Open access peer-reviewed chapter

Quantum Dot Scattering in Monolayer Molybdenum Disulfide

Written By

Rachid Houça, Abdelhadi Belouad, Abdellatif Kamal, El Bouâzzaoui Choubabi and Mohammed El Bouziani

Submitted: 09 May 2022 Reviewed: 07 June 2022 Published: 18 January 2023

DOI: 10.5772/intechopen.105739

From the Edited Volume

Quantum Dots - Recent Advances, New Perspectives and Contemporary Applications

Edited by Jagannathan Thirumalai

Chapter metrics overview

95 Chapter Downloads

View Full Metrics

Abstract

This chapter looks at how electrons propagate in a circular quantum dot (QD) of monolayer molybdenum disulfide (MoS2) that is exposed to an electric potential. Mathematical formulas for the eigenstates, scattering coefficients, scattering efficiency, and radial component of the reflected current and electron density are presented using the continuum model. As a function of physical characteristics such as incident electronic energy, potential barrier, and quantum dot radius, we discover two scattering regimes. We demonstrate the presence of scattering resonances for low-energy incoming electrons. We should also point out that the far-field dispersed current has unique favored scattering directions.

Keywords

  • scattering
  • monolayer molybdenum disulfide
  • quantum dot
  • electric potential
  • electron density

1. Introduction

The hunt for novel two-dimensional materials has been sparked by graphene research [1, 2]. Transition metal dichalcogenides (TMDs) are among them. TMD (MX2 such as M=Mo, W; X=S, Se, and T) monolayers have recently emerged as promising nanostructures for optics, electronics, and spintronic applications. For many years, molybdenum disulfide (MoS2) has been a strong material that has gotten a lot of attention because of its intriguing electrical and optical characteristics [3, 4, 5]. It has a straight bandgap in the visible frequency band [6, 7, 8, 9, 10] and high carrier mobility at ambient temperature [11, 12, 13, 14, 15]. As a result, it is a strong contender for future electrical and optoelectronic technologies.

A strong spin-orbit interaction characterizes the MoS2 monolayer with its honeycomb atomic structure. As a result, totally new electron spin characteristics will emerge. On the one hand, conduction or valence electrons should be significantly less susceptible to the ultrafast spin relaxation effects observed in two-dimensional semiconductor structures such as GaAs quantum wells. The absorption of circularly polarized light, on the other hand, might result in a population of spin-polarized electrons (i.e. an imbalance between the number of spin-up and spin-down electrons). Moreover, the circularly polarized excitation allows control of the distribution of these electrons in one of two valleys of reciprocal space, according to [16]. We call this valley polarization, and we are now working to better understand it so that we can use it in information storage and processing uses. The borders of the valence and conduction bands are positioned at the two corners of the Brillouin zone, i.e. the k and k points, making the MoS2 monolayer a semiconductor. This provides an additional degree of freedom for electrons and holes, which may be employed for encoding information and subsequent processing [16, 17, 18, 19].

Monolayer MoS2 quantum dots (QDs), which feature different physical and chemical characteristics, strong quantum confinement properties, edge effects [20], and a direct bandgap, are key MoS2 related nanostructures that have gotten a lot of interest in recent years. Some many approaches for preparing MoS2 QD have been suggested to date, including solvothermal treatment synthesizing [21], hydrothermal synthesis [22], grinding exfoliation [23], liquid exfoliation in organic solvents [24], electrochemical etching [25], and reaction processing [26].

The contours of the valence and conduction bands determine edge states, which are confined on the boundaries and have energies in the bandgap. The electrical behavior of MoS2 quantum dots was studied using the tight-binding model [22]. According to the orbital asymmetry [22], it was demonstrated that quantum dots with the same form but distinct electrical characteristics may be created.

In this work, we investigate electron propagation in the presence of a potential barrier in a circular electrostatically defined quantum dot monolayer molybdenium disulfide MoS2. Different scattering ranges are identified based on the quantum dot radius, potential barrier, electron spin, and electron energy.

The following is the structure of the present study. We give a theoretical investigation of electron propagation wave plane in a circular quantum dot of monolayer molybdenium disulfide MoS2 in Section 2. The spinors of the Dirac equation solutions corresponding to each area of varied scattering parameters are given. The scattering coefficients are calculated using the continuity criterion. We evaluate the scattering efficiency, square modulus of the scattering coefficients, radial component of the far-field scattered current, and electron density in Section 3, and we describe our findings using various plots. The basic findings of the study are presented in Section 4.

Advertisement

2. Theoretical model

As shown in Figure 1, we investigate a quantum dot of radius ρ in the presence of an electric potential U. In the proximity of the valleys k and k by generating the wave functions via the base of conduction and valence bands, the Dirac-Weyl Hamiltonian for low-energy charge carriers in monolayer molybdenum disulfide (MoS2) yields [27, 28]

H=H0+κ2σz+Δso2τsz1σz+UrE1

Figure 1.

The energy ε of Dirac electrons propagate in a circular quantum dot of monolayer molybdenum disulfide (MoS2). The dot is defined by its radius ρ and the bias introduced U. The incoming plane wave with energy ε>U (blue) belongs to a conduction band state (upper cone). The reflected wave (purple) is now in the conduction band, but the transmitting wave (red) is in the valence band (lower cone).

so the H0 is supplied by:

H0=vFσpE2

where vF, p=pxpy and σ=σxσyσz are, respectively, the Fermi velocity, the 2D momentum operator, and Pauli matrices acting on the atomic orbitals, Ur is the potential barrier, and κ=166.103 eV [28] is related to the material bandgap energy, sz=±1 represents the electron spin-up and spin-down and τ=±1 denotes the k and k valleys, Δso=75.103 eV [28] is the splitting of the valence band owing to spin-orbit coupling, and ρ is the dot radius. For simplicity, the values vF==1 shall be assumed.

We then investigate localized-state solutions in our system, which is characterized as a circularly symmetric quantum dot, utilizing the potential barrier Ur and the energy gap κr defined by:

Ur=0,r>ρU,rρ,κr=0,r>ρκ,rρ.E3

We conduct our research using the polar coordinates (r, ϖ), so that the Hamiltonian (1) has the format:

H=U+ππ+U+τszλsoE4

in which the two potentials and two operators have been established

U±=U±κ2,π±=e±ir±1rϖ.E5

The eigenvalue equation is used to determine the energy spectrum:

Hψlrϖ=εψlrϖ.E6

Because Jz=iφ+σz/2 commutes with H by fulfilling HJz=0. The separability of ψl into the radial Φ±r and angular Ϝ±ϖ parts is required for this commutation, and then we obtain [29, 30]

ψlrϖ=Φl+rϜl+ϖ+Φl+1rϜl+1ϖE7

where the two angular components are

Ϝl+ϖ=eilϖ2π10,Ϝl+1ϖ=eil+1ϖ2π01E8

and l is an integer, which denotes the total angular quantum number.

To obtain the energy spectrum solutions, we must first solve the eigenvalue equation:

Hψlrϖ=εψlrϖE9

by taking into account two areas, as shown in Figure 1, outside (r>ρ) and within (rρ) the quantum dot. As a result, we get an incident wave ψi propagating in the x-direction, an outgoing reflected wave ψr, and a transmitted wave ψt within the dot.

The radial components Φl+r and Φl+1r satisfy two linked differential equations outside the dot (r>ρ):

irΦl+r+ilrΦl+r=εΦl+1rE10
irΦl+1ril+1rΦl+1r=εΦl+r.E11

This may be solved by inserting (10) into (11) to get a second differential equation that Φl+r can fulfill

r222r+rr+r2ε2l2Φl+r=0E12

where the last equation’s solutions are the Bessel functions Jlεr. Furthermore, the incident electron’s wave function moving along the x-direction (x=rcosϖ) has the expression:

ψlirϖ=12lilJlkreilϖ10+iJl+1kreil+1ϖ01E13

as well as the reflected wave:

ψlrrϖ=12lilαlHl1kreilϖ10+iHl+11kreil+1ϖ01E14

wherein Hl1kr is the first class of Hankel function [31], αl is the scattering parameters, and k=ε is the wave number.

The radial functions Φl+ and Φl+1 are represented by the following equations inside the dot (rρ):

irlrΦl+r+εUτszΔsoΦl+1r=0E15
ir+l+1rΦl+1r+εU+Φl+r=0E16

Expressing (15) as

Φl+1r=iεUτszΔsorlrΦl+rE17

By substituting the Eq. (17) in (16), we find a differential equation for Φl+r:

r222r+rr+r2γ2l2Φl+r=0E18

where

γ2=εU+εUτszΔso.E19

The solution of (18) can be worked out to get the transmitted wave as:

ψltrϖ=12lilβlJlγreilϖ10+iμJl+1γreil+1ϖ01E20

where the βl denote the transmission coefficients and

μ=εU+εUτszΔso.E21

Requiring the eigenspinors continuity at the boundary r=ρ of the quantum dot,

ψliρ+ψlrρ=ψltρ,E22

to obtain the conditions

Jl+αlH1=βlJlγρ,E23
Jl+1+αlHl+11=μβlJl+1γρ.E24

Solving these equations to get the scattering coefficients:

αl=JlγρJl+1μJl+1γρJlJlγρHl+11μJl+1γρHl1E25

and the transmission coefficients by:

βl=JlHl+11Jl+1Hl1JlγρHl+11μJl+1γρHl1E26

The density is described as j=ψσψ, with ψ=ψr+ψi outside the dot and ψ=ψt inside the dot.

Angular scattering is described by the far-field radial component of the reflecting current, which giving by

jr=ψcosϖσx+sinϖσyψ=ψ0ee0ψE27

the corresponding radial current can be written as:

jrr=12l=0Alkr×0ee0×l=0AlkrE28

where

Alkr=ilHl1krαleilϖαl1eilϖiHl+11krαl1eil+1ϖαlel+1ϖE29

By injecting the asymptotic behavior of the Hankel function of the first kind for kr1

Hlkr2πkreikr2π4,E30

into the Eq. (29), the density of the system (28) may be simplified to the following form:

jrrϖ=4πkrl=01+cos2l+1ϖcl2E31

where

cl=12αl+12+αl212.E32

The scattering cross section σ is defined by [32]:

σ=IrrIi/AuE33

where Irr and Ii/Au denote, respectively, the total reflected flux across a concentric circle and the incident flux per unit area. Furthermore, Irr is defined by:

Irr=02πrjrrϖdϖ=8kl=0cl2.E34

For the incident wave (13), we note that the incident flux is equal to Ii/Au=1.

To go deeper into our investigation of the scattering problem for a plane Dirac electron at various sizes of the circular quantum dot, we define the scattering efficiency Q by splitting the scattering cross section by the geometric cross section. It is written by:

Q=σ2ρ=4l=0cl2.E35
Advertisement

3. Results and discussions

Figure 2a and b depict the scattering efficiency Q for low energies under barriers scattering (n-p junction) ε=0.01,0.02,0.04,0.06<U=1, in terms of quantum dot radius ρ for the spin-up state in the two valleys k(τ=1, sz=1) and k(τ=1, sz=1) and the spin-down state in the two valleys k(τ=1, sz=1) and k(τ=1, sz=1). We see that when ρ0 implies that Q0, and when ρ rises, Q raises to a maximum value Qmax=20,25,39.5,68 for ε=0.06,0.04,0.02,0.01 for the state (k, sz) (Figure 2a) and for the state (k, -sz) (Figure 2b), then the scattering efficiency Q has a highly damped oscillatory behavior with the appearance of net transverse resonant peaks, comparable to graphene quantum dots [32, 33, 34]. Moreover, the height of the peak decreases as the radius ρ rises, but its breadth increases, which shows the peculiarities of the energy dispersion. Furthermore, by comparing Figure 2a and b, we remark that Q’s dependency on the spin-up and spin-down states in the two valleys is symmetrical, i.e. Qτsz=Qτsz.

Figure 2.

Scattering efficiency Q, for the potential U=1, in terms of the dot radius ρ for different values of ε for the spin-up and spin-down states in two valleys k and k.

Figure 2c and d present, respectively, the scattering efficiency Q for energies across the barrier (n-n junction) ε=1.16,1.2,1.3,1.5>U=1, in terms of quantum dot radius ρ for the spin-up state in the two valleys k and k as well as the spin-down state in valleys k and k. However, when ρ grows, the scattering efficiency Q roughly linearly until it reaches a maximum value Qmax=2.9 relates to a certain value of ρ. Furthermore, by raising ρ and for the four energy levels, the four curves display oscillating attitude [35]. In this domain (ε>U), we find attitudes of symmetry Q (Qτsz=Qτsz) with regard to sz and τ comparable with those of the previous domain (ε<U).

To study the scattering for ε<U in more detail, we present in Figure 3 the scattering efficiency as a function of the electron energy ε. In Figure 3a and b, we consider dots with small radius ρ=0.9,01,1.2,1.3, for the spin-up state in the two valleys k(τ=1) and k(τ=1) and the spin-down state in two valleys k(τ=1) and k(τ=1).

Figure 3.

Scattering efficiency Q, for the potential U=1, in terms of the energy ε of the incident electron for distinct values of ρ for the spin-up and spin-down states in two valleys k and k.

Figure 3a shows that for 0ε0.6, Q exhibits a maximum for the spin-up state in the valley k(τ=1, sz=1) with the emergence of a single peak related to ε=0.55 and a minima for the spin-up state in the valley k(τ, sz=1 without any visible peaks. For E>0.6, we observe that Q has a maximum for the spin-up state in the valley k(τ=1, sz=1) with the emergence of a single peak appropriate for ε=0.75 and a minima for the spin-up state in the valley k(τ=1, sz=1) without the emergence of peaks. For E>0.6, we notice that Q has a maximum for the spin-up state in the valley k(τ=1, sz=1) with the emergence of a single peak appropriate for ε=0.75 and a minima for the spin-up state in the valley k(τ=1, sz=1) without the emergence of peaks.

Figure 3b indicates that for 0ε0.6, Q exhibits a maximum for the spin-down state in the valley k(τ=1, sz=1) with the emergence of a single peak relates to ε=0.55 and a minima for the spin-up state in the valley k(τ=1, sz=1) without the emergence of peaks. We note that for ε>0.6, Q shows a maximum for the spin-down state in the valley k(τ=1, sz=1) with the emergence of a single peak appropriate for ε=0.75 and a minima for the spin-down state in the valley k(τ=1, sz=1) without the emergence of peaks. The electron scattering efficiency, on the other hand, is invariant by the transformation QτszQτsz.

Figure 3c and d represent Q in terms of ε for increasing values of ρ=5,6.25,7,8.25 for the spin-up state in the two valleys k(τ=1, sz=1) and k(τ=1, sz=1) and the spin-down state in two valleys k(τ=1, sz=1) and k(τ=1, sz=1), respectively. Additionally, we notice that Q has significant maxima for low energies as well. However, when ε rises, we see the emergence of a peak with damped oscillations for both spin-up and spin-down states. The resonant excitation of the quantum dot’s normal modes causes these sharp peaks. As a result, the dependency of Q on sz in the two valleys k and k is symmetric with regard to ±sz, i.e Qτsz=Qτsz.

The square modulus of the scattering cl2 is presented in Figure 4 for l=0,1,2,3 in terms of the energy ε, for the spinup and spin-down states in the two valleys k(τ=1) and k(τ=1) for various size of the dot radius: Figure 4a and b: ρ=2, Figure 4c and d: ρ=3, Figure 4e and f: ρ=4 within all panels U=1. Except for the situation corresponding to l=0, all scattering coefficients are zero for zero or near to zero energy. Furthermore, when the energy increases, the scattering coefficients cl2 exhibit oscillatory behavior [35]. We may observe that using the spin-orbit interaction results in an increase in the number of oscillations. Furthermore, we see that some energy values, cl2, have strong peaks. Additionally, the resonances of the dot’s normal modes result in the high peaks previously seen for the scattering efficiency Q in terms of energy (Figure 3). These findings indicate that the term spin-orbit interaction in Eq. (1) needs symmetry cl2τsz=cl2τsz.

Figure 4.

Square modulus of the scattering coefficients cl2 in terms of the energy ε at U=1 for l=0 (blue curve), 1 (red curve), 2 (green curve), and 3 (magenta curve). (a): (k, k, spin-up, ρ=2) states. (b): (k, k, spin-down, ρ=2) states. (c): (k, k, spin-up, ρ=3) states. (d): (k, k, spin-down, ρ=3) states, (e): (k, k, spin-up, ρ=7.75) states, and (f): (k, k, spin-down, ρ=7.75) states. Solid curve relates to valley k and dashed curve relates to valley k.

For the spin-up and spin-down states, we graph the angular characteristic of the reflected radial component jrr in terms of varpi as shown in Figure 5. We note that jrr has a maximum for ϖ=0 and a minimum for ϖ=±π. Furthermore, only forward scattering is preferred for the mode c0 (Figure 5a and b). More favored scattering directions appear for higher modes. As a result, there are three preferred scattering directions for l=1 (Figure 5c and d). Additionally, there are five favored scattering directions for l=2 (Figure 5e and f) and seven preferred scattering directions for l=3 (Figure 5g and h). Generally, each mode has a 2l+1 favored scattering direction that is apparent but with a different amplitude [30], although the mode (l=0) has a larger amplitude than the higher modes (l>0). The electron density profile at the dot reflects resonant scattering by only one of the normal modes. As a result, in both the up and down states, the dependency of jrr on τ is symmetric with respect to ±τ and ±sz, i.e. jrrτsz=jrrτsz.

Figure 5.

Radial component of the far-field scattered current jrr in terms of angle ϖ for various l for the spin-up and spin-down states in two valleys k and k with fixed values ρ=7.75, U=1, and ε=0.0704.

The radial component of the far-field scattered current jrr as a function of incident energy is shown in Figure 6 for the states (a): (k(τ=1), k(τ=1), spin-up), and (b): (k(τ=1), k(τ=1), spin-down) for fixed values U=1 and ρ=4. In all ϖ=0 (red curve) and ϖ=2π/3 (blue curve) [30]. When ε0, jrr for the two values of ϖ, when ε grows to the value 0.5, we see the emergence of peaks of resonances with a maximum peak for ϖτsz=011, as shown in Figure 6a). Although jrr presents oscillatory behavior in the regime 0,5<ε<1.5, in the regime ε1.5, jrr exhibits damped oscillatory behavior with the symmetry jrrϖτsz=jrrϖτsz. Figure 6b indicates that when ττ and szsz, in an analogous method to write jrrϖτsz=jrrϖτsz, the behavior of jrr is identical to that of Figure 6a.

Figure 6.

The radial component of the far-field scattered current jrr in terms of the incident energy ε for various ϖ with fixed values of U=1 and ρ=2 for the spin-up and spin-down states in two valleys k and k.

The electron density profile around the dot reflects the resonant scattering of a single mode. The density inside the quantum dot is described by:

ψtψt=βl2jlγr2+(μjl+1γr2.E36

For the modes β0, β1, β2, and β3, we represent the spatial density ψtψt in the quantum dot, as shown in Figure 7. For both spin-up and spin-down states, the modes βl have a maximum electron density at the center of the quantum dot, and as the size of the quantum dot rises, the electron density drops. We also show that the electron density is becoming significant as the angular monument l is steadily increased. Furthermore, the spin-up and spin-down electron densities are not comparable; more interestingly, the electron density inside the dot has considerably risen, indicating momentary particle entrapment at scattering resonances.

Figure 7.

Spatial density profile ψtψt in the vicinity of the quantum dot for various l for the spin-up and spin-down states in two valleys k and k for fixed values of ρ=3, ε=0.078, and U=1.

Advertisement

4. Summary

The scattering of a planar Dirac electron wave on a circular quantum dot defined electrostatically in the MoS2 monolayer has been investigated. The scattering parameters αl and βl, which characterize the features of our systems, were calculated using the continuity equation at the quantum dot’s borders. The radial component of thev current density, as well as the scattering efficiency and square modulus of the scattering coefficient, were computed. In two energy regimes of the incoming electron, ε<U and εU, the scattering of a planar Dirac electron wave has been examined.

In the regime ε<U, where the incoming electron has a low energy, Q exhibits a damped oscillatory behavior with emergent peaks owing to the excitation of the dot’s normal modes; tiny values of ε correlate with high amplitudes of Q. We have proven that Q exhibits an oscillatory behavior in the other regime εU. On the other hand, we discovered that Qτsz=Qτsz has a remarkable valley symmetry.

To find the resonances, we looked at the energy dependence of the square module of the scattering parameters cl2. We discovered that only the lowest scattering coefficient is non-null near ε0, but as ε increased, the remaining coefficients began to notice significant contributions. The consecutive emergence of modes is interspersed with sudden and sharp peaks of various ε, cl2, but for a not large ε, we have shown that the sequential appearance of modes is interspersed with abrupt and sharp peaks of distinct cl2. We discovered that each mode has (2l+1) preferred paths of scattering visible with distinct amplitudes when it comes to the angular feature of the reflected radial component. Furthermore, we have demonstrated that the density of electrons inside the quantum dot has grown significantly, indicating that electrons have been temporarily trapped during the scattering resonances.

References

  1. 1. Novoselov KS, Geim AK, Morozov SV, Jiang D, Katsnelson MI, Grigorieva IV, et al. Two-dimensional gas of massless Dirac fermions in graphene. Nature. 2005;438:197
  2. 2. Geim AK, Novoselov KS. The rise of grapheme. Nature Materials. 2007;6:183
  3. 3. Mak KF, He K, Shan J, Heinz TF. Control of valley polarization in monolayer MoS2 by optical helicity. Nature Nanotechnology. 2012;7:494
  4. 4. Jones AM, Yu H, Ghimire NJ, Wu S, Aivazian G, Ross JS, et al. Optical generation of excitonic valley coherence in monolayer WSe2. Nature Nanotechnology. 2013;8:634
  5. 5. Zaumseil J. Electronic control of circularly polarized light emission. Science. 2014;344:702
  6. 6. Mak KF, Lee C, Hone J, Shan J, Heinz TF. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Physical Review Letters. 2010;105:136805
  7. 7. Splendiani A, Sun L, Zhang Y, Li T, Kim J, Chim CY, et al. Emerging Photoluminescence in Monolayer MoS2. Nano Letters. 2010;10:1271
  8. 8. Tongay S, Zhou J, Ataca C, Lo K, Matthews TS, Li J, et al. Thermally Driven Crossover from Indirect toward Direct Bandgap in 2D Semiconductors: MoSe2 versus MoS2. Nano Letters. 2012;12:5576
  9. 9. Ross JS, Wu S, Yu H, Ghimire N, Jones A, Aivazian G, et al. Electrical control of neutral and charged excitons in a monolayer semiconductor. Nature Communications. 2013;4:1474
  10. 10. Zeng H, Liu GB, Dai J, Yan Y, Zhu B, He R, et al. Optical signature of symmetry variations and spin-valley coupling in atomically thin tungsten dichalcogenides. Scientific Reports. 2013;3:1608
  11. 11. Radisavljevic B, Radenovic A, Brivio J, Giacometti V, Kis A. Single-layer MoS2 transistors. Nature Nanotechnology. 2011;6:147
  12. 12. Lembke D, Kis A. Breakdown of High-Performance Monolayer MoS2 Transistors. ACS Nano. 2012;6:10070
  13. 13. Lin MW, Liu L, Lan Q, Tan X, Dhindsa KS, Zeng P, et al. Mobility enhancement and highly efficient gating of monolayer MoS2 transistors with polymer electrolyte. Journal of Physics D. 2012;45:345102
  14. 14. Bao W, Cai X, Kim D, Sridhara K, Fuhrer MS. High mobility ambipolar MoS2 field-effect transistors: Substrate and dielectric effects. Applied Physics Letters. 2013;102:042104
  15. 15. Larentis S, Fallahazad B, Tutuc E. Field-effect transistors and intrinsic mobility in ultra-thin MoSe2 layers. Applied Physics Letters. 2012;101:223104
  16. 16. Sallen G, Bouet L, Marie X, Wang G, Zhu CR, Han WP, et al. Robust optical emission polarization in MoS2 monolayers through selective valley excitation. Physical Review B. 2012;86:081301
  17. 17. Xiao D, Yao W, Niu Q. Valley-contrasting physics in graphene: Magnetic moment and topological transport. Physical Review Letters. 2007;99:236809
  18. 18. Yao W, Xiao D, Niu Q. Valley-dependent optoelectronics from inversion symmetry breaking. Physical Review B. 2008;77:235406
  19. 19. Gunawan O, Shkolnikov YP, Vakili K, Gokmen T, De Poortere EP, Shayegan M. Valley susceptibility of an interacting two-dimensional electron system. Physical Review Letters. 2006;97:186404
  20. 20. Wilcoxon J, Samara G. Strong quantum-size effects in a layered semiconductor: MoS2 nanoclusters. Physical Review B. 1995;51:7299
  21. 21. Benson J, Li M, Wang S, Wang P, Papakonstantinou P. Microwave-Assisted Reactant-Protecting Strategy toward Efficient MoS2 Electrocatalysts in Hydrogen Evolution Reaction. ACS Applied Materials & Interfaces. 2015;7:14113
  22. 22. Ren X, Pang L, Zhang Y, Ren XD, Fan H, Liu SZ. One-step hydrothermal synthesis of monolayer MoS2 quantum dots for highly efficient electrocatalytic hydrogen evolution. Journal of Materials Chemistry A. 2015;3:10693
  23. 23. Xu S, Li D, Wu P. One-Pot, Facile, and Versatile Synthesis of Monolayer MoS2/WS2 Quantum Dots as Bioimaging Probes and Efficient Electrocatalysts for Hydrogen Evolution Reaction. Advanced Functional Materials. 2015;25:1136
  24. 24. Gopalakrishnan D, Damien D, Shaijumon M. MoS2 Quantum Dot-Interspersed Exfoliated MoS2 Nanosheets. ACS Nano. 2014;8:5297
  25. 25. Gopalakrishnan D, Damien D, Li B, Gullappalli H, Pillai VK, Ajayan PM, et al. Electrochemical synthesis of luminescent MoS2 quantum dots. Chemical Communications. 2015;51:6293
  26. 26. Li BL, Chen LX, Zou HL, Lei JL, Luo HQ, Li NB. Electrochemically induced Fenton reaction of few-layer MoS2 nanosheets: preparation of luminescent quantum dots via a transition of nanoporous morphology. Nanoscale. 2014;6:9831
  27. 27. Oliveira D, Jiyong F, Villegas-Lelovsky L, Dias AC, Fanyao Q. Valley Zeeman energy in monolayer MoS2 quantum rings: Aharonov-Bohm effect. Physical Review B. 2016;93:205422
  28. 28. Xiao D, Liu GB, Feng W, Xu X, Yao W. Coupled Spin and Valley Physics in Monolayers of MoS2 and Other Group-VI Dichalcogenides. Physical Review Letters. 2012;108:196802
  29. 29. Schnez S, Ensslin K, Sigrist M, Ihn T. Analytic model of the energy spectrum of a graphene quantum dot in a perpendicular magnetic field. Physical Review B. 2008;78:195427
  30. 30. Heinisch RL, Bronold FX, Fehske H. Mie scattering analog in graphene: Lensing, particle confinement, and depletion of Klein tunneling. Physical Review B. 2013;87:155409
  31. 31. Berry MV, Mondragon RJ. Neutrino billiards: time-reversal symmetry-breaking without magnetic fields. Proceedings of the Royal Society. London A. 1987;412:53
  32. 32. Grujic M, Zarenia M, Chaves A, Tadie M, Farias GA, Peeters FM. Electronic and optical properties of a circular graphene quantum dot in a magnetic field: influence of the boundary conditions. Physical Review B. 2011;84:205441
  33. 33. Schulz C, Heinisch RL, Fehske H. Scattering of two-dimensional Dirac fermions on gate-defined oscillating quantum dots. Physical Review B. 2015;91:045130
  34. 34. Belouad A, Zahidi Y, Jellal A, Bahlouli H. Electron scattering in gapped graphene quantum dots. EPL. 2018;28002:123
  35. 35. Xue Wu Z, Long CY, Wu J, Z. Wen Long and T. Xu. Scattering in graphene quantum dots under generalized uncertainty principle. International Journal of Modern Physics A. 2019;4:1950212

Written By

Rachid Houça, Abdelhadi Belouad, Abdellatif Kamal, El Bouâzzaoui Choubabi and Mohammed El Bouziani

Submitted: 09 May 2022 Reviewed: 07 June 2022 Published: 18 January 2023