 2755–2761. [Google Scholar] [CrossRef]
  21. Zhang, G.; Liu, H.; Qu, J.; Li, J. Two-dimensional layered MoS2: Rational design, properties and electrochemical applications. Energy Environ. Sci. 2016, 9, 1190–1209. [Google Scholar] [CrossRef]
  22. Sinha, A.; Dhanjai; Tan, B.; Huang, Y.; Zhao, H.; Dang, X.; Chen, J.; Jain, R. MoS2 nanostructures for electrochemical sensing of multidisciplinary targets: A review. TrAC Trends Anal. Chem. 2018, 102, 75–90. [Google Scholar] [CrossRef]
  23. Xu, D.; Duan, L.; Yan, S.; Wang, Y.; Cao, K.; Wang, W.; Xu, H.; Wang, Y.; Hu, L.; Gao, L. Monolayer MoS2-Based Flexible and Highly Sensitive Pressure Sensor with Wide Sensing Range. Micromachines 2022, 13, 660. [Google Scholar] [CrossRef]
  24. Varghese, S.; Varghese, S.; Swaminathan, S.; Singh, K.; Mittal, V. Two-Dimensional Materials for Sensing: Graphene and Beyond. Electronics 2015, 4, 651–687. [Google Scholar] [CrossRef] [Green Version]
  25. Yang, S.; Jiang, C.; Wei, S. Gas sensing in 2D materials. Appl. Phys. Rev. 2017, 4, 021304. [Google Scholar] [CrossRef]
  26. Samy, O.; El Moutaouakil, A. A Review on MoS2 Energy Applications: Recent Developments and Challenges. Energies 2021, 14, 4586. [Google Scholar] [CrossRef]
  27. He, Z.; Que, W. Molybdenum disulfide nanomaterials: Structures, properties, synthesis and recent progress on hydrogen evolution reaction. Appl. Mater. Today 2016, 3, 23–56. [Google Scholar] [CrossRef]
  28. Ding, Q.; Song, B.; Xu, P.; Jin, S. Efficient Electrocatalytic and Photoelectrochemical Hydrogen Generation Using MoS2 and Related Compounds. Chem 2016, 1, 699–726. [Google Scholar] [CrossRef] [Green Version]
  29. Vazirisereshk, M.R.; Martini, A.; Strubbe, D.A.; Baykara, M.Z. Solid Lubrication with MoS2: A Review. Lubricants 2019, 7, 57. [Google Scholar] [CrossRef] [Green Version]
  30. Balasubramaniam, B.; Singh, N.; Kar, P.; Tyagi, A.; Prakash, J.; Gupta, R.K. Engineering of transition metal dichalcogenide-based 2D nanomaterials through doping for environmental applications. Mol. Syst. Des. Eng. 2019, 4, 804–827. [Google Scholar] [CrossRef]
  31. Liu, M.; Zhang, C.; Han, A.; Wang, L.; Sun, Y.; Zhu, C.; Li, R.; Ye, S. Modulation of morphology and electronic structure on MoS2-based electrocatalysts for water splitting. Nano Res. 2022, 15, 6862–6887. [Google Scholar] [CrossRef]
  32. Agrawal, A.V.; Kumar, N.; Kumar, M. Strategy and Future Prospects to Develop Room-Temperature-Recoverable NO2 Gas Sensor Based on Two-Dimensional Molybdenum Disulfide. Nano-Micro Lett. 2021, 13, 38. [Google Scholar] [CrossRef] [PubMed]
  33. Walton, A.S.; Lauritsen, J.V.; Topsøe, H.; Besenbacher, F. MoS2 nanoparticle morphologies in hydrodesulfurization catalysis studied by scanning tunneling microscopy. J. Catal. 2013, 308, 306–318. [Google Scholar] [CrossRef]
  34. Tedstone, A.A.; Lewis, D.J.; O’Brien, P. Synthesis, Properties, and Applications of Transition Metal-Doped Layered Transition Metal Dichalcogenides. Chem. Mater. 2016, 28, 1965–1974. [Google Scholar] [CrossRef] [Green Version]
  35. Loh, L.; Zhang, Z.; Bosman, M.; Eda, G. Substitutional doping in 2D transition metal dichalcogenides. Nano Res. 2021, 14, 1668–1681. [Google Scholar] [CrossRef]
  36. Sovizi, S.; Szoszkiewicz, R. Single atom doping in 2D layered MoS2 from a periodic table perspective. Surf. Sci. Rep. 2022, 77, 100567. [Google Scholar] [CrossRef]
  37. Yee, K.A.; Hughbanks, T. Utility of semilocalized bonding schemes in extended systems: Three-center metal-metal bonding in molybdenum sulfide (MoS2), niobium tantalum sulfide bronze (Hx(Nb,Ta)S2), and zirconium sulfide (ZrS). Inorg. Chem. 1991, 30, 2321–2328. [Google Scholar] [CrossRef]
  38. Ding, Y.; Wang, Y.; Ni, J.; Shi, L.; Shi, S.; Tang, W. First principles study of structural, vibrational and electronic properties of graphene-like MX2 (M = Mo, Nb, W, Ta; X = S, Se, Te) monolayers. Phys. B Condens. Matter 2011, 406, 2254–2260. [Google Scholar] [CrossRef]
  39. Hu, A.-M.; Wang, L.; Meng, B.; Xiao, W.-Z. Ab initio study of magnetism in nonmagnetic metal substituted monolayer MoS2. Solid State Commun. 2015, 220, 67–71. [Google Scholar] [CrossRef]
  40. Abidi, N.; Bonduelle-Skrzypczak, A.; Steinmann, S.N. How to dope the basal plane of 2H-MoS2 to boost the hydrogen evolution reaction? Electrochim. Acta 2023, 439, 141653. [Google Scholar] [CrossRef]
  41. Menezes, M.G.; Ullah, S. Unveiling the multi-level structure of midgap states in Sb-doped MoX2 (X = S, Se, Te) monolayers. Phys. Rev. B 2021, 104, 125438. [Google Scholar] [CrossRef]
  42. Sahoo, A.R.; Jaya, S.M. Structural stability and electronic structure of Cr substituted MoS2: An ab-initio study. AIP Conf. Proc. 2019, 2115, 030393. [Google Scholar]
  43. Andriambelaza, N.F.; Mapasha, R.E.; Chetty, N. First-principles studies of chromium line-ordered alloys in a molybdenum disulfide monolayer. J. Phys. Condens. Matter 2017, 29, 325504. [Google Scholar] [CrossRef]
  44. Xi, J.; Zhao, T.; Wang, D.; Shuai, Z. Tunable Electronic Properties of Two-Dimensional Transition Metal Dichalcogenide Alloys: A First-Principles Prediction. J. Phys. Chem. Lett. 2014, 5, 285–291. [Google Scholar] [CrossRef]
  45. Yang, J.-H.; Yakobson, B.I. Unusual Negative Formation Enthalpies and Atomic Ordering in Isovalent Alloys of Transition Metal Dichalcogenide Monolayers. Chem. Mater. 2018, 30, 1547–1555. [Google Scholar] [CrossRef] [Green Version]
  46. Atthapak, C.; Ektarawong, A.; Pakornchote, T.; Alling, B.; Bovornratanaraks, T. Effect of atomic configuration and spin–orbit coupling on thermodynamic stability and electronic bandgap of monolayer 2H-Mo1−xWxS2 solid solutions. Phys. Chem. Chem. Phys. 2021, 23, 13535–13543. [Google Scholar] [CrossRef]
  47. Gan, L.-Y.; Zhang, Q.; Zhao, Y.-J.; Cheng, Y.; Schwingenschlögl, U. Order-disorder phase transitions in the two-dimensional semiconducting transition metal dichalcogenide alloys Mo1−xWxX2 (X = S, Se and Te). Sci. Rep. 2014, 4, 6691. [Google Scholar] [CrossRef] [Green Version]
  48. Gao, Y.; Liu, J.; Zhang, X.; Lu, G. Unraveling Structural and Optical Properties of Two-Dimensional MoxW1– xS2 Alloys. J. Phys. Chem. C 2021, 125, 774–781. [Google Scholar] [CrossRef]
  49. Kutana, A.; Penev, E.S.; Yakobson, B.I. Engineering electronic properties of layered transition-metal dichalcogenide compounds through alloying. Nanoscale 2014, 6, 5820–5825. [Google Scholar] [CrossRef]
  50. Meng, D.; Li, M.; Long, D.; Cheng, Y.; Ye, P.; Fu, W.; E, W.; Luo, W.; He, Y. Electronic structure and dynamic properties of two-dimensional WxMo1−xS2 ternary alloys from first-principles calculations. Comput. Mater. Sci. 2020, 182, 109797. [Google Scholar] [CrossRef]
  51. Tsuppayakorn-aek, P.; Pungtrakool, W.; Pinsook, U.; Bovornratanaraks, T. The minimal supercells approach for ab-initio calculation in 2D alloying transition metal dichalcoginides with special quasi-random structure. Mater. Res. Express 2020, 7, 086502. [Google Scholar] [CrossRef]
  52. Tan, W.; Wei, Z.; Liu, X.; Liu, J.; Fang, X.; Fang, D.; Wang, X.; Wang, D.; Tang, J.; Fan, X. Ordered and Disordered Phases in Mo1−xWxS2 Monolayer. Sci. Rep. 2017, 7, 15124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Cheng, Y.C.; Zhu, Z.Y.; Mi, W.B.; Guo, Z.B.; Schwingenschlögl, U. Prediction of two-dimensional diluted magnetic semiconductors: Doped monolayer MoS2 systems. Phys. Rev. B 2013, 87, 100401. [Google Scholar] [CrossRef] [Green Version]
  54. Lu, S.-C.; Leburton, J.-P. Electronic structures of defects and magnetic impurities in MoS2 monolayers. Nanoscale Res. Lett. 2014, 9, 676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Chanana, A.; Mahapatra, S. Theoretical Insights to Niobium-Doped Monolayer MoS2 –Gold Contact. IEEE Trans. Electron Devices 2015, 62, 2346–2351. [Google Scholar] [CrossRef]
  56. Mishra, R.; Zhou, W.; Pennycook, S.J.; Pantelides, S.T.; Idrobo, J.-C. Long-range ferromagnetic ordering in manganese-doped two-dimensional dichalcogenides. Phys. Rev. B 2013, 88, 144409. [Google Scholar] [CrossRef]
  57. Ramasubramaniam, A.; Naveh, D. Mn-doped monolayer MoS2: An atomically thin dilute magnetic semiconductor. Phys. Rev. B 2013, 87, 195201. [Google Scholar] [CrossRef]
  58. Yue, Y.; Jiang, C.; Han, Y.; Wang, M.; Ren, J.; Wu, Y. Magnetic anisotropies of Mn-, Fe-, and Co-doped monolayer MoS2. J. Magn. Magn. Mater. 2020, 496, 165929. [Google Scholar] [CrossRef]
  59. Dolui, K.; Rungger, I.; Das Pemmaraju, C.; Sanvito, S. Possible doping strategies for MoS2 monolayers: An ab initio study. Phys. Rev. B 2013, 88, 075420. [Google Scholar] [CrossRef] [Green Version]
  60. Tan, A.M.Z.; Freysoldt, C.; Hennig, R.G. First-principles investigation of charged dopants and dopant-vacancy defect complexes in monolayer MoS2. Phys. Rev. Mater. 2020, 4, 114002. [Google Scholar] [CrossRef]
  61. Ivanovskaya, V.V.; Zobelli, A.; Gloter, A.; Brun, N.; Serin, V.; Colliex, C. Ab initio study of bilateral doping within the MoS2-NbS2 system. Phys. Rev. B 2008, 78, 134104. [Google Scholar] [CrossRef] [Green Version]
  62. Yoshimura, A.; Koratkar, N.; Meunier, V. Substitutional transition metal doping in MoS2: A first-principles study. Nano Express 2020, 1, 010008. [Google Scholar] [CrossRef]
  63. Shu, H.; Luo, P.; Liang, P.; Cao, D.; Chen, X. Layer-Dependent Dopant Stability and Magnetic Exchange Coupling of Iron-Doped MoS2 Nanosheets. ACS Appl. Mater. Interfaces 2015, 7, 7534–7541. [Google Scholar] [CrossRef]
  64. Guerrero, E.; Karkee, R.; Strubbe, D.A. Phase Stability and Raman/IR Signatures of Ni-Doped MoS2 from Density Functional Theory Studies. J. Phys. Chem. C 2021, 125, 13401–13412. [Google Scholar] [CrossRef]
  65. Han, X.; Benkraouda, M.; Amrane, N. S Vacancy Engineered Electronic and Optoelectronic Properties of Ni-Doped MoS2 Monolayer: A Hybrid Functional Study. J. Electron. Mater. 2020, 49, 3234–3241. [Google Scholar] [CrossRef]
  66. Liu, G.; Robertson, A.W.; Li, M.M.-J.; Kuo, W.C.H.; Darby, M.T.; Muhieddine, M.H.; Lin, Y.-C.; Suenaga, K.; Stamatakis, M.; Warner, J.H.; et al. MoS2 monolayer catalyst doped with isolated Co atoms for the hydrodeoxygenation reaction. Nat. Chem. 2017, 9, 810–816. [Google Scholar] [CrossRef]
  67. Kang, J.; Tongay, S.; Li, J.; Wu, J. Monolayer semiconducting transition metal dichalcogenide alloys: Stability and band bowing. J. Appl. Phys. 2013, 113, 143703. [Google Scholar] [CrossRef]
  68. Komsa, H.-P.; Krasheninnikov, A.V. Two-Dimensional Transition Metal Dichalcogenide Alloys: Stability and Electronic Properties. J. Phys. Chem. Lett. 2012, 3, 3652–3656. [Google Scholar] [CrossRef] [PubMed]
  69. Kuc, A.; Heine, T. On the Stability and Electronic Structure of Transition-Metal Dichalcogenide Monolayer Alloys Mo1−xXxS2−ySey with X = W, Nb. Electronics 2015, 5, 1. [Google Scholar] [CrossRef] [Green Version]
  70. Vallinayagam, M.; Posselt, M.; Chandra, S. Electronic structure and thermoelectric properties of Mo-based dichalcogenide monolayers locally and randomly modified by substitutional atoms. RSC Adv. 2020, 10, 43035–43044. [Google Scholar] [CrossRef]
  71. Andriambelaza, N.F.; Mapasha, R.E.; Chetty, N. First-principles studies of Te line-ordered alloys in a MoS2 monolayer. Phys. B Condens. Matter 2018, 535, 162–166. [Google Scholar] [CrossRef] [Green Version]
  72. Andriambelaza, N.F.; Mapasha, R.E.; Chetty, N. Band gap engineering of a MoS2 monolayer through oxygen alloying: An ab initio study. Nanotechnology 2018, 29, 505701. [Google Scholar] [CrossRef]
  73. Enujekwu, F.M.; Zhang, Y.; Ezeh, C.I.; Zhao, H.; Xu, M.; Besley, E.; George, M.W.; Besley, N.A.; Do, H.; Wu, T. N-doping enabled defect-engineering of MoS2 for enhanced and selective adsorption of CO2: A DFT approach. Appl. Surf. Sci. 2021, 542, 148556. [Google Scholar] [CrossRef]
  74. Nie, S.; Li, Z. Strain Effect on the Magnetism of N-Doped Molybdenum Disulfide. Phys. Status Solidi 2019, 256, 1900110. [Google Scholar] [CrossRef]
  75. Xiao, W.; Liu, P.; Zhang, J.; Song, W.; Feng, Y.P.; Gao, D.; Ding, J. Dual-Functional N Dopants in Edges and Basal Plane of MoS2 Nanosheets Toward Efficient and Durable Hydrogen Evolution. Adv. Energy Mater. 2017, 7, 1602086. [Google Scholar] [CrossRef]
  76. Komsa, H.-P.; Kotakoski, J.; Kurasch, S.; Lehtinen, O.; Kaiser, U.; Krasheninnikov, A.V. Two-Dimensional Transition Metal Dichalcogenides under Electron Irradiation: Defect Production and Doping. Phys. Rev. Lett. 2012, 109, 035503. [Google Scholar] [CrossRef] [PubMed]
  77. Dong, F.; Zhang, M.; Xu, X.; Pan, J.; Zhu, L.; Hu, J. Orbital Modulation with P Doping Improves Acid and Alkaline Hydrogen Evolution Reaction of MoS2. Nanomaterials 2022, 12, 4273. [Google Scholar] [CrossRef]
  78. Kotsun, A.A.; Alekseev, V.A.; Stolyarova, S.G.; Makarova, A.A.; Grebenkina, M.A.; Zubareva, A.P.; Okotrub, A.V.; Bulusheva, L.G. Effect of molybdenum disulfide doping with substitutional nitrogen and sulfur vacancies on lithium intercalation. J. Alloys Compd. 2023, 947, 169689. [Google Scholar] [CrossRef]
  79. Miralrio, A.; Rangel, E.; Castro, M. Activation of MoS2 monolayers by substitutional copper and silver atoms embedded in sulfur vacancies: A theoretical study. Appl. Surf. Sci. 2019, 481, 611–624. [Google Scholar] [CrossRef]
  80. Xiao, J.; Wang, J.; Xue, Z.; Zhang, T.; Wang, J.; Li, Q. The study of oxygen reduction reaction on Ge-doped MoS2 monolayer based on first principle. Int. J. Energy Res. 2021, 45, 13748–13759. [Google Scholar] [CrossRef]
  81. Susarla, S.; Kutana, A.; Hachtel, J.A.; Kochat, V.; Apte, A.; Vajtai, R.; Idrobo, J.C.; Yakobson, B.I.; Tiwary, C.S.; Ajayan, P.M. Quaternary 2D Transition Metal Dichalcogenides (TMDs) with Tunable Bandgap. Adv. Mater. 2017, 29, 1702457. [Google Scholar] [CrossRef]
  82. Zhou, J.; Lin, J.; Sims, H.; Jiang, C.; Cong, C.; Brehm, J.A.; Zhang, Z.; Niu, L.; Chen, Y.; Zhou, Y.; et al. Synthesis of Co-Doped MoS2 Monolayers with Enhanced Valley Splitting. Adv. Mater. 2020, 32, 1906536. [Google Scholar] [CrossRef] [PubMed]
  83. Robertson, A.W.; Lin, Y.-C.; Wang, S.; Sawada, H.; Allen, C.S.; Chen, Q.; Lee, S.; Lee, G.-D.; Lee, J.; Han, S.; et al. Atomic Structure and Spectroscopy of Single Metal (Cr, V) Substitutional Dopants in Monolayer MoS2. ACS Nano 2016, 10, 10227–10236. [Google Scholar] [CrossRef]
  84. Yin, G.; Zhu, D.; Lv, D.; Hashemi, A.; Fei, Z.; Lin, F.; Krasheninnikov, A.V.; Zhang, Z.; Komsa, H.-P.; Jin, C. Hydrogen-assisted post-growth substitution of tellurium into molybdenum disulfide monolayers with tunable compositions. Nanotechnology 2018, 29, 145603. [Google Scholar] [CrossRef] [PubMed]
  85. Sharona, H.; Vishal, B.; Bhat, U.; Paul, A.; Mukherjee, A.; Sarma, S.C.; Peter, S.C.; Datta, R. Rich diversity of crystallographic phase formation in 2D RexMo1–xS2 (x < 0.5) alloy. J. Appl. Phys. 2019, 126, 224302. [Google Scholar] [CrossRef] [Green Version]
  86. Zhang, K.; Bersch, B.M.; Joshi, J.; Addou, R.; Cormier, C.R.; Zhang, C.; Xu, K.; Briggs, N.C.; Wang, K.; Subramanian, S.; et al. Tuning the Electronic and Photonic Properties of Monolayer MoS2 via In Situ Rhenium Substitutional Doping. Adv. Funct. Mater. 2018, 28, 1706950. [Google Scholar] [CrossRef]
  87. Gong, Y.; Liu, Z.; Lupini, A.R.; Shi, G.; Lin, J.; Najmaei, S.; Lin, Z.; Elías, A.L.; Berkdemir, A.; You, G.; et al. Band Gap Engineering and Layer-by-Layer Mapping of Selenium-Doped Molybdenum Disulfide. Nano Lett. 2014, 14, 442–449. [Google Scholar] [CrossRef] [PubMed]
  88. Song, J.-G.; Ryu, G.H.; Lee, S.J.; Sim, S.; Lee, C.W.; Choi, T.; Jung, H.; Kim, Y.; Lee, Z.; Myoung, J.-M.; et al. Controllable synthesis of molybdenum tungsten disulfide alloy for vertically composition-controlled multilayer. Nat. Commun. 2015, 6, 7817. [Google Scholar] [CrossRef] [Green Version]
  89. Xia, X.; Loh, S.M.; Viner, J.; Teutsch, N.C.; Graham, A.J.; Kandyba, V.; Barinov, A.; Sanchez, A.M.; Smith, D.C.; Hine, N.D.M.; et al. Atomic and electronic structure of two-dimensional Mo(1−x)WxS2 alloys. J. Phys. Mater. 2021, 4, 025004. [Google Scholar] [CrossRef]
  90. Cai, Z.; Shen, T.; Zhu, Q.; Feng, S.; Yu, Q.; Liu, J.; Tang, L.; Zhao, Y.; Wang, J.; Liu, B.; et al. Dual-Additive Assisted Chemical Vapor Deposition for the Growth of Mn-Doped 2D MoS2 with Tunable Electronic Properties. Small 2019, 16, 1903181. [Google Scholar] [CrossRef]
  91. Cao, B.; Ma, S.; Wang, W.; Tang, X.; Wang, D.; Shen, W.; Qiu, B.; Xu, B.; Li, G. Charge Redistribution in Mg-Doped p-Type MoS2 /GaN Photodetectors. J. Phys. Chem. C 2022, 126, 18893–18899. [Google Scholar] [CrossRef]
  92. Zhong, M.; Shen, C.; Huang, L.; Deng, H.-X.; Shen, G.; Zheng, H.; Wei, Z.; Li, J. Electronic structure and exciton shifts in Sb-doped MoS2 monolayer. npj 2D Mater. Appl. 2019, 3, 1. [Google Scholar] [CrossRef] [Green Version]
  93. Wang, J.; Fang, W.; Hu, Y.; Zhang, Y.; Dang, J.; Wu, Y.; Chen, B.; Zhao, H.; Li, Z. Single atom Ru doping 2H-MoS2 as highly efficient hydrogen evolution reaction electrocatalyst in a wide pH range. Appl. Catal. B Environ. 2021, 298, 120490. [Google Scholar] [CrossRef]
  94. Mattila, S.; Leiro, J.A.; Heinonen, M.; Laiho, T. Core level spectroscopy of MoS2. Surf. Sci. 2006, 600, 5168–5175. [Google Scholar] [CrossRef]
  95. Baker, M.; Gilmore, R.; Lenardi, C.; Gissler, W. XPS investigation of preferential sputtering of S from MoS2 and determination of MoSx stoichiometry from Mo and S peak positions. Appl. Surf. Sci. 1999, 150, 255–262. [Google Scholar] [CrossRef]
  96. Bulusheva, L.G.; Okotrub, A.V.; Yashina, L.V.; Velasco-Velez, J.J.; Usachov, D.Y.; Vyalikh, D.V. X-ray photoelectron spectroscopy study of the interaction of lithium with graphene. Phys. Sci. Rev. 2018, 3, 20180042. [Google Scholar] [CrossRef]
  97. Lv, K.; Suo, W.; Shao, M.; Zhu, Y.; Wang, X.; Feng, J.; Fang, M.; Zhu, Y. Nitrogen doped MoS2 and nitrogen doped carbon dots composite catalyst for electroreduction CO2 to CO with high Faradaic efficiency. Nano Energy 2019, 63, 103834. [Google Scholar] [CrossRef]
  98. Li, M.; Yao, J.; Wu, X.; Zhang, S.; Xing, B.; Niu, X.; Yan, X.; Yu, Y.; Liu, Y.; Wang, Y. P-type Doping in Large-Area Monolayer MoS2 by Chemical Vapor Deposition. ACS Appl. Mater. Interfaces 2020, 12, 6276–6282. [Google Scholar] [CrossRef]
  99. Ledneva, A.Y.; Dalmatova, S.A.; Fedorenko, A.D.; Asanov, I.P.; Enyashin, A.N.; Mazalov, L.N.; Fedorov, V.E. An Xps Study of Solid Solutions Mo1–XNbxS2 (0 < x < 0.15). J. Struct. Chem. 2018, 59, 1833–1840. [Google Scholar] [CrossRef]
  100. Lee, C.; Yan, H.; Brus, L.E.; Heinz, T.F.; Hone, J.; Ryu, S. Anomalous Lattice Vibrations of Single- and Few-Layer MoS2. ACS Nano 2010, 4, 2695–2700. [Google Scholar] [CrossRef] [Green Version]
  101. Sharma, R.; Pandey, J.; Sahoo, K.R.; Rana, K.S.; Biroju, R.K.; Theis, W.; Soni, A.; Narayanan, T.N. Spectroscopic correlation of chalcogen defects in atomically thin MoS2(1−x)Se2x alloys. J. Phys. Mater. 2020, 3, 045001. [Google Scholar] [CrossRef]
  102. Li, H.; Duan, X.; Wu, X.; Zhuang, X.; Zhou, H.; Zhang, Q.; Zhu, X.; Hu, W.; Ren, P.; Guo, P.; et al. Growth of Alloy MoS2 xSe2(1– x) Nanosheets with Fully Tunable Chemical Compositions and Optical Properties. J. Am. Chem. Soc. 2014, 136, 3756–3759. [Google Scholar] [CrossRef] [PubMed]
  103. Schulpen, J.J.P.M.; Verheijen, M.A.; Kessels, W.M.M.; Vandalon, V.; Bol, A.A. Controlling transition metal atomic ordering in two-dimensional Mo1−xWxS2 alloys. 2D Mater. 2022, 9, 025016. [Google Scholar] [CrossRef]
  104. Hu, X.; Hemmat, Z.; Majidi, L.; Cavin, J.; Mishra, R.; Salehi-Khojin, A.; Ogut, S.; Klie, R.F. Controlling Nanoscale Thermal Expansion of Monolayer Transition Metal Dichalcogenides by Alloy Engineering. Small 2020, 16, 1905892. [Google Scholar] [CrossRef] [PubMed]
  105. Hanafusa, S.; Hashimoto, Y.; Ishibashi, K.; Hibino, Y.; Yamazaki, K.; Yokogawa, R.; Wakabayashi, H.; Ogura, A. Evaluation of Mo(1-x)Wxs Alloy Fabricated By Combinatorial Film Deposition. ECS Trans. 2021, 104, 21–28. [Google Scholar] [CrossRef]
  106. Song, Z.; Li, J.; Davis, K.D.; Li, X.; Zhang, J.; Zhang, L.; Sun, X. Emerging Applications of Synchrotron Radiation X-Ray Techniques in Single Atomic Catalysts. Small Methods 2022, 6, 2201078. [Google Scholar] [CrossRef]
  107. Zhang, H.; Yu, L.; Chen, T.; Zhou, W.; Lou, X.W.D. Surface Modulation of Hierarchical MoS2 Nanosheets by Ni Single Atoms for Enhanced Electrocatalytic Hydrogen Evolution. Adv. Funct. Mater. 2018, 28, 1807086. [Google Scholar] [CrossRef]
  108. Suh, J.; Park, T.-E.; Lin, D.-Y.; Fu, D.; Park, J.; Jung, H.J.; Chen, Y.; Ko, C.; Jang, C.; Sun, Y.; et al. Doping against the Native Propensity of MoS2: Degenerate Hole Doping by Cation Substitution. Nano Lett. 2014, 14, 6976–6982. [Google Scholar] [CrossRef]
  109. Kumar, S.; Meng, G.; Mishra, P.; Tripathi, N.; Bannov, A.G. A systematic review on 2D MoS2 for nitrogen dioxide (NO2) sensing at room temperature. Mater. Today Commun. 2023, 34, 105045. [Google Scholar] [CrossRef]
  110. Li, H.; Yin, Z.; He, Q.; Li, H.; Huang, X.; Lu, G.; Fam, D.W.H.; Tok, A.I.Y.; Zhang, Q.; Zhang, H. Fabrication of Single- and Multilayer MoS2 Film-Based Field-Effect Transistors for Sensing NO at Room Temperature. Small 2012, 8, 63–67. [Google Scholar] [CrossRef]
  111. Wang, Z.; Zhang, Y.; Ren, Y.; Wang, M.; Zhang, Z.; Zhao, W.; Yan, J.; Zhai, C.; Yun, J. NO gas adsorption properties of MoS2 from monolayer to trilayer: A first-principles study. Mater. Res. Express 2021, 8, 015024. [Google Scholar] [CrossRef]
  112. Late, D.J.; Huang, Y.-K.; Liu, B.; Acharya, J.; Shirodkar, S.N.; Luo, J.; Yan, A.; Charles, D.; Waghmare, U.V.; Dravid, V.P.; et al. Sensing Behavior of Atomically Thin-Layered MoS2 Transistors. ACS Nano 2013, 7, 4879–4891. [Google Scholar] [CrossRef]
  113. Yao, Y.; Tolentino, L.; Yang, Z.; Song, X.; Zhang, W.; Chen, Y.; Wong, C. High-Concentration Aqueous Dispersions of MoS2. Adv. Funct. Mater. 2013, 23, 3577–3583. [Google Scholar] [CrossRef]
  114. Lee, K.; Gatensby, R.; McEvoy, N.; Hallam, T.; Duesberg, G.S. High-Performance Sensors Based on Molybdenum Disulfide Thin Films. Adv. Mater. 2013, 25, 6699–6702. [Google Scholar] [CrossRef] [Green Version]
  115. Perkins, F.K.; Friedman, A.L.; Cobas, E.; Campbell, P.M.; Jernigan, G.G.; Jonker, B.T. Chemical Vapor Sensing with Monolayer MoS2. Nano Lett. 2013, 13, 668–673. [Google Scholar] [CrossRef]
  116. Liu, B.; Chen, L.; Liu, G.; Abbas, A.N.; Fathi, M.; Zhou, C. High-Performance Chemical Sensing Using Schottky-Contacted Chemical Vapor Deposition Grown Monolayer MoS2 Transistors. ACS Nano 2014, 8, 5304–5314. [Google Scholar] [CrossRef]
  117. Patel, C.; Singh, R.; Dubey, M.; Pandey, S.K.; Upadhyay, S.N.; Kumar, V.; Sriram, S.; Than Htay, M.; Pakhira, S.; Atuchin, V.V.; et al. Large and Uniform Single Crystals of MoS2 Monolayers for ppb-Level NO2 Sensing. ACS Appl. Nano Mater. 2022, 5, 9415–9426. [Google Scholar] [CrossRef]
  118. Bermudez, V.M. Computational Study of the Adsorption of NO2 on Monolayer MoS2. J. Phys. Chem. C 2020, 124, 15275–15284. [Google Scholar] [CrossRef]
  119. Koroteev, V.O.; Bulushev, D.A.; Chuvilin, A.L.; Okotrub, A.V.; Bulusheva, L.G. Nanometer-Sized MoS2 Clusters on Graphene Flakes for Catalytic Formic Acid Decomposition. ACS Catal. 2014, 4, 3950–3956. [Google Scholar] [CrossRef]
  120. Bulusheva, L.G.; Koroteev, V.O.; Stolyarova, S.G.; Chuvilin, A.L.; Plyusnin, P.E.; Shubin, Y.V.; Vilkov, O.Y.; Chen, X.; Song, H.; Okotrub, A.V. Effect of in-plane size of MoS2 nanoparticles grown over multilayer graphene on the electrochemical performance of anodes in Li-ion batteries. Electrochim. Acta 2018, 283, 45–53. [Google Scholar] [CrossRef]
  121. Cho, S.-Y.; Kim, S.J.; Lee, Y.; Kim, J.-S.; Jung, W.-B.; Yoo, H.-W.; Kim, J.; Jung, H.-T. Highly Enhanced Gas Adsorption Properties in Vertically Aligned MoS2 Layers. ACS Nano 2015, 9, 9314–9321. [Google Scholar] [CrossRef]
  122. Annanouch, F.E.; Alagh, A.; Umek, P.; Casanova-Chafer, J.; Bittencourt, C.; Llobet, E. Controlled growth of 3D assemblies of edge enriched multilayer MoS2 nanosheets for dually selective NH3 and NO2 gas sensors. J. Mater. Chem. C 2022, 10, 11027–11039. [Google Scholar] [CrossRef]
  123. Kumar, R.; Goel, N.; Kumar, M. High performance NO2 sensor using MoS2 nanowires network. Appl. Phys. Lett. 2018, 112, 053502. [Google Scholar] [CrossRef]
  124. Rezende, N.P.; Cadore, A.R.; Gadelha, A.C.; Pereira, C.L.; Ornelas, V.; Watanabe, K.; Taniguchi, T.; Ferlauto, A.S.; Malachias, A.; Campos, L.C.; et al. Probing the Electronic Properties of Monolayer MoS2 via Interaction with Molecular Hydrogen. Adv. Electron. Mater. 2019, 5, 1800591. [Google Scholar] [CrossRef] [Green Version]
  125. Keong Koh, E.W.; Chiu, C.H.; Lim, Y.K.; Zhang, Y.-W.; Pan, H. Hydrogen adsorption on and diffusion through MoS2 monolayer: First-principles study. Int. J. Hydrogen Energy 2012, 37, 14323–14328. [Google Scholar] [CrossRef]
  126. Agrawal, A.V.; Kumar, R.; Venkatesan, S.; Zakhidov, A.; Zhu, Z.; Bao, J.; Kumar, M.; Kumar, M. Fast detection and low power hydrogen sensor using edge-oriented vertically aligned 3-D network of MoS2 flakes at room temperature. Appl. Phys. Lett. 2017, 111, 093102. [Google Scholar] [CrossRef] [Green Version]
  127. Stolyarova, S.G.; Kotsun, A.A.; Shubin, Y.V.; Koroteev, V.O.; Plyusnin, P.E.; Mikhlin, Y.L.; Mel’gunov, M.S.; Okotrub, A.V.; Bulusheva, L.G. Synthesis of Porous Nanostructured MoS2 Materials in Thermal Shock Conditions and Their Performance in Lithium-Ion Batteries. ACS Appl. Energy Mater. 2020, 3, 10802–10813. [Google Scholar] [CrossRef]
  128. Pereira, N.M.; Rezende, N.P.; Cunha, T.H.R.; Barboza, A.P.M.; Silva, G.G.; Lippross, D.; Neves, B.R.A.; Chacham, H.; Ferlauto, A.S.; Lacerda, R.G. Aerosol-Printed MoS2 Ink as a High Sensitivity Humidity Sensor. ACS Omega 2022, 7, 9388–9396. [Google Scholar] [CrossRef]
  129. Tong, Y.; Lin, Z.; Thong, J.T.L.; Chan, D.S.H.; Zhu, C. MoS2 oxygen sensor with gate voltage stress induced performance enhancement. Appl. Phys. Lett. 2015, 107, 123105. [Google Scholar] [CrossRef]
  130. Tongay, S.; Zhou, J.; Ataca, C.; Liu, J.; Kang, J.S.; Matthews, T.S.; You, L.; Li, J.; Grossman, J.C.; Wu, J. Broad-Range Modulation of Light Emission in Two-Dimensional Semiconductors by Molecular Physisorption Gating. Nano Lett. 2013, 13, 2831–2836. [Google Scholar] [CrossRef]
  131. Zhou, Q.; Hong, C.; Yao, Y.; Hussain, S.; Xu, L.; Zhang, Q.; Gui, Y.; Wang, M. Hierarchically MoS2 nanospheres assembled from nanosheets for superior CO gas-sensing properties. Mater. Res. Bull. 2018, 101, 132–139. [Google Scholar] [CrossRef]
  132. Li, H.; Huang, M.; Cao, G. Markedly different adsorption behaviors of gas molecules on defective monolayer MoS2: A first-principles study. Phys. Chem. Chem. Phys. 2016, 18, 15110–15117. [Google Scholar] [CrossRef]
  133. Zhao, S.; Xue, J.; Kang, W. Gas adsorption on MoS2 monolayer from first-principles calculations. Chem. Phys. Lett. 2014, 595–596, 35–42. [Google Scholar] [CrossRef] [Green Version]
  134. Shokri, A.; Salami, N. Gas sensor based on MoS2 monolayer. Sens. Actuators B Chem. 2016, 236, 378–385. [Google Scholar] [CrossRef]
  135. Babar, V.; Vovusha, H.; Schwingenschlögl, U. Density Functional Theory Analysis of Gas Adsorption on Monolayer and Few Layer Transition Metal Dichalcogenides: Implications for Sensing. ACS Appl. Nano Mater. 2019, 2, 6076–6080. [Google Scholar] [CrossRef]
  136. Yu, L.; Guo, F.; Liu, S.; Qi, J.; Yin, M.; Yang, B.; Liu, Z.; Fan, X.H. Hierarchical 3D flower-like MoS2 spheres: Post-thermal treatment in vacuum and their NO2 sensing properties. Mater. Lett. 2016, 183, 122–126. [Google Scholar] [CrossRef]
  137. Hung, C.M.; Vuong, V.A.; Van Duy, N.; Van An, D.; Van Hieu, N.; Kashif, M.; Hoa, N.D. Controlled Growth of Vertically Oriented Trilayer MoS2 Nanoflakes for Room-Temperature NO2 Gas Sensor Applications. Phys. Status Solidi 2020, 217, 2000004. [Google Scholar] [CrossRef]
  138. Hermawan, A.; Septiani, N.L.W.; Taufik, A.; Yuliarto, B.; Suyatman; Yin, S. Advanced Strategies to Improve Performances of Molybdenum-Based Gas Sensors. Nano-Micro Lett. 2021, 13, 207. [Google Scholar] [CrossRef]
  139. Shao, L.; Wu, Z.; Duan, H.; Shaymurat, T. Discriminative and rapid detection of ozone realized by sensor array of Zn2+ doping tailored MoS2 ultrathin nanosheets. Sens. Actuators B Chem. 2018, 258, 937–946. [Google Scholar] [CrossRef]
  140. Taufik, A.; Asakura, Y.; Hasegawa, T.; Kato, H.; Kakihana, M.; Hirata, S.; Inada, M.; Yin, S. Surface Engineering of 1T/2H-MoS2 Nanoparticles by O2 Plasma Irradiation as a Potential Humidity Sensor for Breathing and Skin Monitoring Applications. ACS Appl. Nano Mater. 2020, 3, 7835–7846. [Google Scholar] [CrossRef]
  141. Zhang, D.; Wu, J.; Li, P.; Cao, Y. Room-temperature SO2 gas-sensing properties based on a metal-doped MoS2 nanoflower: An experimental and density functional theory investigation. J. Mater. Chem. A 2017, 5, 20666–20677. [Google Scholar] [CrossRef]
  142. Bharathi, P.; Harish, S.; Mathankumar, G.; Krishna Mohan, M.; Archana, J.; Kamalakannan, S.; Prakash, M.; Shimomura, M.; Navaneethan, M. Solution processed edge activated Ni-MoS2 nanosheets for highly sensitive room temperature NO2 gas sensor applications. Appl. Surf. Sci. 2022, 600, 154086. [Google Scholar] [CrossRef]
  143. Taufik, A.; Asakura, Y.; Hasegawa, T.; Yin, S. MoS2–xSex Nanoparticles for NO Detection at Room Temperature. ACS Appl. Nano Mater. 2021, 4, 6861–6871. [Google Scholar] [CrossRef]
  144. Wu, R.; Hao, J.; Zheng, S.; Sun, Q.; Wang, T.; Zhang, D.; Zhang, H.; Wang, Y.; Zhou, X. N dopants triggered new active sites and fast charge transfer in MoS2 nanosheets for full Response-Recovery NO2 detection at room temperature. Appl. Surf. Sci. 2022, 571, 151162. [Google Scholar] [CrossRef]
  145. Ramaraj, S.G.; Nundy, S.; Zhao, P.; Elamaran, D.; Tahir, A.A.; Hayakawa, Y.; Muruganathan, M.; Mizuta, H.; Kim, S.-W. RF Sputtered Nb-Doped MoS2 Thin Film for Effective Detection of NO2 Gas Molecules: Theoretical and Experimental Studies. ACS Omega 2022, 7, 10492–10501. [Google Scholar] [CrossRef] [PubMed]
  146. Zhang, K.; Deng, D.D.; Zheng, B.; Wang, Y.; Perkins, F.K.; Briggs, N.C.; Crespi, V.H.; Robinson, J.A. Tuning Transport and Chemical Sensitivity via Niobium Doping of Synthetic MoS2. Adv. Mater. Interfaces 2020, 7, 2000856. [Google Scholar] [CrossRef]
  147. Zeng, Y.; Lin, S.; Gu, D.; Li, X. Two-Dimensional Nanomaterials for Gas Sensing Applications: The Role of Theoretical Calculations. Nanomaterials 2018, 8, 851. [Google Scholar] [CrossRef] [Green Version]
  148. Silvestrelli, P.L.; Ambrosetti, A. Including screening in van der Waals corrected density functional theory calculations: The case of atoms and small molecules physisorbed on graphene. J. Chem. Phys. 2014, 140, 124107. [Google Scholar] [CrossRef] [Green Version]
  149. Fan, Y.; Zhang, J.; Qiu, Y.; Zhu, J.; Zhang, Y.; Hu, G. A DFT study of transition metal (Fe, Co, Ni, Cu, Ag, Au, Rh, Pd, Pt and Ir)-embedded monolayer MoS2 for gas adsorption. Comput. Mater. Sci. 2017, 138, 255–266. [Google Scholar] [CrossRef]
  150. Szary, M.J. Adsorption of ethylene oxide on doped monolayers of MoS2: A DFT study. Mater. Sci. Eng. B 2021, 265, 115009. [Google Scholar] [CrossRef]
  151. Deng, X.; Liang, X.; Ng, S.-P.; Wu, C.-M.L. Adsorption of formaldehyde on transition metal doped monolayer MoS2: A DFT study. Appl. Surf. Sci. 2019, 484, 1244–1252. [Google Scholar] [CrossRef]
  152. Sharma, A.; Anu; Khan, M.S.; Husain, M.; Khan, M.S.; Srivastava, A. Sensing of CO and NO on Cu-Doped MoS2 Monolayer-Based Single Electron Transistor: A First Principles Study. IEEE Sens. J. 2018, 18, 2853–2860. [Google Scholar] [CrossRef]
  153. Song, J.; Lou, H. Improvement of gas-adsorption performances of Ag-functionalized monolayer MoS2 surfaces: A first-principles study. J. Appl. Phys. 2018, 123, 175303. [Google Scholar] [CrossRef]
  154. Salih, E.; Ayesh, A.I. First principle study of transition metals codoped MoS2 as a gas sensor for the detection of NO and NO2 gases. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 131, 114736. [Google Scholar] [CrossRef]
  155. Zhu, J.; Zhang, H.; Tong, Y.; Zhao, L.; Zhang, Y.; Qiu, Y.; Lin, X. First-principles investigations of metal (V, Nb, Ta)-doped monolayer MoS2: Structural stability, electronic properties and adsorption of gas molecules. Appl. Surf. Sci. 2017, 419, 522–530. [Google Scholar] [CrossRef]
  156. Gui, Y.; Liu, D.; Li, X.; Tang, C.; Zhou, Q. DFT-based study on H2S and SOF2 adsorption on Si-MoS2 monolayer. Results Phys. 2019, 13, 102225. [Google Scholar] [CrossRef]
  157. Szary, M.J.; Bąbelek, J.A.; Florjan, D.M. Adsorption and dissociation of NO2 on MoS2 doped wi. Surf. Sci. 2021, 712, 121893. [Google Scholar] [CrossRef]
  158. Szary, M.J. MoS2 doping for enhanced H2S detection. Appl. Surf. Sci. 2021, 547, 149026. [Google Scholar] [CrossRef]
  159. Yue, Q.; Shao, Z.; Chang, S.; Li, J. Adsorption of gas molecules on monolayer MoS2 and effect of applied electric field. Nanoscale Res. Lett. 2013, 8, 425. [Google Scholar] [CrossRef] [Green Version]
  160. Pramudita, J.C.; Sehrawat, D.; Goonetilleke, D.; Sharma, N. An Initial Review of the Status of Electrode Materials for Potassium-Ion Batteries. Adv. Energy Mater. 2017, 7, 1602911. [Google Scholar] [CrossRef]
  161. Zhang, R.; Qin, Y.; Liu, P.; Jia, C.; Tang, Y.; Wang, H. How does Molybdenum Disulfide Store Charge: A Minireview. ChemSusChem 2020, 13, 1354–1365. [Google Scholar] [CrossRef]
  162. Afanasiev, P. Synthetic approaches to the molybdenum sulfide materials. Comptes Rendus Chim. 2008, 11, 159–182. [Google Scholar] [CrossRef]
  163. Wang, H.; Tran, D.; Qian, J.; Ding, F.; Losic, D. MoS2/Graphene Composites as Promising Materials for Energy Storage and Conversion Applications. Adv. Mater. Interfaces 2019, 6, 1900915. [Google Scholar] [CrossRef]
  164. Xia, D.; Gong, F.; Pei, X.; Wang, W.; Li, H.; Zeng, W.; Wu, M.; Papavassiliou, D.V. Molybdenum and tungsten disulfides-based nanocomposite films for energy storage and conversion: A review. Chem. Eng. J. 2018, 348, 908–928. [Google Scholar] [CrossRef]
  165. Li, X.L.; Li, T.C.; Huang, S.; Zhang, J.; Pam, M.E.; Yang, H.Y. Controllable Synthesis of Two-Dimensional Molybdenum Disulfide (MoS2) for Energy-Storage Applications. ChemSusChem 2020, 13, 1379–1391. [Google Scholar] [CrossRef] [PubMed]
  166. Ullah, N.; Guziejewski, D.; Yuan, A.; Shah, S.A. Recent Advancement and Structural Engineering in Transition Metal Dichalcogenides for Alkali Metal Ions Batteries. Materials 2023, 16, 2559. [Google Scholar] [CrossRef]
  167. Yang, Z.; Zhu, L.; Lv, C.; Zhang, R.; Wang, H.; Wang, J.; Zhang, Q. Defect engineering of molybdenum disulfide for energy storage. Mater. Chem. Front. 2021, 5, 5880–5896. [Google Scholar] [CrossRef]
  168. Li, B.; Cao, W.; Wang, S.; Cao, Z.; Shi, Y.; Niu, J.; Wang, F. N,S-Doped Porous Carbon Nanobelts Embedded with MoS2 Nanosheets as a Self-Standing Host for Dendrite-Free Li Metal Anodes. Adv. Sci. 2022, 9, 2204232. [Google Scholar] [CrossRef]
  169. Sen, U.K.; Johari, P.; Basu, S.; Nayak, C.; Mitra, S. An experimental and computational study to understand the lithium storage mechanism in molybdenum disulfide. Nanoscale 2014, 6, 10243–10254. [Google Scholar] [CrossRef]
  170. Zhang, L.; Sun, D.; Kang, J.; Feng, J.; Bechtel, H.A.; Wang, L.-W.; Cairns, E.J.; Guo, J. Electrochemical Reaction Mechanism of the MoS2 Electrode in a Lithium-Ion Cell Revealed by in Situ and Operando X-ray Absorption Spectroscopy. Nano Lett. 2018, 18, 1466–1475. [Google Scholar] [CrossRef] [Green Version]
  171. Su, Q.; Wang, S.; Feng, M.; Du, G.; Xu, B. Direct Studies on the Lithium-Storage Mechanism of Molybdenum Disulfide. Sci. Rep. 2017, 7, 7275. [Google Scholar] [CrossRef]
  172. Li, X.; Zai, J.; Xiang, S.; Liu, Y.; He, X.; Xu, Z.; Wang, K.; Ma, Z.; Qian, X. Regeneration of Metal Sulfides in the Delithiation Process: The Key to Cyclic Stability. Adv. Energy Mater. 2016, 6, 1601056. [Google Scholar] [CrossRef]
  173. Zhu, Z.; Xi, S.; Miao, L.; Tang, Y.; Zeng, Y.; Xia, H.; Lv, Z.; Zhang, W.; Ge, X.; Zhang, H.; et al. Unraveling the Formation of Amorphous MoS2 Nanograins during the Electrochemical Delithiation Process. Adv. Funct. Mater. 2019, 29, 1904843. [Google Scholar] [CrossRef]
  174. Stolyarova, S.G.; Koroteev, V.O.; Shubin, Y.V.; Plyusnin, P.E.; Makarova, A.A.; Okotrub, A.V.; Bulusheva, L.G. Pressure-Assisted Interface Engineering in MoS2 /Holey Graphene Hybrids for Improved Performance in Li-ion Batteries. Energy Technol. 2019, 7, 1900659. [Google Scholar] [CrossRef]
  175. Deng, C.; Wang, Z.; Wang, S.; Yu, J. Inhibition of polysulfide diffusion in lithium–sulfur batteries: Mechanism and improvement strategies. J. Mater. Chem. A 2019, 7, 12381–12413. [Google Scholar] [CrossRef]
  176. Guo, P.; Sun, K.; Liu, D.; Cheng, P.; Lv, M.; Liu, Q.; He, D. Molybdenum disulfide nanosheets embedded in hollow nitrogen-doped carbon spheres for efficient lithium/sodium storage with enhanced electrochemical kinetics. Electrochim. Acta 2018, 283, 646–654. [Google Scholar] [CrossRef]
  177. Liu, M.; Fan, H.; Zhuo, O.; Du, X.; Yang, L.; Wang, P.; Yang, L.; Wu, Q.; Wang, X.; Hu, Z. Vertically Grown Few-Layer MoS2 Nanosheets on Hierarchical Carbon Nanocages for Pseudocapacitive Lithium Storage with Ultrahigh-Rate Capability and Long-Term Recyclability. Chem. A Eur. J. 2019, 25, 3843–3848. [Google Scholar] [CrossRef]
  178. Koroteev, V.O.; Stolyarova, S.G.; Kotsun, A.A.; Modin, E.; Makarova, A.A.; Shubin, Y.; Plyusnin, P.E.; Okotrub, A.V.; Bulusheva, L.G. Nanoscale coupling of MoS2 and graphene via rapid thermal decomposition of ammonium tetrathiomolybdate and graphite oxide for boosting capacity of Li-ion batteries. Carbon N. Y. 2021, 173, 194–204. [Google Scholar] [CrossRef]
  179. Wang, L.; Zhang, Q.; Zhu, J.; Duan, X.; Xu, Z.; Liu, Y.; Yang, H.; Lu, B. Nature of extra capacity in MoS2 electrodes: Molybdenum atoms accommodate with lithium. Energy Storage Mater. 2019, 16, 37–45. [Google Scholar] [CrossRef]
  180. Shu, H.; Li, F.; Hu, C.; Liang, P.; Cao, D.; Chen, X. The capacity fading mechanism and improvement of cycling stability in MoS2 -based anode materials for lithium-ion batteries. Nanoscale 2016, 8, 2918–2926. [Google Scholar] [CrossRef]
  181. Feng, C.; Ma, J.; Li, H.; Zeng, R.; Guo, Z.; Liu, H. Synthesis of molybdenum disulfide (MoS2) for lithium ion battery applications. Mater. Res. Bull. 2009, 44, 1811–1815. [Google Scholar] [CrossRef]
  182. Liu, Q.; Weijun, X.; Wu, Z.; Huo, J.; Liu, D.; Wang, Q.; Wang, S. The origin of the enhanced performance of nitrogen-doped MoS2 in lithium ion batteries. Nanotechnology 2016, 27, 175402. [Google Scholar] [CrossRef]
  183. Li, S. N-doped MoS2 Nano-Flowers as High-Performance Anode Electrode for Excellent Lithium Storage. Int. J. Electrochem. Sci. 2019, 14, 7507–7515. [Google Scholar] [CrossRef] [Green Version]
  184. Jiao, J.; Du, K.; Wang, Y.; Sun, P.; Zhao, H.; Tang, P.; Fan, Q.; Tian, H.; Li, Q.; Xu, Q. N plasma treatment on graphene oxide-MoS2 composites for improved performance in lithium ion batteries. Mater. Chem. Phys. 2020, 240, 122169. [Google Scholar] [CrossRef]
  185. Tang, P.; Jiao, J.; Fan, Q.; Wang, X.; Agrawal, V.; Xu, Q. Interlayer spacing engineering in N doped MoS2 for efficient lithium ion storage. Mater. Chem. Phys. 2021, 261, 124166. [Google Scholar] [CrossRef]
  186. Cho, J.; Ryu, S.; Gong, Y.J.; Pyo, S.; Yun, H.; Kim, H.; Lee, J.; Yoo, J.; Kim, Y.S. Nitrogen-doped MoS2 as a catalytic sulfur host for lithium-sulfur batteries. Chem. Eng. J. 2022, 439, 135568. [Google Scholar] [CrossRef]
  187. Fayed, M.G.; Attia, S.Y.; Barakat, Y.F.; El-Shereafy, E.E.; Rashad, M.M.; Mohamed, S.G. Carbon and nitrogen co-doped MoS2 nanoflakes as an electrode material for lithium-ion batteries and supercapacitors. Sustain. Mater. Technol. 2021, 29, e00306. [Google Scholar] [CrossRef]
  188. Gong, S.; Zhao, G.; Lyu, P.; Sun, K. A Pseudolayered MoS2 as Li-Ion Intercalation Host with Enhanced Rate Capability and Durability. Small 2018, 14, 1803344. [Google Scholar] [CrossRef]
  189. Lu, L.; Min, F.; Luo, Z.; Wang, S.; Teng, F.; Li, G.; Feng, C. Synthesis and electrochemical properties of tin-doped MoS2 (Sn/MoS2) composites for lithium ion battery applications. J. Nanoparticle Res. 2016, 18, 357. [Google Scholar] [CrossRef]
  190. Lei, Y.; Fujisawa, K.; Zhang, F.; Briggs, N.; Aref, A.R.; Yeh, Y.-T.; Lin, Z.; Robinson, J.A.; Rajagopalan, R.; Terrones, M. Synthesis of V-MoS2 Layered Alloys as Stable Li-Ion Battery Anodes. ACS Appl. Energy Mater. 2019, 2, 8625–8632. [Google Scholar] [CrossRef]
  191. Cai, G.; Peng, L.; Ye, S.; Huang, Y.; Wang, G.; Zhang, X. Defect-rich MoS2(1−x) Se2x few-layer nanocomposites: A superior anode material for high-performance lithium-ion batteries. J. Mater. Chem. A 2019, 7, 9837–9843. [Google Scholar] [CrossRef]
  192. Wang, S.; Liu, B.; Zhi, G.; Gong, X.; Gao, Z.; Zhang, J. Relaxing volume stress and promoting active sites in vertically grown 2D layered mesoporous MoS2(1-x)Se2x/rGO composites with enhanced capability and stability for lithium ion batteries. Electrochim. Acta 2018, 268, 424–434. [Google Scholar] [CrossRef]
  193. Francis, M.K.; Rajesh, K.; Bhargav, P.B.; Ahmed, N. Binder-free phosphorus-doped MoS2 flexible anode deposited on carbon cloth for high-capacity Li-ion battery applications. J. Mater. Sci. 2023, 58, 4054–4069. [Google Scholar] [CrossRef]
  194. Wang, J.; Zhang, L.; Sun, K.; He, J.; Zheng, Y.; Xu, C.; Zhang, Y.; Chen, Y.; Li, M. Improving ionic/electronic conductivity of MoS2 Li-ion anode via manganese doping and structural optimization. Chem. Eng. J. 2019, 372, 665–672. [Google Scholar] [CrossRef]
  195. Qi, K.; Yuan, Z.; Hou, Y.; Zhao, R.; Zhang, B. Facile synthesis and improved Li-storage performance of Fe-doped MoS2/reduced graphene oxide as anode materials. Appl. Surf. Sci. 2019, 483, 688–695. [Google Scholar] [CrossRef]
  196. Xu, S.; Zhu, Q.; Chen, T.; Chen, W.; Ye, J.; Huang, J. Hydrothermal synthesis of Co-doped-MoS2/reduced graphene oxide hybrids with enhanced electrochemical lithium storage performances. Mater. Chem. Phys. 2018, 219, 399–410. [Google Scholar] [CrossRef]
  197. Li, X.; Feng, Z.; Zai, J.; Ma, Z.-F.; Qian, X. Incorporation of Co into MoS2/graphene nanocomposites: One effective way to enhance the cycling stability of Li/Na storage. J. Power Sources 2018, 373, 103–109. [Google Scholar] [CrossRef]
  198. Han, M.; Chen, J.; Cai, Y.; Wei, L.; Zhao, T. Magnetic-atom strategy enables unilamellar MoS2-C interoverlapped superstructure with ultrahigh capacity and ultrafast ion transfer capability in Li/Na/K-ion batteries. Chem. Eng. J. 2023, 454, 140137. [Google Scholar] [CrossRef]
  199. Sun, X.; Wang, Z. Adsorption and diffusion of lithium on heteroatom-doped monolayer molybdenum disulfide. Appl. Surf. Sci. 2018, 455, 911–918. [Google Scholar] [CrossRef]
  200. Yang, F.; Feng, X.; Glans, P.-A.; Guo, J. MoS2 for beyond lithium-ion batteries. APL Mater. 2021, 9, 050903. [Google Scholar] [CrossRef]
  201. Li, Q.; Yao, Z.; Wu, J.; Mitra, S.; Hao, S.; Sahu, T.S.; Li, Y.; Wolverton, C.; Dravid, V.P. Intermediate phases in sodium intercalation into MoS2 nanosheets and their implications for sodium-ion batteries. Nano Energy 2017, 38, 342–349. [Google Scholar] [CrossRef]
  202. Du, X.; Huang, J.; Guo, X.; Lin, X.; Huang, J.-Q.; Tan, H.; Zhu, Y.; Zhang, B. Preserved Layered Structure Enables Stable Cyclic Performance of MoS2 upon Potassium Insertion. Chem. Mater. 2019, 31, 8801–8809. [Google Scholar] [CrossRef]
  203. Wang, P.; Gou, W.; Jiang, T.; Zhao, W.; Ding, K.; Sheng, H.; Liu, X.; Xu, Q.; Fan, Q. An interlayer spacing design approach for efficient sodium ion storage in N-doped MoS2. Nanoscale Horiz. 2023, 8, 473–482. [Google Scholar] [CrossRef] [PubMed]
  204. Cai, Y.; Yang, H.; Zhou, J.; Luo, Z.; Fang, G.; Liu, S.; Pan, A.; Liang, S. Nitrogen doped hollow MoS2/C nanospheres as anode for long-life sodium-ion batteries. Chem. Eng. J. 2017, 327, 522–529. [Google Scholar] [CrossRef]
  205. Tao, P.; He, J.; Shen, T.; Hao, Y.; Yan, J.; Huang, Z.; Xu, X.; Li, M.; Chen, Y. Nitrogen-Doped MoS2 Foam for Fast Sodium Ion Storage. Adv. Mater. Interfaces 2019, 6, 1900460. [Google Scholar] [CrossRef]
  206. Lim, H.; Kim, H.; Kim, S.-O.; Choi, W. Self-assembled N-doped MoS2/carbon spheres by naturally occurring acid-catalyzed reaction for improved sodium-ion batteries. Chem. Eng. J. 2020, 387, 124144. [Google Scholar] [CrossRef]
  207. Lim, H.; Yu, S.; Choi, W.; Kim, S.-O. Hierarchically Designed Nitrogen-Doped MoS2/Silicon Oxycarbide Nanoscale Heterostructure as High-Performance Sodium-Ion Battery Anode. ACS Nano 2021, 15, 7409–7420. [Google Scholar] [CrossRef]
  208. Li, J.; He, T.; Zhao, Y.; Zhang, X.; Zhong, W.; Zhang, X.; Ren, J.; Chen, Y. In-situ N-doped ultrathin MoS2 anchored on N-doped carbon nanotubes skeleton by Mo-N bonds for fast pseudocapacitive sodium storage. J. Alloys Compd. 2022, 897, 163170. [Google Scholar] [CrossRef]
  209. Liu, Y.; Lei, Z.; Li, X.; Lin, C.; Liu, R.; Cao, C.; Chen, Q.; Wei, M.; Zeng, L.; Qian, Q. Sb-Doped metallic 1T-MoS2 nanosheets embedded in N-doped carbon as high-performance anode materials for half/full sodium/potassium-ion batteries. Dalt. Trans. 2022, 51, 11685–11692. [Google Scholar] [CrossRef]
  210. Fan, B.-B.; Fan, H.-N.; Chen, X.-H.; Gao, X.-W.; Chen, S.; Tang, Q.-L.; Luo, W.-B.; Deng, Y.; Hu, A.-P.; Hu, W. Metallic-State MoS2 Nanosheets with Atomic Modification for Sodium Ion Batteries with a High Rate Capability and Long Lifespan. ACS Appl. Mater. Interfaces 2021, 13, 19894–19903. [Google Scholar] [CrossRef]
  211. Zong, J.; Wang, F.; Liu, H.; Zhao, M.; Yang, S. Te-doping induced C@MoS2-xTex@C nanocomposites with improved electronic structure as high-performance anode for sodium-based dual-ion batteries. J. Power Sources 2022, 535, 231462. [Google Scholar] [CrossRef]
  212. Woo, S.H.; Yadgarov, L.; Rosentsveig, R.; Park, Y.; Song, D.; Tenne, R.; Hong, S.Y. Fullerene-like Re-Doped MoS2 Nanoparticles as an Intercalation Host with Fast Kinetics for Sodium Ion Batteries. Isr. J. Chem. 2015, 55, 599–603. [Google Scholar] [CrossRef]
  213. Zheng, T.; Li, G.; Dong, J.; Sun, Q.; Meng, X. Self-assembled Mn-doped MoS2 hollow nanotubes with significantly enhanced sodium storage for high-performance sodium-ion batteries. Inorg. Chem. Front. 2018, 5, 1587–1593. [Google Scholar] [CrossRef]
  214. Yue, X.; Wang, J.; Patil, A.M.; An, X.; Xie, Z.; Hao, X.; Jiang, Z.; Abudula, A.; Guan, G. A novel vanadium-mediated MoS2 with metallic behavior for sodium ion batteries: Achieving fast Na+ diffusion to enhance electrochemical kinetics. Chem. Eng. J. 2021, 417, 128107. [Google Scholar] [CrossRef]
  215. Yue, X.; Wang, J.; Xie, Z.; He, Y.; Liu, Z.; Liu, C.; Hao, X.; Abudula, A.; Guan, G. Controllable Synthesis of Novel Orderly Layered VMoS2 Anode Materials with Super Electrochemical Performance for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2021, 13, 26046–26054. [Google Scholar] [CrossRef]
  216. Fan, H.; Wang, X.; Yu, H.; Gu, Q.; Chen, S.; Liu, Z.; Chen, X.; Luo, W.; Liu, H. Enhanced Potassium Ion Battery by Inducing Interlayer Anionic Ligands in MoS1.5Se0.5 Nanosheets with Exploration of the Mechanism. Adv. Energy Mater. 2020, 10, 1904162. [Google Scholar] [CrossRef]
  217. Kang, W.; Wang, Y.; An, C. Interlayer engineering of MoS2 nanosheets for high-rate potassium-ion storage. New J. Chem. 2020, 44, 20659–20664. [Google Scholar] [CrossRef]
  218. Gao, J.; Wang, G.; Liu, Y.; Li, J.; Peng, B.; Jiao, S.; Zeng, S.; Zhang, G. Ternary molybdenum sulfoselenide based hybrid nanotubes boost potassium-ion diffusion kinetics for high energy/power hybrid capacitors. J. Mater. Chem. A 2020, 8, 13946–13954. [Google Scholar] [CrossRef]
  219. He, H.; Huang, D.; Gan, Q.; Hao, J.; Liu, S.; Wu, Z.; Pang, W.K.; Johannessen, B.; Tang, Y.; Luo, J.-L.; et al. Anion Vacancies Regulating Endows MoSSe with Fast and Stable Potassium Ion Storage. ACS Nano 2019, 13, 11843–11852. [Google Scholar] [CrossRef]
  220. Kang, W.; Xie, R.; Wang, Y.; An, C.; Li, C. Te–S covalent bond induces 1T&2H MoS2 with improved potassium-ion storage performance. Nanoscale 2020, 12, 24463–24470. [Google Scholar] [CrossRef]
  221. Zhang, K.; Zhang, H.; Li, C.; Ma, X. Vertically oriented MoSe0.4S1.6/N-doped C nanostructures directly grown on carbon nanotubes as high-performance anode for potassium-ion batteries. J. Electroanal. Chem. 2022, 923, 116846. [Google Scholar] [CrossRef]
  222. Raghavendra, K.V.G.; Vinoth, R.; Zeb, K.; Muralee Gopi, C.V.V.; Sambasivam, S.; Kummara, M.R.; Obaidat, I.M.; Kim, H.J. An intuitive review of supercapacitors with recent progress and novel device applications. J. Energy Storage 2020, 31, 101652. [Google Scholar] [CrossRef]
  223. Mohan, M.; Shetti, N.P.; Aminabhavi, T.M. Phase dependent performance of MoS2 for supercapacitor applications. J. Energy Storage 2023, 58, 106321. [Google Scholar] [CrossRef]
  224. Koroteev, V.O.; Kuznetsova, I.V.; Kurenya, A.G.; Kanygin, M.A.; Fedorovskaya, E.O.; Mikhlin, Y.L.; Chuvilin, A.L.; Bulusheva, L.G.; Okotrub, A.V. Enhanced supercapacitance of vertically aligned multi-wall carbon nanotube array covered by MoS2 nanoparticles. Phys. Status Solidi 2016, 253, 2451–2456. [Google Scholar] [CrossRef]
  225. Wang, L.; Wu, J.; Wang, X.; Fu, S. Defect engineering of MoS2-based materials as supercapacitors electrode: A mini review. J. Alloys Compd. 2023, 959, 170548. [Google Scholar] [CrossRef]
  226. Pisal, K.B.; Babar, B.M.; Mujawar, S.H.; Kadam, L.D. Overview of molybdenum disulfide based electrodes for supercapacitors: A short review. J. Energy Storage 2021, 43, 103297. [Google Scholar] [CrossRef]
  227. Rohith, R.; Prasannakumar, A.T.; Mohan, R.R.; V., M.; Varma, S.J. Advances in 2D Molybdenum Disulfide-Based Functional Materials for Supercapacitor Applications. ChemistrySelect 2022, 7, e202203068. [Google Scholar] [CrossRef]
  228. Kanade, C.; Arbuj, S.; Kanade, K.; Kim, K.S.; Yeom, G.Y.; Kim, T.; Kale, B. Hierarchical nanostructures of nitrogen-doped molybdenum sulphide for supercapacitors. RSC Adv. 2018, 8, 39749–39755. [Google Scholar] [CrossRef]
  229. Le, K.; Zhang, X.; Zhao, Q.; Liu, Y.; Yi, P.; Xu, S.; Liu, W. Controllably Doping Nitrogen into 1T/2H MoS2 Heterostructure Nanosheets for Enhanced Supercapacitive and Electrocatalytic Performance by Low-Power N2 Plasma. ACS Appl. Mater. Interfaces 2021, 13, 44427–44439. [Google Scholar] [CrossRef] [PubMed]
  230. Singha, S.S.; Rudra, S.; Mondal, S.; Pradhan, M.; Nayak, A.K.; Satpati, B.; Pal, P.; Das, K.; Singha, A. Mn incorporated MoS2 nanoflowers: A high performance electrode material for symmetric supercapacitor. Electrochim. Acta 2020, 338, 135815. [Google Scholar] [CrossRef]
  231. Wang, Z.; Wang, J.; Wang, F.; Zhang, X.; He, X.; Liu, S.; Zhang, Z.; Zhang, Z.; Wang, X. One-step hydrothermal synthesis of high-performance stable Ni-doped 1T-MoS2 electrodes for supercapacitors. J. Alloys Compd. 2023, 947, 169505. [Google Scholar] [CrossRef]
  232. Prakash, K.; Harish, S.; Kamalakannan, S.; Logu, T.; Shimomura, M.; Archana, J.; Navaneethan, M. Perspective on ultrathin layered Ni-doped MoS2 hybrid nanostructures for the enhancement of electrochemical properties in supercapacitors. J. Energy Chem. 2023, 80, 335–349. [Google Scholar] [CrossRef]
  233. Rohith, R.; Manuraj, M.; Jafri, R.I.; Rakhi, R.B. Co-MoS2 nanoflower coated carbon fabric as a flexible electrode for supercapacitor. Mater. Today Proc. 2022, 50, 1–6. [Google Scholar] [CrossRef]
  234. Sun, A.; Xie, L.; Wang, D.; Wu, Z. Enhanced energy storage performance from Co-decorated MoS2 nanosheets as supercapacitor electrode materials. Ceram. Int. 2018, 44, 13434–13438. [Google Scholar] [CrossRef]
  235. Pan, U.N.; Sharma, V.; Kshetri, T.; Singh, T.I.; Paudel, D.R.; Kim, N.H.; Lee, J.H. Freestanding 1T-MnxMo1– xS2– ySey and MoFe2S4–zSez Ultrathin Nanosheet-Structured Electrodes for Highly Efficient Flexible Solid-State Asymmetric Supercapacitors. Small 2020, 16, 2001691. [Google Scholar] [CrossRef] [PubMed]
  236. Sun, T.; Li, Z.; Liu, X.; Ma, L.; Wang, J.; Yang, S. Oxygen-incorporated MoS2 microspheres with tunable interiors as novel electrode materials for supercapacitors. J. Power Sources 2017, 352, 135–142. [Google Scholar] [CrossRef]
  237. Lee, S.; Hwang, J.; Kim, D.; Ahn, H. Oxygen incorporated in 1T/2H hybrid MoS2 nanoflowers prepared from molybdenum blue solution for asymmetric supercapacitor applications. Chem. Eng. J. 2021, 419, 129701. [Google Scholar] [CrossRef]
  238. Sellam, A.; Jenjeti, R.N.; Sampath, S. Ultrahigh-Rate Supercapacitors Based on 2-Dimensional, 1T MoS2 xSe2(1– x) for AC Line-Filtering Applications. J. Phys. Chem. C 2018, 122, 14186–14194. [Google Scholar] [CrossRef]
  239. Manuraj, M.; Mohan, V.V.; Assa Aravindh, S.; Sarath Kumar, S.R.; Narayanan Unni, K.N.; Rakhi, R.B. Can mixed anion transition metal dichalcogenide electrodes enhance the performance of electrochemical energy storage devices? The case of MoS2xSe2(1-x). Chem. Eng. J. 2022, 443, 136451. [Google Scholar] [CrossRef]
  240. Xu, J.; Dong, L.; Li, C.; Tang, H. Facile synthesis of Mo0.91W0.09S2 ultrathin nanosheets/amorphous carbon composites for high-performance supercapacitor. Mater. Lett. 2016, 162, 126–130. [Google Scholar] [CrossRef]
  241. Bello, I.T.; Otun, K.O.; Nyongombe, G.; Adedokun, O.; Kabongo, G.L.; Dhlamini, M.S. Synthesis, Characterization, and Supercapacitor Performance of a Mixed-Phase Mn-Doped MoS2 Nanoflower. Nanomaterials 2022, 12, 490. [Google Scholar] [CrossRef]
  242. Palanisamy, S.; Periasamy, P.; Subramani, K.; Shyma, A.P.; Venkatachalam, R. Ultrathin sheet structure Ni-MoS2 anode and MnO2/water dispersion graphene cathode for modern asymmetrical coin cell supercapacitor. J. Alloys Compd. 2018, 731, 936–944. [Google Scholar] [CrossRef]
  243. Panchu, S.J.; Raju, K.; Singh, P.; Johnson, D.D.; Swart, H.C. High Mass Loading of Flowerlike Ni-MoS2 Microspheres toward Efficient Intercalation Pseudocapacitive Electrodes. ACS Appl. Energy Mater. 2023, 6, 2187–2198. [Google Scholar] [CrossRef]
  244. Vikraman, D.; Hussain, S.; Karuppasamy, K.; Kathalingam, A.; Jo, E.-B.; Sanmugam, A.; Jung, J.; Kim, H.-S. Engineering the active sites tuned MoS2 nanoarray structures by transition metal doping for hydrogen evolution and supercapacitor applications. J. Alloys Compd. 2022, 893, 162271. [Google Scholar] [CrossRef]
  245. Falola, B.D.; Fan, L.; Wiltowski, T.; Suni, I.I. Electrodeposition of Cu-Doped MoS2 for Charge Storage in Electrochemical Supercapacitors. J. Electrochem. Soc. 2017, 164, D674–D679. [Google Scholar] [CrossRef]
  246. Lu, Q.; Wei, Z.; Liang, J.; Li, L.; Ma, J.; Li, C. Electrochemical properties of Mo0.7Co0.3S2/g-C3N4 nanocomposites prepared by solvothermal method. J. Mater. Sci. Mater. Electron. 2021, 32, 28152–28162. [Google Scholar] [CrossRef]
  247. Lu, Q.; Wei, Z.; Liang, J.; Li, C.; Li, L.; Ma, J. High-Performance Supercapacitor Electrode Materials of Mo1-xCox S2 Nanoparticles Prepared by Hydrothermal Method. ChemistrySelect 2022, 7, e202102794. [Google Scholar] [CrossRef]
  248. Shao, J.; Li, Y.; Zhong, M.; Wang, Q.; Luo, X.; Li, K.; Zhao, W. Enhanced-performance flexible supercapacitor based on Pt-doped MoS2. Mater. Lett. 2019, 252, 173–177. [Google Scholar] [CrossRef]
Figure 1. Structures show the MoS2 fragment in top and side projections. Elements considered as heteroatoms in MoS2 in theoretical works are located in the shaded cells of the periodic table. Elements located in green and cream cells cause p- and n-type doping, respectively. Elements in yellow cells are isovalent to molybdenum or sulfur, which are found in light yellow cells. Gray color of the cell with Sb indicates the metallic type due to doping. Elements of red color were introduced into MoS2 experimentally.
Figure 1. Structures show the MoS2 fragment in top and side projections. Elements considered as heteroatoms in MoS2 in theoretical works are located in the shaded cells of the periodic table. Elements located in green and cream cells cause p- and n-type doping, respectively. Elements in yellow cells are isovalent to molybdenum or sulfur, which are found in light yellow cells. Gray color of the cell with Sb indicates the metallic type due to doping. Elements of red color were introduced into MoS2 experimentally.
Nanomaterials 13 02182 g001
Figure 2. Binding energies of metal atoms from the four, five, and six periods with a molybdenum vacancy in monolayer MoS2, taken with permission from [53]; Copyright 2013, American Physical Society.
Figure 2. Binding energies of metal atoms from the four, five, and six periods with a molybdenum vacancy in monolayer MoS2, taken with permission from [53]; Copyright 2013, American Physical Society.
Nanomaterials 13 02182 g002
Figure 3. (a) Electron density distribution of impurity state associated with the introduction of Nb atom (blue sphere) into MoS2; taken with permission from [61]; Copyright 2008, American Physical Society. (b) Model of MoS2 with 33% Nb concentration as triatomic clusters centered on a hole of a hexagon.
Figure 3. (a) Electron density distribution of impurity state associated with the introduction of Nb atom (blue sphere) into MoS2; taken with permission from [61]; Copyright 2008, American Physical Society. (b) Model of MoS2 with 33% Nb concentration as triatomic clusters centered on a hole of a hexagon.
Nanomaterials 13 02182 g003
Figure 4. Partial density of states (top figures) calculated for ideal MoS2 monolayer and that with two substitutional N atoms and monovacancy (models in bottom). Dashed vertical lines correspond to the top of valence band, the numbers specify band gap values. Taken with permission from [78]; Copyright 2023, Elsevier B.V.
Figure 4. Partial density of states (top figures) calculated for ideal MoS2 monolayer and that with two substitutional N atoms and monovacancy (models in bottom). Dashed vertical lines correspond to the top of valence band, the numbers specify band gap values. Taken with permission from [78]; Copyright 2023, Elsevier B.V.
Nanomaterials 13 02182 g004
Figure 5. (a) Experimental (left column) and simulated (right column) atom-resolved STEM images for triatomic Co cluster (top row) and isolated Co atom (bottom row) replacing Mo atoms in a MoS2 monolayer. Purple, blue, and yellow balls in the models correspond to Mo, Co, and S, respectively. (b) EELS spectrum taken from places presented in (a), the arrow shows the Co L2,3 edge, enlarged in the inset. Taken with permission from [82]; Copyright 2020, John Wiley and Sons.
Figure 5. (a) Experimental (left column) and simulated (right column) atom-resolved STEM images for triatomic Co cluster (top row) and isolated Co atom (bottom row) replacing Mo atoms in a MoS2 monolayer. Purple, blue, and yellow balls in the models correspond to Mo, Co, and S, respectively. (b) EELS spectrum taken from places presented in (a), the arrow shows the Co L2,3 edge, enlarged in the inset. Taken with permission from [82]; Copyright 2020, John Wiley and Sons.
Nanomaterials 13 02182 g005
Figure 6. (a) Response of undoped MoS2 and doped MoS2 sensors toward varying concentrations of SO2; (b) selectivity of Ni-doped MoS2 sensor for 500 ppm of various gases; taken with permission from [141]; Copyright 2017, The Royal Society of Chemistry.
Figure 6. (a) Response of undoped MoS2 and doped MoS2 sensors toward varying concentrations of SO2; (b) selectivity of Ni-doped MoS2 sensor for 500 ppm of various gases; taken with permission from [141]; Copyright 2017, The Royal Society of Chemistry.
Nanomaterials 13 02182 g006
Figure 7. (a) Response of MoS2−xSex (x = 0, 0.2, 1, 1.8, 2) sensors for varying concentration of NO; (b) Selectivity of MoSSe sensor toward various gases; taken with permission from [143]; Copyright 2021, American Chemical Society.
Figure 7. (a) Response of MoS2−xSex (x = 0, 0.2, 1, 1.8, 2) sensors for varying concentration of NO; (b) Selectivity of MoSSe sensor toward various gases; taken with permission from [143]; Copyright 2021, American Chemical Society.
Nanomaterials 13 02182 g007
Figure 8. (a) Top and side view of NO2 located above Ag (green ball) substituent in MoS2. The values show the distances (in Å) between Ag and O atoms of NO2; taken with permission from [153]; Copyright 2023, AIP Publishing; (b) charge density difference plots and value of charge transferred for Ag-doped MoS2 (left) and undoped MoS2 (right) with adsorbed NO2. Blue (yellow) in left and green (red) in right correspond to depletion (excess) of electron density; taken with permission from [159]; Copyright 2013 by the authors.
Figure 8. (a) Top and side view of NO2 located above Ag (green ball) substituent in MoS2. The values show the distances (in Å) between Ag and O atoms of NO2; taken with permission from [153]; Copyright 2023, AIP Publishing; (b) charge density difference plots and value of charge transferred for Ag-doped MoS2 (left) and undoped MoS2 (right) with adsorbed NO2. Blue (yellow) in left and green (red) in right correspond to depletion (excess) of electron density; taken with permission from [159]; Copyright 2013 by the authors.
Nanomaterials 13 02182 g008
Figure 9. (a) Scanning electron microscopy image of 7.9% Sn-doped MoS2 flowers; (b) cycling performance of undoped MoS2 and MoS2 doped with 4.3% (Sn/MoS2-1), 7.9% (Sn/MoS2-2), and 14.9% (Sn/MoS2-3) of tin in LIBs at a current density of 0.2 Ah∙g−1; taken with permission from [189]; Copyright 2016, Springer–Verlag.
Figure 9. (a) Scanning electron microscopy image of 7.9% Sn-doped MoS2 flowers; (b) cycling performance of undoped MoS2 and MoS2 doped with 4.3% (Sn/MoS2-1), 7.9% (Sn/MoS2-2), and 14.9% (Sn/MoS2-3) of tin in LIBs at a current density of 0.2 Ah∙g−1; taken with permission from [189]; Copyright 2016, Springer–Verlag.
Nanomaterials 13 02182 g009
Figure 10. (a) Rate performance of undoped MoS2/carbon cloth (MSC) and Mn-doped (MMSC) electrodes in LIBs; (b) diffusion barrier calculated for Li atoms over Mn-doped MoS2 monolayer and pristine MoS2 monolayer; taken with permission from [194]; Copyright 2019, Elsevier B.V.
Figure 10. (a) Rate performance of undoped MoS2/carbon cloth (MSC) and Mn-doped (MMSC) electrodes in LIBs; (b) diffusion barrier calculated for Li atoms over Mn-doped MoS2 monolayer and pristine MoS2 monolayer; taken with permission from [194]; Copyright 2019, Elsevier B.V.
Nanomaterials 13 02182 g010
Figure 11. (a) HAADF/STEM image and HRTEM image (inset) of Co1/3Mo2/3S2/rGO; (b) cycling performance of Co1/3Mo2/3S2/rGO electrode at various current densities in SIB; taken with permission from [197]; Copyright 2018, Elsevier S.A.
Figure 11. (a) HAADF/STEM image and HRTEM image (inset) of Co1/3Mo2/3S2/rGO; (b) cycling performance of Co1/3Mo2/3S2/rGO electrode at various current densities in SIB; taken with permission from [197]; Copyright 2018, Elsevier S.A.
Nanomaterials 13 02182 g011
Figure 12. (a) Fourier transform magnitudes of EXAFS spectra measured for MoS2, MoSe2, and MoSSe; (b) cycling performance of MoS2, MoSe2, and MoSSe electrodes at various current densities in PIBs; taken with permission from [219]; Copyright 2019, American Chemical Society.
Figure 12. (a) Fourier transform magnitudes of EXAFS spectra measured for MoS2, MoSe2, and MoSSe; (b) cycling performance of MoS2, MoSe2, and MoSSe electrodes at various current densities in PIBs; taken with permission from [219]; Copyright 2019, American Chemical Society.
Nanomaterials 13 02182 g012
Figure 13. (a) CV curves measured under different bending angles of Pt-doped MoS2/carbon cloth electrode at 100 mV s−1; taken with permission from [248]; Copyright 2019, Elsevier B.V.; (b) CV curves of 1T-MnxMo1−xS2−ySey/carbon cloth after bending and twisting of a supercapacitor; taken with permission from [235]; Copyright 2020, Wiley-VCH Verlag GmbH and Co; KGaA.
Figure 13. (a) CV curves measured under different bending angles of Pt-doped MoS2/carbon cloth electrode at 100 mV s−1; taken with permission from [248]; Copyright 2019, Elsevier B.V.; (b) CV curves of 1T-MnxMo1−xS2−ySey/carbon cloth after bending and twisting of a supercapacitor; taken with permission from [235]; Copyright 2020, Wiley-VCH Verlag GmbH and Co; KGaA.
Nanomaterials 13 02182 g013
Table 1. Comparison of heteroatom-doped MoS2 gas sensors with the best undoped MoS2 sensors.
Table 1. Comparison of heteroatom-doped MoS2 gas sensors with the best undoped MoS2 sensors.
NanomaterialAnalyteDetection Limit, TemperatureReference
5% Zn-doped MoS2 nanoparticlesO317 ppb, 25 °C[139]
NO2215 ppb, 25 °C
Ni-doped MoS2 nanoparticlesSO2250 ppb, room[141]
7% Ni-doped MoS2 nanoparticlesNO2311 ppb, room[142]
MoSSe nanoparticlesNO270 ppb, 25 °C[143]
N-doped MoS2 nanosheetsNO2125 ppb, 25 °C[144]
Nb-doped MoS2 monolayerNO25 ppm, 100 °C[145]
6% Nb-doped MoS2 monolayerN(CH2CH3)315 ppb, room[146]
mechanically detached MoS2NO800 ppb, room[110]
CVD MoS2 monolayerNO21.4 ppb[117]
CVD MoS2 monolayerNO2
NH3
20 ppb, room
1 ppm, room
[116]
CVD MoS2 filmsNH3300 ppb, room[114]
Edge-enriched MoS2NO2
NH3
20 ppb, 100 °C
2 ppm, room
[122]
MoS2 nanowiresNO24.6 ppb, 60 °C[123]
mechanically detached MoS2N(CH2CH3)3~1 ppm, 20 °C[115]
Table 2. DFT calculated adsorption energy (Eads), charge transfer (ΔQ), and shortest distance (d) for molecules on the surface of heteroatom-doped MoS2.
Table 2. DFT calculated adsorption energy (Eads), charge transfer (ΔQ), and shortest distance (d) for molecules on the surface of heteroatom-doped MoS2.
HeteroatomGas MoleculeEads (eV)ΔQ (e)d (Å)CorrectionReference
metal
FeCO−1.60−0.271.86D2[149]
NO−2.84−0.321.68
O2−2.10−0.70-
NO2−2.29−0.662.11
NH3−1.520.162.12
CoCO−1.71−0.191.80
NO−2.95−0.251.64
O2−1.61−0.54-
NO2−1.79−0.622.04
NH3−1.380.142.07
NiCO−1.69−0.16-
NO−1.88−0.22-
O2−0.98−0.44-
NO2−1.32−0.42-
NH3−1.330.15-
C2H4O−0.79−0.192.13D3[150]
CH2O−1.080.021.85D2[151]
CuCO−1.27−0.02-D2[149]
−1.25−0.081.86D2[152]
NO−1.5−0.22-D2[149]
−1.44−0.261.81D2[152]
O2−1.16−0.46 D2[149]
NO2−1.88−0.64-
NH3−1.470.18-
AgCO−0.79−0.01-
−0.810.022.09D2[153]
NO−0.98−0.20-D2[149]
0.38−0.132.18D2[154]
O2−0.62−0.38-D2[149]
NO2−1.49−0.61-
−1.85−0.442.18D2[154]
−2.83−0.612.35D2[153]
NH3−1.120.14-D2[149]
−1.130.142.26D2[153]
H2S−0.850.192.52
SO2−0.71−0.352.44
CH2O−0.730.012.27
AuCO−1.06−0.02-D2[149]
NO−1.24−0.20-
−0.72−0.112.16D2[154]
O2−0.74−0.39-D2[149]
NO2−1.70−0.54-
−1.60−0.422.08D2[154]
NH3−1.170.23-D2[149]
Au-AgNO−0.55−0.222.17D2[154]
NO2−2.60−0.452.19
RhCO−1.49−0.16-D2[149]
NO−2.74−0.20-
O2−1.29−0.45-
NO2−1.64−0.29-
NH3−1.200.19-
PdCO−1.13−0.06-
NO−1.05−0.14-
O2−0.39−0.17-
NO2−0.77−0.34-
NH3−0.940.11-
C2H4O−0.530.132.31D3[150]
CH2O−0.570.092.13D2[151]
PtCO−1.60−0.08-D2[149]
NO−1.36−0.17-
O2−0.47−0.26-
NO2−0.99−0.31-
NH3−1.080.16-
CH2O−0.73−0.032.1D2[151]
IrCO−2.04−0.16-D2[149]
NO−3.36−0.26-
O2−1.68−0.55-
NO2−2.08−0.40-
NH3−1.330.17-
TiC2H4O−1.230.282.09D3[150]
CH2O−1.59−0.371.86D2[151]
VCO−1.25−0.24-D3[155]
NO2−2.59−0.66-
H2O−0.810.03-
NH3−1.300.1-
NbCO−1.36−0.33-
NO2−3.88−0.69-
H2O−0.920.04-
NH3−1.240.1-
TaCO−1.70−0.35-
NO2−3.64−0.72-
H2O−1.440.05-
NH3−1.710.1-
non-metal
SiC2H4O−1.650.291.85D3[150]
H2S−0.68−0.162.40vdW[156]
SOF2−3.630.831.64
NO2−3.56−0.302.37D3[157]
GeC2H4O−0.240.023.23D3[150]
N2−0.110.013.88D3[158]
O2−0.220.122.52
H2S−0.15−0.013.84
NO2−1.81−0.172.82D3[157]
PC2H4O−0.130.032.85D3[150]
N2−0.060.013.21D3[158]
O2−0.160.063.01
H2S−0.37−0.202.84
NO2−2.49−0.662.42D3[157]
ClC2H4O−0.210.023.26D3[150]
N2−0.090.013.41D3[158]
O2−0.010.302.52
H2S−0.130.012.98
NO2−1.53−0.142.53D3[157]
SeNO2−0.23−0.042.54
Table 3. Performance of anodes from heteroatom-doped MoS2-based nanomaterials in lithium-ion batteries.
Table 3. Performance of anodes from heteroatom-doped MoS2-based nanomaterials in lithium-ion batteries.
NanomaterialInitial Specific Capacity (Current Density)Reversible Capacity
(Current Density)
Reference
N-doped MoS2 nanoparticles1130.8 mAh∙g−1 (0.2 C)800 mAh∙g−1 (0.2 C)
after 40 cycles
[182]
N-doped MoS2 nanoparticles1186 mAh∙g−1 (0.1 C)738 mAh∙g−1 (0.5 C)
after 100 cycles
[183]
N-rGO/MoS2726.9 mAh∙g−1 (1 C)592.7 mAh∙g−1 (1 C)
after 100 cycles
[184]
N-rGO/MoS21725.6 mAh∙g−1 (0.1 A∙g−1)~630 mAh∙g−1 (0.5 A∙g−1)
after 400 cycles
[185]
N, C-doped MoS2 nanoparticles1280 mAh∙g−1 (0.1 A∙g−1)400 mAh∙g−1 (0.1 A∙g−1)
after 60 cycles
[187]
7.9% Sn-doped MoS2 flowers1087 mAh∙g−1 (0.2 A∙g−1)1087 mAh∙g−1 (0.2 A∙g−1)
after 100 cycles
[189]
17.6% V-doped MoS2814 mAh∙g−1 (1 A∙g−1)350 mAh∙g−1 (1 A∙g−1)
after 300 cycles
[190]
MoS1.75Se0.25~1640 mAh∙g−1 (0.1 C)500 mAh∙g−1 (5 C)
after 350 cycles
[191]
meso-MoS1.12Se0.88/rGO~1330 mAh∙g−1 (0.1 A∙g−1)830 mAh∙g−1 (0.1 A∙g−1)
after 150 cycles
[192]
P-doped MoS2/carbon cloth2700 mAh∙g−1 (0.1 A∙g−1)713 mAh∙g−1 (0.5 A∙g−1)
after 500 cycles
[193]
Mn-doped MoS2/carbon cloth~1280 mAh∙g−1 (0.1 A∙g−1)~1130 mAh∙g−1 (0.1 A∙g−1)
after 200 cycles
[194]
Fe-doped MoS2/rGO1671 mAh∙g−1 (0.1 A∙g−1)946 mAh∙g−1 (0.1 A∙g−1)
after 100 cycles
[195]
Co-doped MoS2/rGO1385.3 mAh∙g−1 (0.1 A∙g−1)1223 mAh∙g−1 (0.1 A∙g−1)
after 100 cycles
[196]
Co1/3Mo2/3S2/rGO~1800 mAh∙g−1 (0.2 A∙g−1)1200 mAh∙g−1 (0.2 A∙g−1)
after 700 cycles
[197]
FeCo-doped MoS2/carbon1874.4 mAh∙g−1 (0.1 C)971.2 mAh∙g−1 (5 C)
after 3000 cycles
[198]
Table 4. Performance of anodes from heteroatom-doped MoS2-based nanomaterials in sodium-ion batteries.
Table 4. Performance of anodes from heteroatom-doped MoS2-based nanomaterials in sodium-ion batteries.
NanomaterialInitial Specific Capacity (Current Density)Reversible Capacity
(Current Density)
Reference
N-doped MoS2/C hollow nanostructures972 mAh∙g−1 (0.1 A∙g−1)128 mAh∙g−1 (2 A∙g−1)
after 5000 cycles
[204]
N-doped MoS2 foam1193 mAh∙g−1 (0.02 A∙g−1)312.4 mAh∙g−1 (2 A∙g−1)
after 100 cycles
[205]
N-doped MoS2/C spheres745 mAh∙g−1 (0.1 C = 0.067 A∙g−1)401 mAh∙g−1 (0.2 C)
after 200 cycles
[206]
N-doped MoS2/C@SiOC716.6 mAh∙g−1 (0.05 A∙g−1)~680 mAh∙g−1 (0.1 A∙g−1)
after 200 cycles
[207]
N-MoS2/N-carbon nanotubes658 mAh∙g−1 (0.1 A∙g−1)372.3 mAh∙g−1 (2 A∙g−1)
after 100 cycles
[208]
N-rGO/MoS21100 mAh∙g−1 (0.1 A∙g−1)542 mAh∙g−1 (0.2 A∙g−1)
after 300 cycles
[203]
Sb-doped MoS2/N-carbon~920 mAh∙g−1 (0.1 A∙g−1)253 mAh∙g−1 (1 A∙g−1)
after 2200 cycles
[209]
C@MoS2−xTex@C630.7 mAh∙g−1 (0.2 A∙g−1)365.3 mAh∙g−1 (1 A∙g−1)
after 300 cycles
[211]
Re-doped fullerene-like MoS2~160 mAh∙g−1 (0.1 C = 0.02 A∙g−1)74 mAh∙g−1 (20 C)
after 30 cycles
[212]
Mn-doped MoS2 nanotubes778 mAh∙g−1 (0.1 A∙g−1)160 mAh∙g−1 (1 A∙g−1)
after 1000 cycles
[213]
VMoS2 flowers580.1 mAh∙g−1 (0.1 A∙g−1)451.6 mAh∙g−1 (2 A∙g−1)
after 800 cycles
[214]
orderly layered VMoS2791.8 mAh∙g−1 (0.2 A∙g−1)534.1 mAh∙g−1 (2 A∙g−1)
after 190 cycles
[215]
Co1/3Mo2/3S2/rGO~1050 mAh∙g−1 (0.1 A∙g−1)529 mAh∙g−1 (1 A∙g−1)
after 200 cycles
[197]
Ni-MoS2@porous carbon930.1 mAh∙g−1 (0.1 A∙g−1)337.5 mAh∙g−1 (1 A∙g−1)
after 500 cycles
[210]
FeCo-doped MoS2/carbon982.7 mAh∙g−1 (0.1 C)473.3 mAh∙g−1 (5 C)
after 3000 cycles
[198]
Table 5. Performance of anodes from heteroatom-doped MoS2-based nanomaterials in potassium-ion batteries.
Table 5. Performance of anodes from heteroatom-doped MoS2-based nanomaterials in potassium-ion batteries.
NanomaterialInitial Specific Capacity (Current Density)Reversible Capacity
(Current Density)
Reference
1T Sb-doped MoS2/N-carbon~700 mAh∙g−1 (0.1 A∙g−1)343 mAh∙g−1 (0.1 A∙g−1)
after 100 cycles
[209]
Se-doped MoS2 nanosheets~320 mAh∙g−1 (0.05 A∙g−1)140 mAh∙g−1 (1 A∙g−1)
after 100 cycles
[217]
MoS1.3Se0.7/N,P-carbon nanotubes686 mAh∙g−1 (0.2 A∙g−1)237 mAh∙g−1 (0.5 A∙g−1)
after 300 cycles
[218]
MoS1.5Se0.5/NC~880 mAh∙g−1 (0.2 A∙g−1)531.6 mAh∙g−1 (0.2 A∙g−1)
after 1000 cycles
[216]
MoS1.6Se0.4/N-carbon1671 mAh∙g−1 (0.1 A∙g−1)143.7 mAh∙g−1 (2 A∙g−1)
after 1500 cycles
[221]
Vacancy-rich MoSSe701.6 mAh∙g−1 (0.1 A∙g−1)~310 mAh∙g−1 (2 A∙g−1)
after 1000 cycles
[219]
Te-doped MoS2723.4 mAh∙g−1 (0.1 A∙g−1)301 mAh∙g−1 (2 A∙g−1)
after 1000 cycles
[220]
FeCo-doped MoS2/carbon542.3 mAh∙g−1 (0.1 C)209.8 mAh∙g−1 (5 C)
after 3000 cycles
[198]
Table 6. Supercapacitor performance of heteroatom-doped MoS2-based nanomaterial electrodes.
Table 6. Supercapacitor performance of heteroatom-doped MoS2-based nanomaterial electrodes.
NanomaterialElectrolyteSpecific CapacitanceEnergy DensityPower DensityReference
N-doped MoS21 M H2SO474.4 F∙g−1 at 2 A∙g−1[228]
1T/2H N-doped MoS2 on carbon cloth0.5 M H2SO4410 F∙g−1 at 1 A∙g−1[229]
N,C-doped MoS2 nanoflakes6 M KOH1400 F∙g−1 at 1 A∙g−145 Wh∙kg−1912 W∙kg−1[187]
O-doped MoS2 microspheres1 M KCl744.2 F∙g−1 at 1 A∙g−1[236]
1T/2H O-doped MoS2/graphite foil1 M Na2SO4280 F∙g−1 at 1 A∙g−139.7 Wh∙kg−1450 W∙kg−1[237]
1T-MoSSe6 M KOH36 F∙g−1 at 0.5 A∙g−1~12.1 Wh∙kg−1~842.5 W∙kg−1[238]
MoSSe6 M KOH1020 F∙g−1 at 10 A∙g−151 Wh∙kg−16000 W∙kg−1[239]
Mo0.91W0.09S2/C41 M Na2SO4432.7 F∙g−1 at 1 A∙g−1[240]
Mn-doped MoS2 nanoflowers0.5 M Na2SO4351 F∙g−1 at 1 A∙g−148.9 Wh∙kg−15000 W∙kg−1[230]
binder-free Mn-doped MoS21 M KOH70.37 F∙g−1 at 1 A∙g−13.14 Wh∙kg−14346.35 W∙kg−1[241]
1T Mn-doped MoS21 M KOH980 F∙g−1 at 1 A∙g−1[231]
1T MnxMo1−xS2−ySey/carbon cloth3 M KOH~288 mAh∙g−1 at 1 mA cm−2~69 Wh∙kg−1985 W∙kg−1[235]
Ni-doped MoS21 M Na2SO4291 F∙g−1 at 0.5 A∙g−1[242]
Ni-doped MoS2 microspheres1 M Na2SO4425 F∙g−1 at 5 mV∙s−19 Wh∙kg−10.5 W∙kg−1[243]
6% Ni-doped MoS23 M KOH528.7 F∙g−1 at 1 A∙g−1140.9 Wh∙kg−111,520 W∙kg−1[232]
1T Ni-doped MoS21 M KOH2461.2 F∙g−1 at 1 A∙g−165.96 Wh∙kg−1700 W∙kg−1[231]
Ni-doped MoS26 M KOH285 F∙g−1 at 1 A∙g−14.83 Wh∙kg−12660 W∙kg−1[244]
Fe-doped MoS26 M KOH211 F∙g−1 at 1 A∙g−14.08 Wh∙kg−16000 W∙kg−1[244]
Cu-doped MoS26 M KOH353 F∙g−1 at 1 A∙g−15.58 Wh∙kg−16000 W∙kg−1[244]
Cu-doped MoS2 film1 M Na2SO4502 F∙g−1 at 1 A∙g−1[245]
Co-doped MoS22 M KOH510 F∙g−1 at 1 A∙g−1[234]
Mo0.7Co0.3S2/g-C3N45 M KOH1063.22 F∙g−1 at 0.5 A∙g−1[246]
Mo0.7Co0.3S2 822.1 F∙g−1 at 0.5 A∙g−1[247]
Co-doped MoS2 nanoflowers1 M KOH* 86 F∙g−1 at 1 A∙g−14.3 Wh∙kg−1600 W∙kg−1[233]
1T Co-doped MoS21 M KOH1270.8 F∙g−1 at 1 A∙g−1[231]
flexible Pt-doped MoS21 M Na2SO4250 F∙g−1 at 0.5 A∙g−1 [248]
* obtained for two-electrode system.
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

Bulusheva, L.G.; Semushkina, G.I.; Fedorenko, A.D. Heteroatom-Doped Molybdenum Disulfide Nanomaterials for Gas Sensors, Alkali Metal-Ion Batteries and Supercapacitors. Nanomaterials 2023, 13, 2182. https://doi.org/10.3390/nano13152182

AMA Style

Bulusheva LG, Semushkina GI, Fedorenko AD. Heteroatom-Doped Molybdenum Disulfide Nanomaterials for Gas Sensors, Alkali Metal-Ion Batteries and Supercapacitors. Nanomaterials. 2023; 13(15):2182. https://doi.org/10.3390/nano13152182

Chicago/Turabian Style

Bulusheva, Lyubov G., Galina I. Semushkina, and Anastasiya D. Fedorenko. 2023. "Heteroatom-Doped Molybdenum Disulfide Nanomaterials for Gas Sensors, Alkali Metal-Ion Batteries and Supercapacitors" Nanomaterials 13, no. 15: 2182. https://doi.org/10.3390/nano13152182

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