Next Article in Journal
Evolution of Mn1−xGexBi2Te4 Electronic Structure under Variation of Ge Content
Previous Article in Journal
Reduced Graphene Oxide Embedded with ZnS Nanoparticles as Catalytic Cathodic Material for Li-S Batteries
Previous Article in Special Issue
Unlocking the Power of Artificial Intelligence: Accurate Zeta Potential Prediction Using Machine Learning
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Second Harmonic Generation in Janus Transition Metal Chalcogenide Oxide Monolayers: A First-Principles Investigation

State Key Laboratory of Information Photonics and Optical Communications, Beijing University of Posts and Telecommunications, Beijing 100876, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(14), 2150; https://doi.org/10.3390/nano13142150
Submission received: 18 June 2023 / Revised: 11 July 2023 / Accepted: 18 July 2023 / Published: 24 July 2023

Abstract

:
Due to the unique optical responses induced by vertical atomic asymmetry inside a monolayer, two-dimensional Janus structures have been conceived as promising building blocks for nanoscale optical devices. In this paper, second harmonic generation (SHG) in Janus transition metal chalcogenide oxide monolayers is systematically investigated by the first-principles calculations. Second-order nonlinear susceptibilities are theoretically determined for Janus MXO (M = Mo/W, X = S/Se/Te) monolayers. The calculated values are comparable in magnitude with Janus MoSSe monolayer. X-M-O symmetry breaking leads to non-zero components in vertical direction, compared with the non-Janus structure. Focusing on the SHG induced by incident light at 1064 nm, polarization-dependent responses of six Janus MXO monolayers are demonstrated. The symmetry of p-polarization changes from six-fold to three-fold with acute incidence angle. Moreover, the effects of biaxial strain on band structures and SHG are further investigated, taking MoSO as an exemplary case. We expect these results to bring in recipes for designing nonlinear optical devices based on Janus transition metal chalcogenide oxide monolayers.

1. Introduction

Nonlinear optics, an essential discipline for regulating and controlling light, have made incredible advances in theory and application [1,2,3]. Nonlinear optics are widely used in laser frequency conversion, optical communications, and optical computing [4,5]. However, nonlinear optical devices also face the challenge of integration and miniaturization. With the continuous development of nanotechnology, two-dimensional (2D) materials have shown extraordinary properties in nonlinear optics [6,7]. Mikhailov et al. investigated the nonlinear high-harmonic conversion efficiency and optical properties of graphene from the perspective of the electromagnetics and linear response theory. They discovered that graphene exhibited a significantly higher efficiency in high-harmonic conversion at terahertz frequencies, due to its non-harmonic, charge–carrier dispersion relationship [8]. Cox et al. investigated the nonlinear polarization response of graphene nanosheets near the plasmonic response wavelength of graphene using a combination of a tight-binding model and motion equations of a density matrix. They employed a classical model of nonharmonic oscillators to describe the polarization response under the influence of external optical field intensity. The study revealed that graphene nanodots exhibited several orders of magnitude that were higher nonlinear second-order and third-order polarization responses compared to metallic nanoparticles of similar dimensions [7].
Among the many 2D materials, transition metal dichalcogenides (TMDs) have received much attention. Malard et al. conducted measurements of the second harmonic generation (SHG) spectrum of MoS2 using femtosecond pulses. Their experimental results revealed that the SHG susceptibility of MoS2 monolayer reached a maximum value of 13 × 104 pm2/V when the photonic energy ranged from 2 eV to 4 eV [9]. In the same year, Li et al. experimentally measured the optical response of MoS2 monolayer and a strong SHG was observed [10]. Since then, other studies have come to the same conclusion [11]. Moreover, MoS2 showed an intense nonlinear optical behavior in the third harmonic [12] and four-wave mixing [13,14]. Furthermore, Murray et al. experimentally investigated SHG in WS2 with synthetically or focused ion beam-induced structural defects. The results demonstrated the significant SHG capability of WS2 monolayer and that, despite the potential defects having a relatively minor impact on the second-order nonlinearity, photoluminescence (PL) is significantly affected [15].
However, due to the D3h symmetry of MoS2 (WS2) monolayer, they do not possess χ   ( 2 ) related to the vertical direction. In recent years, the optical properties of two-dimensional materials with broken symmetry structures have gradually attracted attention. Zhang et al. successfully synthesized the Janus MoSSe monomolecular layer based on MoSe2 by employing a controlled sulfide method to make S atoms substitute Se atoms in one layer, while Se atoms in the other layer remained unaltered [16]. The newly synthesized (2D) material breaks the out-of-plane symmetry of TMDs, converting the initial higher symmetry point group of D3h to the lower symmetry point group, C3V, resulting in a host of excellent physical properties. For instance, the built-in dipole moment enhances the wave function separation of electron and hole, reinforcing the electron−phonon interactions, and ultimately leads to the longer exciton radiative recombination lifetime [17]. The Janus MoSSe nanoribbon with spontaneous curling can further reduce symmetry [18,19,20]. Sun et al. conducted quantum transport simulations to investigate the generation of photogalvanic effect (PGE) photocurrent in six Janus transition metal dichalcogenide (TMD) compounds. Under the influence of spontaneous curvature, the device symmetry is reduced from C3v symmetry to Cs symmetry. The PGE increases in the zigzag and armchair directions for Janus TMD photodetectors with non-collinear electrodes [21]. Furthermore, it has been shown that the optical signal can be modified by introducing a different atom into an extended/molecular system [22,23,24]. Mocci et al. reported that the absorption spectra of C24H12 and C32H14 can be modified by substituting Si atoms in them [22]. The second-order nonlinear susceptibility strongly depends on the symmetry of crystals [25]. This means that the inherent low symmetry of Janus TMDs provides a platform for strong SHG. Two-dimensional materials avoid the requirement for phase matching because their crystal thickness is far less than the coherence length [26]. The second harmonic effect of Janus MoSSe had been investigated by first-principles calculations, which discovered the extraordinary nonlinear optical responses of Janus MoSSe with monolayer and bilayer heterostructures [27,28]. Meanwhile, Bian et al. conducted experimental measurements to investigate the power and wavelength dependence of SHG in MoSSe monolayer. By tuning the excitation wavelength, they discovered that the enhancement of SHG coincided with the energy of the C-exciton state in the linear absorption spectrum of monolayer MoSSe. Therefore, it is inferred that the interaction between light and matter is particularly strong when there is resonance between two-photon transitions and the C-exciton state, resulting in a large second-order susceptibility. The Janus MoSSe monolayer is a tunable nonlinear medium in two-dimensional limits, with the potential for on-chip, frequency-doubling applications [29]. And Li et al. investigated the nonlinear optical properties of MoSSe nanosheets using the open-aperture Z-scan technique. The results showed that the nonlinear optical response of MoSSe in the visible range was superior to that in the near-infrared range [30].
Like the Janus TMD system, the phonon spectra of the MXO (M = W/Mo, X = S/Te/Se) monolayer indicates that their structures are stable [31,32], and the MXO monolayers also show excellent optoelectronic properties [31,32,33,34,35]. Varjovi et al. investigated three structural phases (1H, 1T, and 1T′) of Janus WXO (X = S, Se, and Te) monolayers and studied their vibrational, thermal, elastic, piezoelectric, and electronic properties using first-principles methods. The results demonstrated the multifunctional mechanical and electronic properties of Janus WXO monolayers, along with a large piezoelectric response [32]. Recently, Xu et al. oxidized monolayer MoS2 by lowering the concentration of H2O2, which can form the monolayer of Janus MoSO [36]. In addition, Kang et al. and Shioya et al. also proposed the oxidation of single-molecule films WX2 (X = S/Se) using ultraviolet-ozone and laser heating methods, respectively [37,38]. Nguyen et al. conducted a comprehensive study on the tunable electronic and magnetic properties of monolayer MoSO based on density functional theory (DFT) [34]. Additionally, Li et al. identified Janus MoSO monolayer as a potential candidate material for optoelectronic applications [31]. Waheed et al. found that MoSO had an absorption efficiency of up to 90% in the infrared to the ultraviolet region of the spectrum, so MoSO can be an effective candidate for photocatalysis and solar cells [39]. Falahati et al. investigated the optical properties of Janus WSO monolayer using first-principle calculations based on DFT. The findings suggested that the Janus WSO monolayer presents new opportunities for electronic, optoelectronic, and photonic applications [40]. The physical properties of Janus MXOs have been studied in various aspects, but there is a gap in the research of Janus MXOs in the field of nonlinear optics.
In this paper, the second harmonic generation in Janus MXO (M = Mo/W, X = S/Se/Te) monolayers is investigated, based on first-principles calculations. The second-order nonlinear susceptibilities of MXO are obtained using Random-Phase Approximation (RPA) [41]. Our findings suggest that monolayer MXO provides a potential platform for nonlinear optics. Furthermore, we study the biaxial strain effect on the bandgap and SHG of Janus MoSO monolayer. The computational results indicate that second-order nonlinear susceptibility of Janus MoSO monolayer exhibits strain dependence under biaxial strain.

2. Methods

All calculations are performed using the PWmat package, which is a plane wave pseudopotential software based on density-functional theory with GPU acceleration [42,43]. The exchange-correlation function adopts generalized gradient approximation (GGA) of Perdew–Burke–Ernzerhof (PBE) [44]. The spin orbit coupling (SOC) is included in all property calculations. The plane wave cutoff energy of the wave function is set to 70 Ry. The vacuum layer of 30 Å is set in models to avoid the influence of periodic boundary conditions on the calculation. In the primitive cell optimization for Janus MXO, the conjugate gradient (CG) algorithm is used for atomic relaxation [45]. The convergence standard for structural relaxation is set 0.001 eV/Å for atomic forces and less than 0.02 eV/atom for lattice stresses, and van der Waals interactions are considered by the DFT-D3 method [46]. For the structure optimizations and band structure calculations, we set the Monkhost-Pack k-point sampling grid to 15 × 15 × 1. In addition, to avoid the impact of the bandgap underestimation caused by the PBE form, we also adopt the Heyd–Scuseria–Ernzerhof (HSE06) heterogeneous generalization function in the band structure calculations [47].
For calculating nonlinear optical properties, we use the RPA-based method SHG calculation package developed in-house by PWmat [48]. Calculating the nonlinear optical properties requires a dense k-point grid. After testing, we use Monkhost-Pack k-point sampling with a 70 × 70 × 1 grid and 140 electron energy band numbers. We sample 20,000 energy points in the range of 0–6 eV. Methods dealing with bandgap underestimation in the calculation of optical properties include GW, Bethe-Salpeter equation (BSE) and scissor operators, etc. [28,49,50,51]. We adopt the scissor operator to overcome the bandgap underestimation problem caused by the PBE, and the HSE calculation results provide a reference for the correction value of the scissor operator. The expression of second-order nonlinear susceptibility ( χ a b c 2 ω , ω , ω ) was derived by Sharma [52]. Since the calculation method is based on RPA, exciton effects are not included [53]. The second-order susceptibility is contributed by interband transitions, intraband transitions and modulation of interband terms by intraband terms, which can be described as [48,52]
χ inter a b c ( 2 ω , ω , ω ) = e 3 2 n m l d k 4 π 3 r n m a { r m l b r l n c } ( ω l n ω m l ) { 2 f n m ( ω m n 2 ω ) + f m l ( ω m l ω ) + f l n ( ω l n ω ) }
χ intra   a b c ( 2 ω , ω , ω ) = e 3 2 d k 4 π 3 [ n m l ω m n r n m a { r m l b r l n c } { f n l ω l n 2 ( ω l n ω ) f l m ω m l 2 ( ω m l ω ) } 8 i n m f n m r n m a { Δ m n b r m n c } ω m n 2 ( ω m n 2 ω )   + 2 n m l f n m r n m a { r m l b r l n c } ( ω m l ω l n ) ω m n 2 ( ω m n 2 ω ) ]
χ mod a b c ( 2 ω , ω , ω ) = e 3 2 2 d k 4 π 3 [ n m l f n m ω m n 2 ( ω m n ω ) { ω n l r l m a { r m n b r n l c } ω l m r n l a { r l m b r m n c } } + i n m f n m r n m a { r m n b Δ m n c } ω m n 2 ( ω m n ω ) ]
where
r n m a ( k ) = p n m a ( k ) i m ω n m ( k )
Δ n m a ( k ) = v n n a ( k ) v m m a ( k )
In order to avoid the effect of vacuum layer thickness, we multiply the results of the in-plane component ( χ y y y ( 2 ) , χ x x z ( 2 ) ) by a factor φ , which is the ratio of the vacuum layer thickness c to the atomic thickness t. For the out-of-plane components ( χ z x x ( 2 ) , χ z z z ( 2 ) ), we obtain χ a b s ( 2 ) by multiplying χ r e a l ( 2 ) by 1/ φ and χ i m a g ( 2 ) by φ [54,55,56,57].

3. Results and Discussion

As shown in Figure 1, the Janus MXO (M = Mo/W, X = S/Se/Te) monolayer is a hexagonal structure with C3v symmetry. The transition metal atom layer is sandwiched between the oxygen atom layer and the sulfur group atom layer. The theoretically optimized structural parameters are summarized in Table 1. The results show that the thickness of crystals increases with the increase of the X-atom radius, and similarly, the size of the atomic radius affects the lattice constant. The lengths of W-O and Mo-O bond remain almost the same in all systems. Our results are also compared with literatures to verify the reliability of our method. The van der Waals interaction is considered in the calculations. As an example, the optimized lattice parameter of MoSeO is 3.05 Å. Since the radius of the O atom is smaller than that of the Se atom, the lattice constant is smaller than that of MoSe2 [58]. Moreover, we investigate the electronic properties of the Janus MXO family. The Fermi level is set to zero. The band structures from PBE are shown in Figure S1. To reduce the underestimation of PBE bandgap, we resort to HSE06. As shown in Figure 2, the results indicate that all six monolayers are indirect-gap semiconductors, and their bandgaps are between MX2 and MO2. The bandgap of Janus WSO is 1.62 eV, which locates between WO2 (1.52 eV) [58] and WS2 (1.81 eV) [59]. The valance band maximums (VBM) are all located at the Γ point. The conduction band minimum (CBM) of MoTeO is located between Γ and M points, while the CBM of the other five materials is located on the K point. When spin-orbit coupling (SOC) is considered, the valence band edge of MoSO monolayer shows an energy split of 153 meV at the K-high symmetry point, 172 meV for MoSeO monolayer, 193 meV for MoTeO monolayer, 474 meV for WSO monolayer, 473 meV for WSeO monolayer and 447 meV for WTeO monolayer. However, they all split little at the Γ point. In SHG calculation, we use the difference between the bandgap of PBE and HSE06 as a scissor correction. Thus, the scissor operators for the MXO monolayers are 0.56 eV (MoSO), 0.27 eV (MoSeO), 0.47 eV (MoTeO), 0.39 eV (WSO), 0.37 eV (WSeO) and 0.41 eV (WTeO), respectively.
Considering that the crystal symmetry of the Janus MXO monolayers is C3V, there are eleven nonzero components of the second-order nonlinear susceptibility: χ y x x ( 2 ) = χ x x y ( 2 ) = χ x y x ( 2 ) = χ y y y ( 2 ) , χ x x z ( 2 ) = χ x z x ( 2 ) = χ y y z ( 2 ) = χ y z y ( 2 ) , χ z x x ( 2 ) = χ z y y ( 2 ) and χ z z z ( 2 ) , as shown in Figure 3. The largest χ y y y ( 2 ) for the Janus MXO monolayer is caused by materials with minimal bandgap energy and parallel electronic band. The similar trend is also observed in Janus TMDs monolayer [60]. The results show that the principal polarization component χ y y y ( 2 ) generally increases with the increase of atomic number of the sulfur level in the following order: MoSO (WSO) < MoSeO (WSeO) < MoTeO (WTeO). The second-order harmonic response of the Mo-based monolayer is greater than that of W-based monolayer, while the sulfur group elements are identical. The peak position of χ y y y ( 2 ) is red shifted with the increase of the atomic number of X and the maximum of χ y y y ( 2 ) increases. For the Mo-based monolayers, the first peaks of χ y y y ( 2 ) of MoSO, MoSeO, MoTeO locate at 1.46 eV, 1.35 eV and 0.9 eV, respectively. For the W-based monolayers, the first peaks of WSO, WSeO, and WTeO locate at 1.70 eV, 1.46 eV, and 1.02 eV, respectively. Moreover, the Janus structure breaks the out-of-plane symmetry, resulting in larger SHG in the out-of-plane direction, namely χ z x x ( 2 ) , χ z y y ( 2 ) , and χ z z z ( 2 ) . As illustrated in Figure 3a–c, the peak values of χ z x x ( 2 ) of MoSO, MoSeO and MoTeO are 420 pm/V, 900 pm/V, 2346 pm/V, respectively. The susceptibility χ z x x ( 2 ) of MoTeO monolayer is almost five times higher than that of MoSO monolayer. In Figure 3d–f, the peak of susceptibility χ z x x ( 2 ) of W-base is relatively lower than that of Mo-base, with only 364 pm/V, 505 pm/V, 1116 pm/V for WSO, WSeO, and WTeO, respectively. It is worth noting that the charge transfer is not uniform due to the difference in electronegativity of the elements, which results in the largest intensity of the second-order susceptibility in all directions for MoTeO monolayer. The out-of-plane polarization component χ z x x ( 2 ) exceeds the in-plane component χ y y y ( 2 ) when the photon energy reaches about 2 eV, indicating that MoTeO monolayer gives rise to the potential applications for NLO devices. In addition, we compare the results of SHG of MoSSe monolayer calculations with other work [27,28,61,62], as shown in Figure S2. The first peak of the MoSSe monolayer is located near 0.9 eV. The MoSSe monolayer has a maximum magnitude of 1700 pm/V for χ y y y ( 2 ) and 550 pm/V for χ z x x ( 2 ) , which is consistent with what has been reported in other work. In terms of the SHG magnitude of MoXO, MoSO < MoSeO < MoTeO. Based on this, we take MoSSe as a benchmark and further investigated the SHG in the MoSO monolayer and MoTeO monolayer as two representatives, and compared them with MoSSe. Considering a wavelength of 1064 nm laser, Janus MoSO monolayer ( χ y y y ( 2 ) 0.5   × MoSSe, χ z x x ( 2 ) 0.3   ×   M o S S e ) and Janus MoTeO monolayer ( χ y y y ( 2 ) 0.8   × MoSSe, χ z x x ( 2 ) 16   ×   M o S S e ) show excellent out-of-plane responses and in-plane responses in SHG. The Janus transition metal chalcogenide oxide monolayers can play as a promising material platform to investigate SHG.
According to the C3v symmetry of Janus MXO, the two polarization components of nonlinear optical response can be expressed as [27]
I s ( 1 / 2 χ y y y cos 3 ϕ ) 2 I 0 2
I p ( 1 / 2 χ y y y cos θ sin 3 ϕ 1 / 2 χ z y y sin θ ) 2 I 0 2
where θ is the angle between the incident light and the normal line of the monolayer plane, ϕ is the angle between the incident light and the zigzag direction of the in-plane Janus MXO. According to Equation (5), the polarization response is not only related to the polarization angle ϕ but also to the incident angle θ and magnitude of χ z y y ( 2 ) . We set the angle of the incident light as 45° ( θ = 45 ° ). When the wavelength of the incident laser is 1064 nm (1.16 eV), the magnitudes of χ y y y ( 2 ) and χ z y y ( 2 ) for Janus MoTeO monolayer are 5.11 × 102 pm/V and 2.06 × 102 pm/V, respectively. The generated polarizations for all Janus MXO monolayers are shown in Figure 4. The s-polarization has six-fold rotational symmetry while p-polarization possesses triple rotational symmetry. For the Janus MoTeO monolayer, it has the largest SHG intensity of polarization response among the studied Janus MXO. In addition, the response of s-polarization reaches a maximum and the p-polarization response is almost nonexistent for ϕ = 60°, which is determined by the magnitudes of χ y y y ( 2 ) and χ z y y ( 2 ) . It should be noted the intensities of s-polarization and p-polarization response vary at different incident light wavelengths.
Next, the impact of biaxial tensile (compressive) strain is investigated, taking the Janus MoSO monolayer as an exemplary case. In our work, biaxial strain is achieved by adjusting the lattice constants a and b. For structures under different strains, we fix the length of the cell in the c-direction and the lattice constants, and then, relax the positions of all atoms. For example, in the absence of strain, the lattice constants of MoSO monolayer are a0 = b0 = 2.987, and the atomic distance between Se and S is 2.78 Å. When a tensile strain of 2% is applied, we modify the lattice constants to a = b = 3.047, and the atomic distance between Mo and S is reduced to 2.75 Å as a result. The bandgaps of Janus MoSO monolayer under tensile and compressive strain are illustrated in Figure S3. The bandgap decreases considerably when the strain shifts from compression to tension. As can be seen from Figure 5, the bandgap decreases fastest when the strain is −2% to 2%. In addition, we find the CBM locates at the K point when 2% compressive strain is applied to MoSO monolayer. With the compressive strain increasing, the CBM gradually shifts from the K point to a point in K-Γ path and the energy at K point on the conduction band continues to move to higher value. Nevertheless, the VBM remains on Γ point. On the other hand, with the tensile strain increasing, the CBM and VBM remain locate at the K and Γ point, respectively. Meanwhile, the CBM and VBM respectively shift downward and upward, which clearly decreases the bandgap.
We finally calculate the SHG under biaxial strain. In the absence strain, χ y y y ( 2 ) is the dominant component in the Janus MoSO monolayer, and the C3V symmetry of Janus MoSO monolayer yields out-of-plane polarization components. Thus, we mainly calculate the χ y y y ( 2 ) ,   χ z x x ( 2 ) and χ z z z ( 2 ) under strain. As shown in Figure 6, the bandgap of the band structure changes when strain is introduced, which results in a shift of all the peak positions of χ y y y ( 2 ) . For example, in Figure 6a,c, the first peak position shifts 0.1 eV at 2% strain. As the strain increases, the peak shifts more. The magnitude of χ y y y ( 2 ) in the low-energy region increases remarkably with the increase of tensile strain. In Figure 6b,d, as the compressive strain increases, it causes an obvious increase in the magnitude of χ z x x ( 2 ) in photon energy range of 3–4 eV. However, the magnitude of χ z x x ( 2 ) remained essentially constant in the low-energy region with the increase of compressive strain. Interestingly, by applying a 6% compressive strain, we observe that the peak magnitude of the out of plane component χ z x x ( 2 ) at 3.5 eV is three times larger than that of the in-plane component χ y y y ( 2 ) . The other two independent components ( χ x x z ( 2 ) , χ z z z ( 2 ) ) are shown in Figure S4. Applying different strains result in significant changes in second-order nonlinear susceptibility, which exhibits a strong strain dependence.

4. Conclusions

Based on DFT, we investigate the structural and electronic properties of Janus MXO (M = Mo/W, X = S/Se/Te) monolayers and further the second harmonic generation. The results show the out-of-plane symmetry breaking of Janus MXO monolayer leads to non-zero second-order nonlinear susceptibilities in vertical direction. Among these materials, a larger out-of-plane component is produced compare to the Janus MoSSe monolayer. In addition, we calculate the s- and p-polarization of nonlinear optical response of all Janus MXO monolayers. The symmetry of p-polarization changes from six-fold to three-fold with acute incidence angle. We also study the effect of biaxial strain on the band structures and SHG in Janus MoSO monolayer. The bandgap can be tuned by applying biaxial strain and we observe the four independent components of nonlinear susceptibility with different variations. By applying compressive strain, the out of plane component χ z x x ( 2 ) have peaks exceeding the in-plane component χ y y y ( 2 ) within the range of 3–4 eV photon energy. In this sense, the SHG in MoSO monolayer exhibits sensitivity to strain. The results show large SHG in Janus transition metal chalcogenide oxide monolayers and this material family can be a potential platform for nanoscale nonlinear optical devices.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13142150/s1, Figure S1. Band structures of Janus MXO monolayers calculated with PBE + SOC. Figure S2. Calculated second-order nonlinear susceptibility components of MoSSe (a) χ y y y ( 2 ) , (b) χ x x z ( 2 ) , (c) χ z x x ( 2 ) , (d) χ z z z ( 2 ) . Our calculated result (black line) are compared to Ref. [61] in green line, Ref. [28] in blue line, Ref. [27] in pink line, Ref. [60] in yellow line, Ref. [62] in purple line. Figure S3. Bandgap of Janus MoSO monolayer as a function of biaxial strain. Figure S4. Second-order nonlinear susceptibility   χ z z z ( 2 ) for MoSO under (a) tensile strain and (c) compressive strain. Second-order nonlinear susceptibility χ x x z ( 2 ) for MoSO under (b) tensile strain and (d) compressive strain.

Author Contributions

Conceptualization, H.Y.; methodology, H.Y., P.S. and N.S.; software, H.Y.; writing—original draft preparation, P.S. and N.S.; writing—review and editing, H.Y., S.L. and H.Z.; funding acquisition, H.Y. and H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by the National Natural Science Foundation of China (Grant No. 11974003 and 62075015).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dadap, J.I.; Karlsson, M.; Panoiu, N.C. Focus Issue Introduction: Nonlinear Optics 2013. Opt. Express 2013, 21, 31176–31178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Yelin, S.F.; Hemmer, P.R. Resonantly Enhanced Nonlinear Optics in Semiconductor Quantum Wells: An Application to Sensitive Infrared Detection. Phys. Rev. A 2002, 66, 013803. [Google Scholar] [CrossRef] [Green Version]
  3. Levenson, M. The Principles of Nonlinear Optics. IEEE J. Quantum Electron. 1985, 21, 400. [Google Scholar] [CrossRef]
  4. Manykin, E.A.; Melnichenko, E.V. Quantum Computing on Base of Technology of Nonlinear Optical Information Processing. In International Workshop on Quantum Optics 2003; Samartsev, V.V., Ed.; SPIE: Bellingham, WA, USA, 2004; pp. 130–139. [Google Scholar]
  5. Krijnen, G.J.M.; Popma, T.J.A.; Lambeck, P.V.; van Schoot, J.B.P.; Hoekstra, H.J.W.M.; Offrein, B.J.; Driessen, A.; Horst, F. All-Optical Integrated Optic Devices: A Hybrid Approach. IEE Proc.-Optoelectron. 1998, 145, 227–235. [Google Scholar] [CrossRef]
  6. Sipe, J.E.; Ghahramani, E. Nonlinear Optical Response of Semiconductors in the Independent-Particle Approximation. Phys. Rev. B 1993, 48, 11705–11722. [Google Scholar] [CrossRef]
  7. Cox, J.D.; Javier García de Abajo, F. Electrically Tunable Nonlinear Plasmonics in Graphene Nanoislands. Nat. Commun. 2014, 5, 5725. [Google Scholar] [CrossRef] [Green Version]
  8. Mikhailov, S.A. Non-Linear Electromagnetic Response of Graphene. Europhys. Lett. 2007, 79, 27002. [Google Scholar] [CrossRef] [Green Version]
  9. Malard, L.M.; Alencar, T.V.; Barboza, A.P.M.; Mak, K.F.; de Paula, A.M. Observation of Intense Second Harmonic Generation from MoS2 Atomic Crystals. Phys. Rev. B 2013, 87, 201401. [Google Scholar] [CrossRef] [Green Version]
  10. Li, Y.; Rao, Y.; Mak, K.F.; You, Y.; Wang, S.; Dean, C.R.; Heinz, T.F. Probing Symmetry Properties of Few-Layer MoS2 and h-BN by Optical Second-Harmonic Generation. Nano Lett. 2013, 13, 3329–3333. [Google Scholar] [CrossRef] [PubMed]
  11. Lin, K.-Q.; Bange, S.; Lupton, J.M. Quantum Interference in Second-Harmonic Generation from Monolayer WSe2. Nat. Phys. 2019, 15, 242–246. [Google Scholar] [CrossRef] [Green Version]
  12. Wang, R.; Chien, H.-C.; Kumar, J.; Kumar, N.; Chiu, H.-Y.; Zhao, H. Third-Harmonic Generation in Ultrathin Films of MoS2. ACS Appl. Mater. Interfaces 2014, 6, 314–318. [Google Scholar] [CrossRef] [PubMed]
  13. Jakubczyk, T.; Delmonte, V.; Koperski, M.; Nogajewski, K.; Faugeras, C.; Langbein, W.; Potemski, M.; Kasprzak, J. Radiatively Limited Dephasing and Exciton Dynamics in MoSe2 Monolayers Revealed with Four-Wave Mixing Microscopy. Nano Lett. 2016, 16, 5333–5339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Li, D.; Xiong, W.; Jiang, L.; Xiao, Z.; Rabiee Golgir, H.; Wang, M.; Huang, X.; Zhou, Y.; Lin, Z.; Song, J.; et al. Multimodal Nonlinear Optical Imaging of MoS2 and MoS2-Based van Der Waals Heterostructures. ACS Nano 2016, 10, 3766–3775. [Google Scholar] [CrossRef] [PubMed]
  15. Murray, W.; Lucking, M.; Kahn, E.; Zhang, T.; Fujisawa, K.; Perea-Lopez, N.; Laura Elias, A.; Terrones, H.; Terrones, M.; Liu, Z. Second Harmonic Generation in Two-Dimensional Transition Metal Dichalcogenides with Growth and Post-Synthesis Defects. 2D Mater. 2020, 7, 045020. [Google Scholar] [CrossRef]
  16. Zhang, J.; Jia, S.; Kholmanov, I.; Dong, L.; Er, D.; Chen, W.; Guo, H.; Jin, Z.; Shenoy, V.B.; Shi, L.; et al. Janus Monolayer Transition-Metal Dichalcogenides. ACS Nano 2017, 11, 8192–8198. [Google Scholar] [CrossRef] [Green Version]
  17. Zuo, X.; Chang, K.; Zhao, J.; Xie, Z.; Tang, H.; Li, B.; Chang, Z. Bubble-Template-Assisted Synthesis of Hollow Fullerene-like MoS2 Nanocages as a Lithium Ion Battery Anode Material. J. Mater. Chem. A 2016, 4, 51–58. [Google Scholar] [CrossRef]
  18. Sun, N.; Wang, M.; Quhe, R.; Liu, Y.; Liu, W.; Guo, Z.; Ye, H. Armchair Janus MoSSe Nanoribbon with Spontaneous Curling: A First-Principles Study. Nanomaterials 2021, 11, 3442. [Google Scholar] [CrossRef]
  19. Ye, H.; Zhou, J.; Er, D.; Price, C.C.; Yu, Z.; Liu, Y.; Lowengrub, J.; Lou, J.; Liu, Z.; Shenoy, V.B. Toward a Mechanistic Understanding of Vertical Growth of van Der Waals Stacked 2D Materials: A Multiscale Model and Experiments. ACS Nano 2017, 11, 12780–12788. [Google Scholar] [CrossRef]
  20. Ye, H.; Zhang, Y.; Wei, A.; Han, D.; Liu, Y.; Liu, W.; Yin, Y.; Wang, M. Intrinsic-Strain-Induced Curling of Free-Standing Two-Dimensional Janus MoSSe Quantum Dots. Appl. Surf. Sci. 2020, 519, 146251. [Google Scholar] [CrossRef]
  21. Sun, N.; Ye, H.; Quhe, R.; Liu, Y.; Wang, M. Prediction of Photogalvanic Effect Enhancement in Janus Transition Metal Dichalcogenide Monolayers Induced by Spontaneous Curling. Appl. Surf. Sci. 2023, 619, 156730. [Google Scholar] [CrossRef]
  22. Mocci, P.; Cardia, R.; Cappellini, G. Si-Atoms Substitutions Effects on the Electronic and Optical Properties of Coronene and Ovalene. New J. Phys. 2018, 20, 113008. [Google Scholar] [CrossRef]
  23. Pérez-Jiménez, Á.J.; Sancho-García, J.C. Using Circumacenes to Improve Organic Electronics and Molecular Electronics: Design Clues. Nanotechnology 2009, 20, 475201. [Google Scholar] [CrossRef] [PubMed]
  24. Mocci, P.; Cardia, R.; Cappellini, G. Inclusions of Si-Atoms in Graphene Nanostructures: A Computational Study on the Ground-State Electronic Properties of Coronene and Ovalene. J. Phys. Conf. Ser. 2018, 956, 012020. [Google Scholar] [CrossRef]
  25. Shwartz, S.; Weil, R.; Segev, M.; Lakin, E.; Zolotoyabko, E.; Menon, V.M.; Forrest, S.R.; EL-Hanany, U. Light-Induced Symmetry Breaking and Related Giant Enhancement of Nonlinear Properties in CdZnTe:V Crystals. Opt. Express 2006, 14, 9385–9390. [Google Scholar] [CrossRef] [PubMed]
  26. Zhao, M.; Ye, Z.; Suzuki, R.; Ye, Y.; Zhu, H.; Xiao, J.; Wang, Y.; Iwasa, Y.; Zhang, X. Atomically Phase-Matched Second-Harmonic Generation in a 2D Crystal. Light Sci. Appl. 2016, 5, e16131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Wei, Y.; Xu, X.; Wang, S.; Li, W.; Jiang, Y. Second Harmonic Generation in Janus MoSSe a Monolayer and Stacked Bulk with Vertical Asymmetry. Phys. Chem. Chem. Phys. 2019, 21, 21022–21029. [Google Scholar] [CrossRef] [PubMed]
  28. Strasser, A.; Wang, H.; Qian, X. Nonlinear Optical and Photocurrent Responses in Janus MoSSe Monolayer and MoS2–MoSSe van Der Waals Heterostructure. Nano Lett. 2022, 22, 4145–4152. [Google Scholar] [CrossRef] [PubMed]
  29. Bian, C.; Shi, J.; Liu, X.; Yang, Y.; Yang, H.; Gao, H. Optical Second-Harmonic Generation of Janus MoSSe Monolayer. Chin. Phys. B 2022, 31, 097304. [Google Scholar] [CrossRef]
  30. Li, X.; Chen, X.; Wei, N.; Chen, C.; Yang, Z.; Xie, H.; He, J.; Dong, N.; Dan, Y.; Wang, J. Nonlinear Absorption and Integrated Photonics Applications of MoSSe. Opt. Express 2022, 30, 32924. [Google Scholar] [CrossRef]
  31. Li, Y.-Q.; Wang, X.-Y.; Zhu, S.-Y.; Tang, D.-S.; He, Q.-W.; Wang, X.-C. Active Asymmetric Electron-Transfer Effect on the Enhanced Piezoelectricity in MoTO (T = S, Se, or Te) Monolayers and Bilayers. J. Phys. Chem. Lett. 2022, 13, 9654–9663. [Google Scholar] [CrossRef]
  32. Varjovi, M.J.; Yagmurcukardes, M.; Peeters, F.M.; Durgun, E. Janus Two-Dimensional Transition Metal Dichalcogenide Oxides: First-Principles Investigation of W X O Monolayers with X = S, Se, and Te. Phys. Rev. B 2021, 103, 195438. [Google Scholar] [CrossRef]
  33. Van On, V.; Nguyen, D.K.; Guerrero-Sanchez, J.; Hoat, D.M. Exploring the Electronic Band Gap of Janus MoSeO and WSeO Monolayers and Their Heterostructures. New J. Chem. 2021, 45, 20776–20786. [Google Scholar] [CrossRef]
  34. Nguyen, D.K.; Guerrero-Sanchez, J.; Van On, V.; Rivas-Silva, J.F.; Ponce-Pérez, R.; Cocoletzi, G.H.; Hoat, D.M. Tuning MoSO Monolayer Properties for Optoelectronic and Spintronic Applications: Effect of External Strain, Vacancies and Doping. RSC Adv. 2021, 11, 35614–35623. [Google Scholar] [CrossRef]
  35. Yagmurcukardes, M.; Peeters, F.M. Stable Single Layer of Janus MoSO: Strong out-of-Plane Piezoelectricity. Phys. Rev. B 2020, 101, 155205. [Google Scholar] [CrossRef]
  36. Xu, C.-Y.; Qin, J.-K.; Yan, H.; Li, Y.; Shao, W.-Z.; Zhen, L. Homogeneous Surface Oxidation and Triangle Patterning of Monolayer MoS2 by Hydrogen Peroxide. Appl. Surf. Sci. 2018, 452, 451–456. [Google Scholar] [CrossRef]
  37. Shioya, H.; Tsukagoshi, K.; Ueno, K.; Oiwa, A. Selective Oxidation of the Surface Layer of Bilayer WSe2 by Laser Heating. Jpn. J. Appl. Phys. 2019, 58, 120903. [Google Scholar] [CrossRef]
  38. Kang, M.; Yang, H.I.; Choi, W. Oxidation of WS2 and WSe2 Monolayers by Ultraviolet-Ozone Treatment. J. Phys. D Appl. Phys. 2019, 52, 505105. [Google Scholar] [CrossRef]
  39. Waheed, H.S.; Asghar, M.; Ahmad, H.S.; Abbas, T.; Ullah, H.; Ali, R.; Khan, M.J.I.; Iqbal, M.W.; Shin, Y.-H.; Khan, M.S.; et al. Janus MoSO and MoSSe Monolayers: A Promising Material for Solar Cells and Photocatalytic Applications. Phys. Status Solidi (b) 2023, 260, 2200267. [Google Scholar] [CrossRef]
  40. Falahati, K.; Khatibi, A.; Shokri, B. Light-Matter Interaction in Tungsten Sulfide-Based Janus Monolayers: A First-Principles Study. Appl. Surf. Sci. 2022, 599, 153967. [Google Scholar] [CrossRef]
  41. Pešić, J.; Vujin, J.; Tomašević-Ilić, T.; Spasenović, M.; Gajić, R. DFT Study of Optical Properties of MoS2 and WS2 Compared to Spectroscopic Results on Liquid Phase Exfoliated Nanoflakes. Opt. Quantum Electron. 2018, 50, 291. [Google Scholar] [CrossRef]
  42. Jia, W.; Cao, Z.; Wang, L.; Fu, J.; Chi, X.; Gao, W.; Wang, L.-W. The Analysis of a Plane Wave Pseudopotential Density Functional Theory Code on a GPU Machine. Comput. Phys. Commun. 2013, 184, 9–18. [Google Scholar] [CrossRef]
  43. Jia, W.; Fu, J.; Cao, Z.; Wang, L.; Chi, X.; Gao, W.; Wang, L.-W. Fast Plane Wave Density Functional Theory Molecular Dynamics Calculations on Multi-GPU Machines. J. Comput. Phys. 2013, 251, 102–115. [Google Scholar] [CrossRef]
  44. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  45. Hestenes, M.R.; Stiefel, E. Methods of Conjugate Gradients for Solving Linear Systems. J. Res. Natl. Bur. Stand. 1952, 49, 409–436. [Google Scholar] [CrossRef]
  46. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid Functionals Based on a Screened Coulomb Potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
  48. Rashkeev, S.N.; Lambrecht, W.R.L.; Segall, B. Efficient Ab Initio Method for the Calculation of Frequency-Dependent Second-Order Optical Response in Semiconductors. Phys. Rev. B 1998, 57, 3905–3919. [Google Scholar] [CrossRef]
  49. Bernardi, M.; Palummo, M.; Grossman, J.C. Extraordinary Sunlight Absorption and One Nanometer Thick Photovoltaics Using Two-Dimensional Monolayer Materials. Nano Lett. 2013, 13, 3664–3670. [Google Scholar] [CrossRef]
  50. Cappellini, G.; Furthmüller, J.; Cadelano, E.; Bechstedt, F. Electronic and Optical Properties of Cadmium Fluoride: The Role of Many-Body Effects. Phys. Rev. B 2013, 87, 075203. [Google Scholar] [CrossRef] [Green Version]
  51. Cadelano, E.; Furthmüller, J.; Cappellini, G.; Bechstedt, F. One- and Two-Particle Effects in the Electronic and Optical Spectra of Barium Fluoride. J. Phys. Condens. Matter 2014, 26, 125501. [Google Scholar] [CrossRef]
  52. Sharma, S.; Ambrosch-Draxl, C. Second-Harmonic Optical Response from First Principles. Phys. Scr. 2004, T109, 128. [Google Scholar] [CrossRef] [Green Version]
  53. Chang, E.K.; Shirley, E.L.; Levine, Z.H. Excitonic Effects on Optical Second-Harmonic Polarizabilities of Semiconductors. Phys. Rev. B 2001, 65, 035205. [Google Scholar] [CrossRef]
  54. Pike, N.A.; Pachter, R. Second-Order Nonlinear Optical Properties of Monolayer Transition-Metal Dichalcogenides by Computational Analysis. J. Phys. Chem. C 2021, 125, 11075–11084. [Google Scholar] [CrossRef]
  55. Radziuk, D.; Möhwald, H. Ultrasonically Treated Liquid Interfaces for Progress in Cleaning and Separation Processes. Phys. Chem. Chem. Phys. 2016, 18, 21–46. [Google Scholar] [CrossRef]
  56. Laturia, A.; Van De Put, M.L.; Vandenberghe, W.G. Dielectric Properties of Hexagonal Boron Nitride and Transition Metal Dichalcogenides: From Monolayer to Bulk. npj 2D Mater. Appl. 2018, 2, 6. [Google Scholar] [CrossRef] [Green Version]
  57. Osanloo, M.R.; Van De Put, M.L.; Saadat, A.; Vandenberghe, W.G. Identification of Two-Dimensional Layered Dielectrics from First Principles. Nat. Commun. 2021, 12, 5051. [Google Scholar] [CrossRef] [PubMed]
  58. Fang, W.; Xiao, X.; Wei, H.; Chen, Y.; Li, M.; He, Y. The Elastic, Electron, Phonon, and Vibrational Properties of Monolayer XO2 (X = Cr, Mo, W) from First Principles Calculations. Mater. Today Commun. 2022, 30, 103183. [Google Scholar] [CrossRef]
  59. Yakovkin, I. Dirac Cones in Graphene, Interlayer Interaction in Layered Materials, and the Band Gap in MoS2. Crystals 2016, 6, 143. [Google Scholar] [CrossRef] [Green Version]
  60. Pike, N.A.; Pachter, R. Angular Dependence of the Second-Order Nonlinear Optical Response in Janus Transition Metal Dichalcogenide Monolayers. J. Phys. Chem. C 2022, 126, 16243–16252. [Google Scholar] [CrossRef]
  61. Taghizadeh, A.; Thygesen, K.S.; Pedersen, T.G. Two-Dimensional Materials with Giant Optical Nonlinearities near the Theoretical Upper Limit. ACS Nano 2021, 15, 7155–7167. [Google Scholar] [CrossRef]
  62. Lucking, M.C.; Beach, K.; Terrones, H. Large Second Harmonic Generation in Alloyed TMDs and Boron Nitride Nanostructures. Sci. Rep. 2018, 8, 10118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. (a) Janus MXO monolayers with (a) top view and (b) side view.
Figure 1. (a) Janus MXO monolayers with (a) top view and (b) side view.
Nanomaterials 13 02150 g001
Figure 2. Band structures of Janus MXO monolayers calculated with HSE06 + SOC.
Figure 2. Band structures of Janus MXO monolayers calculated with HSE06 + SOC.
Nanomaterials 13 02150 g002
Figure 3. Second-order nonlinear susceptibility χ y y y ( 2 ) (black line), χ x x z ( 2 ) (red line), χ z x x ( 2 ) (blue line), χ z z z ( 2 ) (orange line) for (a) MoSO, (b) MoSeO, (c) MoTeO, (d) WSO, (e) WSeO and (f) WTeO.
Figure 3. Second-order nonlinear susceptibility χ y y y ( 2 ) (black line), χ x x z ( 2 ) (red line), χ z x x ( 2 ) (blue line), χ z z z ( 2 ) (orange line) for (a) MoSO, (b) MoSeO, (c) MoTeO, (d) WSO, (e) WSeO and (f) WTeO.
Nanomaterials 13 02150 g003
Figure 4. s- and p-polarization of nonlinear optical response of monolayer (a) MoSO, (b) MoSeO, (c) MoTeO, (d) WSO, (e) WSeO and (f) WTeO with respect to s-polarized incident light, the wavelengths of all cases are 1064 nm (1.16 eV). The incident angle is 45 ° .
Figure 4. s- and p-polarization of nonlinear optical response of monolayer (a) MoSO, (b) MoSeO, (c) MoTeO, (d) WSO, (e) WSeO and (f) WTeO with respect to s-polarized incident light, the wavelengths of all cases are 1064 nm (1.16 eV). The incident angle is 45 ° .
Nanomaterials 13 02150 g004
Figure 5. Band structures of MoSO monolayer under biaxial strains.
Figure 5. Band structures of MoSO monolayer under biaxial strains.
Nanomaterials 13 02150 g005
Figure 6. Second-order nonlinear susceptibility χ y y y ( 2 ) for MoSO under (a) tensile strain and (c) compressive strain. Second-order nonlinear susceptibility χ z x x ( 2 ) for MoSO under (b) tensile strain and (d) compressive strain.
Figure 6. Second-order nonlinear susceptibility χ y y y ( 2 ) for MoSO under (a) tensile strain and (c) compressive strain. Second-order nonlinear susceptibility χ z x x ( 2 ) for MoSO under (b) tensile strain and (d) compressive strain.
Nanomaterials 13 02150 g006
Table 1. Lattice parameters and bandgaps of Janus MXO monolayers.
Table 1. Lattice parameters and bandgaps of Janus MXO monolayers.
This WorkOther Works
Lattice Parameter (Å)Bandgap
(eV)
Lattice Parameter (Å)Bandgap
(eV)
a, bPBEHSEa, bPBEHSERef.
MoSO2.9871.171.733.0001.091.67[39]
MoSeO3.0500.871.143.0770.821.32[33]
MoTeO3.1720.240.713.1700.250.73[31]
WSO2.9821.622.013.0501.482.06[60]
WSeO3.0421.341.713.0711.321.89[33]
WTeO3.160.550.963.1900.520.97[60]
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

Su, P.; Ye, H.; Sun, N.; Liu, S.; Zhang, H. Second Harmonic Generation in Janus Transition Metal Chalcogenide Oxide Monolayers: A First-Principles Investigation. Nanomaterials 2023, 13, 2150. https://doi.org/10.3390/nano13142150

AMA Style

Su P, Ye H, Sun N, Liu S, Zhang H. Second Harmonic Generation in Janus Transition Metal Chalcogenide Oxide Monolayers: A First-Principles Investigation. Nanomaterials. 2023; 13(14):2150. https://doi.org/10.3390/nano13142150

Chicago/Turabian Style

Su, Peng, Han Ye, Naizhang Sun, Shining Liu, and Hu Zhang. 2023. "Second Harmonic Generation in Janus Transition Metal Chalcogenide Oxide Monolayers: A First-Principles Investigation" Nanomaterials 13, no. 14: 2150. https://doi.org/10.3390/nano13142150

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