Giant Optical Second Harmonic Generation in Two ... - ACS Publications

Jul 3, 2017 - Department of Materials Science and Engineering, College of Engineering ... SnSe, and SnS, denoted by MX with M = Ge, Sn and X = Se, S)...
0 downloads 0 Views 28MB Size
Subscriber access provided by NEW YORK UNIV

Communication

Giant Optical Second Harmonic Generation in 2D Multiferroics Hua Wang, and Xiaofeng Qian Nano Lett., Just Accepted Manuscript • DOI: 10.1021/acs.nanolett.7b02268 • Publication Date (Web): 03 Jul 2017 Downloaded from http://pubs.acs.org on July 4, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Nano Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Giant Optical Second Harmonic Generation in 2D Multiferroics Hua Wang† and Xiaofeng Qian †, *

AUTHOR ADDRESS: † Department of Materials Science and Engineering, College of Engineering and College of Science, Texas A&M University, College Station, TX 77843, USA * E-mail: [email protected]

KEYWORDS: 2D Materials; Multiferroics; Second Harmonic Generation; First-Principles Theory; Group IV Monochalcogenides; Nonlinear Optical Properties

ACS Paragon Plus Environment

1

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 33

ABSTRACT

Nonlinear optical properties of materials such as second and higher order harmonic generation play pivotal roles in lasers, frequency conversion, electro-optic modulators and switches, etc. The strength of nonlinear optical responses highly depends on intrinsic crystal symmetry, transition dipole moments, specific optical excitation, and local environment. Using first-principles electronic structure theory, here we predict giant second harmonic generation (SHG) in recently discovered two-dimensional (2D) ferroelectric-ferroelastic multiferroics – group IV monochalcogenides (i.e., GeSe, GeS, SnSe and SnS). Remarkably, the strength of SHG susceptibility in GeSe and SnSe monolayers is more than one order of magnitude higher than that in monolayer MoS2, and two orders of magnitude higher than that in monolayer hexagonal BN. Their extraordinary SHG is dominated by the large residual of two opposite intraband contributions in the SHG susceptibility. More importantly, the SHG polarization anisotropy is strongly correlated with the intrinsic ferroelastic and ferroelectric orders in group IV monochalcogenide monolayers. Our present findings provide a microscopic understanding of the large SHG susceptibility in 2D group IV monochalcogenide multiferroics from first-principles theory and open up a variety of new avenues for 2D ferroelectrics, multiferroics, and nonlinear optoelectronics, for example, realizing active electrical/mechanical switching of ferroic orders in 2D multiferroics and in situ ultrafast optical characterization of local electronic structures using noncontact noninvasive optical SHG techniques.

ACS Paragon Plus Environment

2

Page 3 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Materials under intense optical excitation may generate large nonlinear optical responses,1-3 depending on the intrinsic crystalline symmetry, microscopic transition dipole matrix, and specific frequency and orientation of optical field applied. In centrosymmetric crystals all even order electric susceptibility tensors vanish, as both electric field and electric polarization are polar vectors with odd parity. Noncentrosymmetric materials with large second-order electric susceptibilities such as optical second harmonic generation (SHG),4 linear electro-optic or Pockels effect, optical rectification, and sum frequency generation are highly valuable for a variety of applications such as lasers, frequency conversion, electro-optic modulators and switches, etc. Large SHG and third harmonic generation were recently discovered in a number of two-dimensional (2D) materials, including monolayer/multilayer MoS2,5-9 MoSe2,10-12 WS2,1315

WSe2,13,

16

hexagonal BN (h-BN),6,

17

GaSe18 and InSe.19 All of them belong to the

noncentrosymmetric point group D3h ( 62 ) with effectively only one independent SHG susceptibility tensor element. It has also been observed along the edge of MoS2.20 The SHG responses can be significantly enhanced by electric control21 and resonant excitonic excitation.22 Despite their monolayer or few layer nature, the effective nonlinear SHG susceptibility of h-BN is close to the well-developed lithium niobate nonlinear crystals widely used in integrated optical waveguide, and in MoS2 monolayer it is even one order magnitude higher than that of h-BN and lithium niobate.6 These discoveries have enkindled further investigations of 2D nonlinear optical materials that are free of phase-matching bottleneck and promising for applications in 2D nonlinear optics, e.g. ultrathin nonlinear optical devices and spectroscopies.21

Very recently, group IV monochalcogenides (i.e., GeSe, GeS, SnSe and SnS, denoted by MX with M=Ge, Sn and X=Se, S) were predicted to be 2D ferroelastic-ferroelectric multiferroics23-26 with giant spontaneous electric polarization23,

24, 27

and spontaneous lattice

ACS Paragon Plus Environment

3

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 33

strain23-26 as well as sizable piezoelectricity.24, 28 Ferroelectricity was experimentally observed by Chang et al.29 in an MX’s cousin, SnTe monolayer, with a similar type of spontaneous lattice distortion and ferroelectric polarization. Moreover, these 2D multiferroic materials possess coupled structure-optical properties owing to their correlated noncentrosymmetry and ferroelastic order.24, 30-33 The presence of ferroelectricity necessitates the breaking of inversion symmetry, therefore in principle these 2D multiferroics can have finite second-order nonlinear electric susceptibilities, while specific nontrivial elements depend on the underlying point group symmetry. These MX monolayers belong to the noncentrosymmetric point group C2v (mm2), and have up to five independent SHG susceptibility tensor elements. Quite interestingly, their bulk holds a centrosymmetric point group D2h (mmm), hence has vanishing second-order susceptibilities. In fact, MX with odd number of layers can have finite SHG response, conversely MX with even number of layers have vanishing SHG response like their “infinite” bulk. Nevertheless, neither theoretical nor experimental SHG has been reported for these 2D multiferroic materials.

Here, using first-principles electronic structure theory we show that these group IV monochalcogenides possess giant SHG susceptibility, much higher than that in MoS2 and h-BN monolayers. The large SHG polarization anisotropy is pertinent to the ferroelastic order and noncentrosymmetric point group, and their extraordinary SHG responses are dominated by the large residual of two opposite intraband contributions in the SHG susceptibility governed by interband and intraband Berry connections. Remarkably, the SHG susceptibility in GeSe and SnSe monolayer is one order of magnitude higher than that in MoS2 monolayer, and two orders of magnitude higher than that in h-BN monolayer. In contrast, MoS2 and h-BN monolayers are largely determined by a single intraband term different from 2D MX. The unique ferroelasticity

ACS Paragon Plus Environment

4

Page 5 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

and ferroelectricity in these 2D multiferroics allow direct mechanical switching of lattice orientation and electrical switching of electrical polarization, accompanied with instantaneous direction and phase switching in the SHG susceptibility tensor, respectively. Our present findings not only predict colossal SHG nonlinear optical responses in group IV monochalcogenide monolayers and provide a microscopic understanding from first-principles theory, but also open up a variety of new avenues for 2D ferroelectrics, multiferroics and nonlinear optoelectronics, e.g. achieving active control of their ferroic orders with electrical and/or mechanical stimuli and in-situ ultrafast optical sensing of local electronic structures and domain evolution in these 2D multiferroics with noncontact noninvasive linear and nonlinear optical spectroscopy/imaging. Computational

Methods.

Ground-state

crystal

structures

of

group

IV

monochalcogenide monolayers were calculated by first-principles density-functional theory (DFT)34 implemented in the Vienna Ab initio Simulation Package35 with a plane-wave basis and the projector-augmented wave method.36 Here we used the Perdew-Berke-Ernzerhof (PBE)’s form37 of exchange-correlation functional within the generalized-gradient approximation (GGA)38 and a Monkhorst-Pack k-point sampling for the Brillouin zone (BZ) integration. An energy cutoff of 400 eV for the plane-wave basis and a Monkhorst-Pack k-point sampling of 10×10×1 were applied. Dynamical stability of the crystal structures was confirmed by phonon dispersion from density-functional perturbation theory calculations.39 A separate stand-alone package has been developed in the group to compute the second-order SHG susceptibility tensor40-42 interfaced with first-principles packages. A scissor operator was applied to account for the underestimated band gap in the DFT-GGA calculations. The Kramers-Kronig relation of the calculated SHG susceptibility tensors was verified, satisfying the causality condition. Benchmark calculations were performed to test the convergence with respect to number of bands, k-point

ACS Paragon Plus Environment

5

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 33

sampling, etc. A dense k-point sampling of 72x72x1, 40 electronic bands, and total 1000 frequency grids between the energy range of [-6eV, 6eV] are enough for achieving converged SHG susceptibility tensors, except that 2000 frequency grids between [-12eV, 12eV] were used for h-BN due to large optical gap. Only positive frequency range is presented because of the reality condition. The fundamental frequency  in the denominator of susceptibility tensor

carries a small imaginary smearing factor :  →  + , and =0.05 eV was used in this work.

All symmetry-allowed and symmetry-disallowed SHG tensor elements have been extensively checked for noncentrosymmetric point groups C2v (mm2) and D3h (62). That is, the 23 SHG susceptibility elements for D3h and 20 elements for C2v are zero at all frequencies. Finally, as the thickness of 2D materials is not well-defined, sheet SHG susceptibility tensors (with the unit of pm2/V) are reported in this work. This is because every term of  −2; ,  shown in the Supporting Information requires an integration in the first BZ. In practice, this is achieved by the discrete sum with the dense k-point sampling in the first BZ:  



  → ∑ , where Ω is the volume of unit cell. For 2D materials, Ω becomes ill-defined for arbitrary simulation cell length along plane norm (Lz assuming z is parallel to plane normal). A 

more meaningful definition will be to use ∑ , where A is the area of the 2D plane in the unit  cell, and correspondingly the k-point sampling is on the 2D plane of the first BZ. It essentially leads to the following formula for “sheet” SHG susceptibility tensors we reported in this work, which is similar to the sheet conductance/resistance used for thin-film materials and 2D  materials:  !""# −2; ,  = %&' −2; ,  ∗ )* . Physically it means the ability of SHG

response coming from a single 2D material. When comparing with other nonlinear optical bulk materials, one can define an effective Lz. This is often done by including the thickness of 2D material (+, ) and the van der Waals thickness on both sides of 2D material (approximated by

ACS Paragon Plus Environment

6

Page 7 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

~3.4 Å on each side), that is, )* = 3.4 ∗ 2 + +, . However, it should be noted that this is certainly a rough estimation where the effect of symmetry and interlayer coupling has not taken into account. More detailed discussion can be found in Supporting Information. Additional SHG calculations were performed to check the impact of spin-orbit coupling which turns out to be small for the materials studied in this work, and the results are included in Supporting Information as well.

Results and Discussion. The crystal structure of group IV monochalcogenide monolayers consists of two atomic layers that are puckered along either x or y direction, as their nondistorted parent structure has four-fold rotation symmetry and four mirror planes. A representative 2D GeSe crystal structure is shown in Figure 1(a). Here, the z-axis is oriented along the plane normal of 2D MX. Our previous calculation indicates large spontaneous lattice strain η in MX monolayers depicted by a transformation strain based on the Green-Lagrange strain tensor.24 Spontaneous tensile strain varies from 1% to 14% and the associated spontaneous compressive strain ranges from -1% to -7%. For the GeSe monolayer shown in Figure 1(a), its distortion is along the y direction, and consequently ηyy= 4.1% and ηxx= -2.7%. Such large spontaneous lattice strain has a nontrivial consequence in the electronic structure and crystal symmetry. Particularly, the puckering causes inversion symmetry breaking and results in the noncentrosymmetric polar point group C2v (mm2), consequently these MX monolayers exhibit large spontaneous polarization, i.e. 484, 357, 260, and 181 pC/m for GeS, GeSe, SnS, and SnSe, respectively.24 In addition, the puckering also leads to highly anisotropic electronic band structure, as displayed in Figure 1(b) for the two bands near the Fermi level, namely conduction band minimum (CBM) and valence band maximum (VBM). The calculated DFT-GGA band gap is 1.13 eV at two k-points 1 (kx,ky) = [0, ±0.39] in the fractional coordinate unit (marked in ACS Paragon Plus Environment

7

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 33

Figure 1(b, c)). As GeSe monolayer is a 2D semiconductor, the reduced dimensionality enhances electron-hole Coulomb interaction and increases the density of states at the band edge, and consequently, large excitonic optical absorption is expected.24,

30, 43

Figure 1(c) displays the

quasiparticle band structure of GeSe monolayer using many-body quasiparticle approach44 and first-principles quasiatomic orbital method45, 46 with the exciton level marked by a red line. The computed exciton energy at the 1 point is 1.26 eV, which is only 0.13 eV higher than the DFTGGA gap

The small difference is due to the cancellation of quasiparticle energy correction to

the DFT-GGA gap (~0.5 eV) and the large exciton binding energy (~ 0.4 eV). To correct the underestimated band gap obtained from DFT-GGA calculations, we have used the quasiparticle GW gap and exciton binding energy to correct the optical gap of monolayer MX, MoS2, and BN. This is a reasonable remedy as the GW band structures are similar to DFT band structure for the 2D materials investigated in this work. With the major difference taken into account by the scissor shift in optical band gap, the dipole matrix elements after the correction should be reasonable. Future efforts shall be made to include quasiparticle GW electronic structure and exciton effect for monolayer MX in a computation and memory efficient way. The unique electronic structures and multiferroic nature of monolayer MX, particularly the absence of inversion symmetry, suggests a possibility of large nonlinear optical responses in group IV monochalcogenide monolayers. Under incident electric field 2 with fundamental frequency  , the second-order

nonlinear polarization 32 at frequency 2 is determined by a third-rank electric

susceptibility tensor 4: 56 2 = 67 −2; , 87 8  which is further correlated with the emitting SHG field. Figure 2(a-c) show the calculated sheet SHG susceptibility tensors in GeSe, MoS2, and h-BN monolayers, and the calculated SHG data for GeS, SnSe, and SnS are

ACS Paragon Plus Environment

8

Page 9 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

presented in Figures S1 and S2. Both MoS2 and BN belong to the D3h (62) point group with effectively only one independent nontrivial SHG susceptibility tensor element that satisfies 







9:: = :9: = ::9 = − 999 . Group IV monochalcogenides belong to the point group C2v (mm2) as mentioned earlier, thus they have five independent SHG susceptibility tensor elements: 













9:: , 999 , 9;; , :9: = ::9 , and ;;9 = ;9; . The magnitude of the calculated sheet SHG 

susceptibility tensor element 9:: is 5.16×106 pm2/V at 0.66 eV for monolayer GeSe, 3.02×105 pm2/V at 1.67 eV for monolayer MoS2, and 6.38×104 pm2/V at 3.22 eV h-BN, demonstrating much higher SHG response in GeSe than that in MoS2 and h-BN. Moreover, GeSe has another 



independent SHG tensor element :9: = ::9 with a substantial magnitude of more than 4×106 pm2/V, and all the other three independent elements are much smaller. The computed frequency dependent SHG susceptibility satisfies the Kramers-Kronig relation as shown in Figure S3. It is worth noticing that, despite the fact that the energy gaps are corrected with respect to exciton energies, the computed SHG response is sensitive to the smearing factor, and the independentparticle approximation may also affect the accuracy. Consequently, the absolute magnitude may differ from experiment. Indeed, the calculated SHG response is higher than the measured SHG values MoS2 shown as dots in Figure 2(b).

Nonetheless, our results on MoS2 and h-BN

monolayers are in good agreement with recent works within the independent-particle approximation.47-49 Meanwhile, the same smearing factor was used for all calculations of the six 2D materials studied to keep them on an equal footing. Very interestingly, the fundamental 

frequency for the first large peak of SHG susceptibility 9:: covers a wide range of optical spectrum, i.e. 0.66, 1.18, 0.56, and 1.00 eV for GeSe, GeS, SnSe and SnS, respectively as shown in Figure S2.

ACS Paragon Plus Environment

9

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 33

Nonlinear SHG response can be characterized by shedding linearly-polarized laser beam onto the materials and measuring different polarization component of the outgoing SHG response through an analyzer. By inspecting its angular dependence, e.g. by rotating the samples, their crystallographic orientation and SHG polarization anisotropy and intensity can be determined.5-7 The SHG response in an arbitrarily oriented sample stems from all symmetryallowed SHG nonlinear susceptibility tensor elements taking into account the relative orientation between the incident beam and the crystal lattice via proper Euler angles. Here, we consider a normal incidence geometry for the six 2D materials. As a result, the angular dependent SHG susceptibilities for the point group C2v (mm2) and D3h (62) are given by     ∥ =; >?  = @ :9: + 9:: A sin = cos = + 999 sinG =,     H =; >?  = 9:: cos G = + @ 999 − :9: A cos = sin =,     ∥ =; IG!  = − 9:: sin 3=, H =; IG!  = 9:: cos 3=,

where “ ∥ ” (“ ⊥ ”) stands for the polarization components of the SHG response parallel

(perpendicular) to the polarization of incident electric field 2 at fundamental frequency . “=” is the rotation angle between 2 and the crystal lattice (e.g. the x axis here). For GeSe 

9:: = 18.2 − 515.6 ⋅  × 10Q

monolayer, 10Q

RST U



, 999 = 9.0 – 11.2 ⋅  × 10Q



;9; = −1.1 + 0.1 ⋅  × 10Q 



RST U

RST U





U



,



:9: = ::9 = −232.9 − 233.5 ⋅  ×



, 9;; = −3.6 + 5.3 ⋅  × 10Q

for  = 0.66 eV .

::9 = − 999 = −1.7 − 30.2 ⋅  × 10Q 

RST

RST U

RST U



, and ;;9 = 



For MoS2 monolayer, 9:: = :9: = 

for  = 1.67 eV. For h-BN monolayer, 9:: =

:9: = ::9 = − 999 = −2.8 – 5.7 ⋅  × 10Q

RST U

for  = 3.22 eV . The SHG intensity is

ACS Paragon Plus Environment

10

Page 11 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters





  proportional to [ ∥ =[ and [ H =[ , hence their corresponding polar plots are presented in

Figure 3(a-c) together with the corresponding frequencies and maximum values. The red and 



  blue lines are [ ∥ =[ and [ H =[ , respectively. The results clearly show that the GeSe

monolayer exhibits highly-polarized colossal SHG response with its maximum value located on 

the perpendicular component at = = 0 (corresponding to 9:: ), which is also evident in the formula above. The polar plots of MoS2 monolayer in Figure 3(b) precisely reflect their D3h symmetry, in excellent agreement with several recent experiments.5-7 Since SHG susceptibility varies with fundamental frequency, the SHG polarization anisotropy will change accordingly, as illustrated in Figure S4. The giant SHG susceptibility and polarization anisotropy in MX monolayers can be observed by noncontact noninvasive SHG spectroscopy with lower pump beam intensity, thereby achieving accurate determination of their 2D crystal orientation. `%a −2; ,  contains an interband In general, the total SHG susceptibility \]\^_

`%a `%a contribution ( bc\de , a modification by intraband motion ( bc\e^ , and a modulation by interband `%a motion ( S]f  ,40,

41

that is,

`%a `%a `%a `%a −2; ,  = bc\de \]\^_ + bc\e^ + S]f . These three

contributions can be further classified into six terms, according to the location of the poles (either `%a `%a `%a at  or at 2 ): 6g#h` ≡ 6  + 6? 2 + 6h 2, 6g#"h ≡ "  + " 2, and jk ≡

j . Detailed expressions of each term are given in the Supporting Information. Figure 4(a) 

and 4(b) present the frequency dependent SHG susceptibility 9:: of GeSe monolayer and the contributions from the six terms with imaginary and real part, respectively. Among all the six terms, 6? 2 and 6h 2 are two leading ones responsible for the giant SHG susceptibility in the GeSe monolayer governed by interband and intraband Berry connections. The explicit expressions are given below, which were originally derived by Sipe et al. using the generalized

ACS Paragon Plus Environment

11

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 33

position operator for periodic crystals,50, 51 "

6? 2 = ℏT  6h 2 = Here, €gj =

6 

"



Q 

@−8 ∑gj



o pq? r h s tuqh r ? s tv hmn nn nm nm mm T wnm

{−2 ∑gj'  ℏT Q 

o @|? r h s ~u|er ? s ~A hmn n} }m n} }m T wnm

xw

xw

ymn

nm uw

ymn

nm uw

zA,

z.

∗ ‚∇ „j ‚ is the matrix element of position operator between the  ‚ ƒ „g

cell periodic part of eigenstates n and m: „j ‚ = eub‚ †j ‚. The superscripts “a, b, c”

indicate the Cartesian directions. €gj is equivalent to interband Berry connection when ‡ ≠ ,

which is directly related to the interband velocity matrix element ‰j' via ‰j' = j' €j' . ` Furthermore, jg = j − g , and Šjj =

‹wn ‹ o

is the group velocity of eigenstate m along

Cartesian direction a. The curly brackets indicate the intrinsic permutation symmetry `%a −2; ,  = `a% −2; ,  with respect to the Cartesian directions, that is, 

` % ` % % ` qŠj' Œ'g t =  pŠj' Œ'g + Œj' Š'g v. Finally, jg ≡ g − j , where g is occupation number of state

n according to the Fermi-Dirac distribution. The above two equations clearly show that 6? 2

% a and Šgg ), where represents the susceptibility modified by intraband motion (evident by Šjj % a 6h 2 contains the susceptibility modified by interband motion (evident by Šj' and Š'g ). As

shown in Figure 4(a-b), these two terms have large but opposite contribution. However their sum 6 2 ≡ 6? 2 + 6h 2 remains very large, which accounts for the major part of total SHG 

susceptibility 9:: compared to other terms including interband " 2 , interband "  , intraband χb ω etc.

ACS Paragon Plus Environment

12

Page 13 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

To further understand the origin of the extraordinary SHG susceptibility in MX 

monolayers, we examine the detailed distribution of the total susceptibility 9:: (Figure 5(a,b))

and the two dominant SHG susceptibility terms, 6? 2 (Figure 5(c,d)) and 6h 2 (Figure 

5(e,f)) in the first BZ. The total susceptibility 9:: is unambiguously concentrated around two

valleys where the bandgap is located. One of the two major terms, 6? 2, is distributed at four spots in the first BZ close to but a bit deviated from the two valleys (Figure 5(c,d)), and they are largely governed by VBM and CBM near the Fermi level. The yxx component of 6? 2 can "



‘ u6 hmn hnm z ’:jg ~, nm wnm uw



? 2 = T   |∑gj 8gj x T be rearranged as follows: 6,9:: ℏ Q w

where ’jg ≡

‰SS − ‰cc is the velocity difference between states m and n. Figure 5(g) and Figure 5(h,i) 

‘ u6 hmn hnm z, nm wnm uw

present the vector field of ’jg and the imaginary and real part of x T w The

velocity

difference

plot

shows

that

close

to

the

respectively. 1

points,

’:jg p“: , Λ 9 v = −’:jg p−“: , Λ 9 v, and the similar behavior can be seen for the other term owing to the mirror symmetry with a (001) mirror plane. Therefore, the product of the two vector fields for specific yxx component leads to the result shown in Figure 5(c,d). Even though the 9

: numerator Œgj Œjg is only constrained by the symmetry and strength of interband Berry

  connection, the denominator jg jg − 2 becomes relatively much larger away from the

bandgap at Λ. Furthermore, ’jg vanishes at two 1 points as the velocity of both VBM and CBM

vanishes. All these factors combined give rise to four spots concentrated around two valleys in the first BZ. For 6h 2, a detailed analysis indicates that it primarily arises from n=VBM, m=CBM, and l=VBM-2, CBM+1, and CBM+2, and the results are shown in Figure 5(j,k)   accordingly. Similarly, the denominator jg jg − 2 largely dominate the response and

concentrates it around the band gap at the two 1 points. In summary, our results identify that the

ACS Paragon Plus Environment

13

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 33

giant SHG response in GeSe monolayer is controlled by two intraband contributions 6? 2 and

6h 2 located near the two valleys. Despite the opposite contributions, the residual SHG 

susceptibility tensor elements 9:: remains extraordinary due to the much larger 6h 2 term. It is worth to note that bulk MX belong to centrosymmetric point group mmm, hence their nature forbids SHG response. The difference between monolayer and bulk is therefore solely due to the stacking induced emerging inversion symmetry of individual noncentrosymmetric layers. As a result, multilayer MX materials shall exhibit alternating even-odd behavior. We have carried out first-principles calculations for multilayer GeSe and SnSe and the results are shown in Figure S6 of Supporting Information, which demonstrates that monolayer and trilayer GeSe (SnSe) have very similar SHG strength and shape at low energy range, and they start to deviate more at relatively higher frequency range with additional oscillations and peaks, largely due to the weak van der Waals stacking of multiple layers.

Group IV monochalcogenide monolayers are 2D ferroelastic-ferroelectric multiferroics with a few intrinsic order parameters, include the ferroelectric spontaneous polarization vector (P), the 2nd-rank ferroelastic spontaneous strain tensor (η), and the SHG second-order nonlinear 

susceptibility (4^•– −2; , , a 3rd-rank tensor. These order parameters are intrinsically coupled with each other. As illustrated in Figure 6(a), there are four possible configurations upon ferroelastic and/or ferroelectric transition. Each configuration has a peculiar set of order parameters that can be switched via ferroelastic/ferroelectric phase transition. For example, the structure at the top of Figure 6(a) may experience ferroelastic transition to the one on the right with order parameters evolving from p59 , —99 , 9:: ≠ 0, :99 = 0v to p5: , —:: , 9:: = 0, :99 ≠

ACS Paragon Plus Environment

14

Page 15 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

0v where :99 right = 9:: top. Their SHG susceptibility tensor will change accordingly, displayed in Figure 6(c) and Figure 6(e). A subsequent ferroelectric transition from the right to the left will change it from p5: , —:: , 9:: = 0, :99 ≠ 0v to p−5: , —:: , 9:: = 0, − :99 ≠ 0v, and its corresponding SHG susceptibility is shown in Figure 6(b). The switching of all the above order parameters is simply due to its inherent symmetry including lattice, sites, and electric polarization. Very importantly, our results demonstrate that the SHG susceptibility in four ferroelectric-ferroelastic multiferroic states are distinctly different from each other, and hence the SHG intensity and phase information52 can be adopted as a unique noncontact measure for accessing multiferroic states. Although linear optical absorption may be used to determine the orientation of crystal lattice based on the anisotropic optical response,24 it cannot distinguish different ferroelectric polarization. The great sensitivity of SHG on the inversion symmetry breaking and the possibility of obtaining the SHG phase information are therefore highly advantageous.

Conclusions.

In summary, our present findings provide a theoretical prediction of

extraordinary SHG nonlinear optical responses and a microscopic understanding in group IV monochalcogenide monolayers. More importantly, it will open up a variety of new opportunities for 2D ferroelectrics, multiferroics, and nonlinear optoelectronics. For example, it could be of fundamental interest to see the effect of photostriction31 on the first and higher-order optical responses described hereby. Moreover, as the ultimate nanometer thickness of 2D materials is much smaller than their coherent wavelength, these 2D nonlinear optical materials will not suffer from the crucial phase-matching problem generally encountered in 3D nonlinear optical crystals. The extraordinary nonlinear responses together with excellent phase-matching make 2D multiferroics particularly promising for nonlinear optoelectronic devices.53 Very excitingly, the

ACS Paragon Plus Environment

15

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 33

unique ferroelasticity and ferroelectricity in these 2D multiferroics, as illustrated in Figure 6(a), allow direct mechanical switching of lattice orientation and electrical switching of electrical polarization, accompanied with instantaneous direction and sign switching of the giant SHG susceptibilities. One can envisage, for example, to achieve active control of the ferroic orders in these 2D multiferroics with electrical, optical and/or mechanical stimuli while in-situ characterizing their local electronic structures and domain evolution using noncontact noninvasive ultrafast optical SHG techniques.

ACS Paragon Plus Environment

16

Page 17 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

ASSOCIATED CONTENT Supporting Information Available The Supporting Information is available free of charge on the ACS Publications website at http://pubs.acs.org. Calculation details, sheet SHG susceptibilities in all six monolayers, benchmark calculation of the Kramers-Kronig relation, SHG polarization anisotropy at different frequencies, the effect of spin-orbit coupling on SHG, and the results of monolayer and trilayer MX as well as detailed definition of sheet SHG susceptibility are available. AUTHOR INFORMATION Corresponding Author * Email: [email protected] Author Contributions X.Q. conceived the idea and supervised the project. H.W. and X.Q. wrote the computational package for SHG nonlinear susceptibility tensors, performed the calculations, and analyzed the results. X.Q. wrote the initial draft of the manuscript with contributions from H.W. Notes The authors declare no competing financial interests. ACKNOWLEDGMENT H.W. and X.Q. acknowledge the start-up funds from Texas A&M University. Portions of this research were conducted with the advanced computing resources provided by Texas A&M High Performance Research Computing.

ACS Paragon Plus Environment

17

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 33

REFERENCES 1. Shen, Y.-R., The Principles of Nonlinear Optics. 1984. 2. Butcher, P. N.; Cotter, D., The Elements of Nonlinear Optics. Cambridge university press: 1991; Vol. 9. 3. Boyd, R. W., Nonlinear Optics. 3rd ed.; Elsevier Inc.: 2008. 4. Franken, P.; Hill, A. E.; Peters, C.; Weinreich, G. Phys. Rev. Lett. 1961, 7, 118. 5. Kumar, N.; Najmaei, S.; Cui, Q. N.; Ceballos, F.; Ajayan, P. M.; Lou, J.; Zhao, H. Phys. Rev. B 2013, 87, 161403. 6. Li, Y. L.; Rao, Y.; Mak, K. F.; You, Y. M.; Wang, S. Y.; Dean, C. R.; Heinz, T. F. Nano Lett. 2013, 13, 3329-3333. 7. Malard, L. M.; Alencar, T. V.; Barboza, A. P. M.; Mak, K. F.; de Paula, A. M. Phys. Rev. B 2013, 87, 201401. 8. Wang, R.; Chien, H. C.; Kumar, J.; Kumar, N.; Chiu, H. Y.; Zhao, H. ACS Applied Materials & Interfaces 2014, 6, 314-318. 9. Jiang, T.; Liu, H.; Huang, D.; Zhang, S.; Li, Y.; Gong, X.; Shen, Y.-R.; Liu, W.-T.; Wu, S. Nat. Nanotech. 2014, 9, 825-829. 10. Gong, Z. R.; Liu, G. B.; Yu, H. Y.; Xiao, D.; Cui, X. D.; Xu, X. D.; Yao, W. Nat. Commun. 2013, 4, 2053. 11. Kim, D. H.; Lim, D. Journal of the Korean Physical Society 2015, 66, 816-820. 12. Wang, G.; Gerber, I. C.; Bouet, L.; Lagarde, D.; Balocchi, A.; Vidal, M.; Amand, T.; Marie, X.; Urbaszek, B. 2D Mater. 2015, 2, 045005. 13. Zeng, H. L.; Liu, G. B.; Dai, J. F.; Yan, Y. J.; Zhu, B. R.; He, R. C.; Xie, L.; Xu, S. J.; Chen, X. H.; Yao, W.; Cui, X. D. Scientific Reports 2013, 3, 1608. 14. Janisch, C.; Wang, Y. X.; Ma, D.; Mehta, N.; Elias, A. L.; Perea-Lopez, N.; Terrones, M.; Crespi, V.; Liu, Z. W. Scientific Reports 2014, 4, 5530. 15. Ye, Z. L.; Cao, T.; O'Brien, K.; Zhu, H. Y.; Yin, X. B.; Wang, Y.; Louie, S. G.; Zhang, X. Nature 2014, 513, 214-218. 16. Ribeiro-Soares, J.; Janisch, C.; Liu, Z.; Elias, A. L.; Dresselhaus, M. S.; Terrones, M.; Cancado, L. G.; Jorio, A. 2D Mater. 2015, 2, 045015. 17. Kim, C. J.; Brown, L.; Graham, M. W.; Hovden, R.; Havener, R. W.; McEuen, P. L.; Muller, D. A.; Park, J. Nano Lett. 2013, 13, 5660-5665. 18. Karvonen, L.; Saynatjoki, A.; Mehravar, S.; Rodriguez, R. D.; Hartmann, S.; Zahn, D. R. T.; Honkanen, S.; Norwood, R. A.; Peyghambarian, N.; Kieu, K.; Lipsanen, H.; Riikonen, J. Scientific Reports 2015, 5, 10334. 19. Deckoff-Jones, S.; Zhang, J.; Petoukhoff, C. E.; Man, M. K. L.; Lei, S.; Vajtai, R.; Ajayan, P. M.; Talbayev, D.; Madéo, J.; Dani, K. M. Scientific Reports 2016, 6, 22620. 20. Yin, X. B.; Ye, Z. L.; Chenet, D. A.; Ye, Y.; O'Brien, K.; Hone, J. C.; Zhang, X. Science 2014, 344, 488-490. 21. Seyler, K. L.; Schaibley, J. R.; Gong, P.; Rivera, P.; Jones, A. M.; Wu, S. F.; Yan, J. Q.; Mandrus, D. G.; Yao, W.; Xu, X. D. Nat. Nanotech. 2015, 10, 407-411. 22. Wang, G.; Marie, X.; Gerber, I.; Amand, T.; Lagarde, D.; Bouet, L.; Vidal, M.; Balocchi, A.; Urbaszek, B. Phys. Rev. Lett. 2015, 114, 097403. 23. Wu, M.; Zeng, X. C. Nano Lett. 2016, 16, 3236-3241. 24. Wang, H.; Qian, X. 2D Mater. 2017, 4, 015042. 25. Mehboudi, M.; Dorio, A. M.; Zhu, W.; van der Zande, A.; Churchill, H. O.; PachecoSanjuan, A. A.; Harriss, E. O.; Kumar, P.; Barraza-Lopez, S. Nano Lett. 2016, 16, 1704-1712.

ACS Paragon Plus Environment

18

Page 19 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

26. Mehboudi, M.; Fregoso, B. M.; Yang, Y.; Zhu, W.; van der Zande, A.; Ferrer, J.; Bellaiche, L.; Kumar, P.; Barraza-Lopez, S. Phys. Rev. Lett. 2016, 117, 246802. 27. Fei, R.; Kang, W.; Yang, L. Phys. Rev. Lett. 2016, 117, 097601. 28. Fei, R.; Li, W.; Li, J.; Yang, L. Appl. Phys. Lett. 2015, 107, 173104. 29. Chang, K.; Liu, J.; Lin, H.; Wang, N.; Zhao, K.; Zhang, A.; Jin, F.; Zhong, Y.; Hu, X.; Duan, W.; Zhang, Q.; Fu, L.; Xue, Q.-K.; Chen, X.; Ji, S.-H. Science 2016, 353, 274-278. 30. Shi, G.; Kioupakis, E. Nano Lett. 2015, 15, 6926-6931. 31. Haleoot, R.; Paillard, C.; Kaloni, T. P.; Mehboudi, M.; Xu, B.; Bellaiche, L.; BarrazaLopez, S. Phys. Rev. Lett. 2017, 118, 227401. 32. Cook, A. M.; M. Fregoso, B.; de Juan, F.; Coh, S.; Moore, J. E. Nat. Commun. 2017, 8, 14176. 33. Rangel, T.; Fregoso, B. M.; Mendoza, B. S.; Morimoto, T.; Moore, J. E.; Neaton, J. B. arXiv preprint arXiv:1610.06589 2016. 34. Kohn, W.; Sham, L. J. Phys. Rev. 1965, 140, A1133-A1138. 35. Kresse, G.; Furthmüller, J. Phys. Rev. B 1996, 54, 11169-11186. 36. Blöchl, P. E. Phys. Rev. B 1994, 50, 17953-17979. 37. Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865-3868. 38. Langreth, D. C.; Mehl, M. J. Phys. Rev. B 1983, 28, 1809-1834. 39. Baroni, S.; de Gironcoli, S.; Dal Corso, A.; Giannozzi, P. Rev. Mod. Phys. 2001, 73, 515562. 40. Aversa, C.; Sipe, J. E. Phys. Rev. B 1995, 52, 14636-14645. 41. Hughes, J. L. P.; Sipe, J. E. Phys. Rev. B 1996, 53, 10751-10763. 42. Sharma, S.; Dewhurst, J. K.; Ambrosch-Draxl, C. Phys. Rev. B 2003, 67, 165332. 43. Gomes, L. C.; Trevisanutto, P. E.; Carvalho, A.; Rodin, A. S.; Castro Neto, A. H. Phys. Rev. B 2016, 94, 155428. 44. Hedin, L. Phys. Rev. 1965, 139, A796-A823. 45. Qian, X.; Li, J.; Qi, L.; Wang, C. Z.; Chan, T. L.; Yao, Y. X.; Ho, K. M.; Yip, S. Phys. Rev. B 2008, 78, 245112. 46. Qian, X.; Li, J.; Yip, S. Phys. Rev. B 2010, 82, 195442. 47. Guo, G. Y.; Lin, J. C. Phys. Rev. B 2005, 72, 075416. 48. Grüning, M.; Attaccalite, C. Phys. Rev. B 2014, 89, 081102. 49. Wang, C.-Y.; Guo, G.-Y. The Journal of Physical Chemistry C 2015, 119, 13268-13276. 50. Adams, E. N.; Blount, E. I. J. Phys. Chem. Solids 1959, 10, 286-303. 51. Sipe, J. E.; Ghahramani, E. Phys. Rev. B 1993, 48, 11705-11722. 52. Kemnitz, K.; Bhattacharyya, K.; Hicks, J. M.; Pinto, G. R.; Eisenthal, B.; Heinz, T. F. Chem. Phys. Lett. 1986, 131, 285-290. 53. Zhao, M.; Ye, Z.; Suzuki, R.; Ye, Y.; Zhu, H.; Xiao, J.; Wang, Y.; Iwasa, Y.; Zhang, X. Light: Science and Applications 2016, 5, e16131.

ACS Paragon Plus Environment

19

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 33

FIGURES

Figure 1. Atomic configuration and electronic structures of group IV monochalcogenide MX monolayers. (a) Top and side views of 2D MX. Dashed orange lines indicate the primitive cell with spontaneous polarization along the y direction. (b) 2D surface plot of electronic band structure in GeSe monolayer in the first Brillouin zone: CBM (upper) and VBM (lower). “Λ” denotes the location of energy gap in the reciprocal unit cell, 1p“: , “9 v = 0, ž0.39Ÿ in the fractional coordinate unit. (c) Quasiparticle band structure of GeSe monolayer with the quasiparticle band gap, the lowest exciton, and the exciton binding energy labeled in the plot.

ACS Paragon Plus Environment

20

Page 21 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Figure 2. Sheet SHG susceptibilities in GeSe, MoS2, and h-BN monolayers with their magnitude, real, and imaginary component. (a) GeSe monolayer which has seven non-zero susceptibility tensor elements with five independent ones due to its point group C2v. (b) MoS2 monolayer which has four SHG elements with only one independent element due to its point group D3h. (c) h-BN monolayer with only one independent element due to the same point group. Black dots indicate the experimental values.7

ACS Paragon Plus Environment

21

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 33

Figure 3. Polarization anisotropy of sheet SHG susceptibilities in (a) GeSe, (b) MoS2, and (c) BN marked by red and blue line, respectively. “∥” (“⊥”) stands for the polarization components of the SHG response parallel (perpendicular) to the polarization 2 of the incident electric

field. “=” is the rotation angle between 2 and the crystal lattice (i.e. the x axis here for all

three 2D materials). Their maximum values are indicated by the green circles in the polar plot and listed at the bottom.

ACS Paragon Plus Environment

22

Page 23 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Figure 4. Frequency-dependent (a) real and (b) imarginar part of the interband ( " , " 2), intraband ( 6 , 6 2), and modulation ( j ) contributions to the sheet SHG 

susceptibility 9:: . “” and “2” denote the poles of the specific terms. The blue and red dots indicate two major 2 terms from the intraband contribution, 6? 2 and 6h 2. Total real 

and imaginary part of the SHG susceptibility 9:: are marked by black line.

ACS Paragon Plus Environment

23

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 33



Figure 5. The k-point dependent distribution of sheet SHG susceptibility 9:: and 2 intraband terms 6? 2 and 6h 2 in the first BZ of GeSe monolayer. (a, b) Real and imaginary part of 

9:: dominated by the contributions from two equivalent valleys at the two bandgap locations

(marked by Λ). (c-f) Real and imaginary part of 2 intraband terms 6? 2 and 6h 2. (g)

The distribution of velocity difference ’jg ≡ ‰SS − ‰cc between CBM and VBM in the first

BZ. (h-i) Real and imaginary part of non velocity terms in 6? 2. (j, k) Real and imaginary

part of 6h 2 from three dominating sets of bands.

ACS Paragon Plus Environment

24

Page 25 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Nano Letters

Figure 6. Strong coupling among ferroelectric (5, ferroelastic (—) order parameters, and sheet SHG susceptibilities  . (a) Four different multiferroic states in the GeSe monolayer: p−5: , —:: , 9:: = 0, − :99 ≠ 0v , p59 , —99 , 9:: ≠ 0, :99 = 0v , p−59 , —99 , 9:: = 0, − :99 ≠ 0v , and

p5: , —:: , 9:: = 0, :99 ≠ 0v marked by cyan, red, magenta, and green box, respectively. (b-e)

Frequency dependent sheet SHG susceptibilities 9:: (solid line) and :99 (dashed line) of the above four different multiferroic states in the unit of 104 pm2/V.

ACS Paragon Plus Environment

25

Nano Letters

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 33

For Table of Contents Only

ACS Paragon Plus Environment

26

(a)

(b)

Page 27 of 33

Y

(c)

R

GeSe

5 4

X

3

x

Λ Λ

1 E QP g 0

-1

binding

E exc E exc Λ

-2 -3

(Ge, Sn) (S, Se)

Energy (eV)

2

y z

Λ Γ

Energy (eV)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31

GeSe

Nano Letters

kx ACS Paragon Plus Environment

ky

-4 -5 Γ

Y

R

X

Γ

R

|χ (2)| (-2 ω; ω, ω) (10 4 pm 2 /V) Re χ (2) (-2 ω; ω, ω)

Im χ (2) (-2 ω; ω, ω)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

(a)

GeSe χ (2) yxx

500 400

χ (2) yyy

40

χ (2) = χ (2) xyx xxy

20

χ (2) yzz

300 200 0 500

(b)

Nano Letters

(2) (2) (2) χ (2) = χ = χ = χ yxx xyx xxy yyy

2 Malard et al. (2013)

0

0 4

0

0

0 -4

-40

-500 500

40

0

4

0

0 -4

-40

-500 1

2

3

ω (eV)

4

5

Page 28 of 33

(2) (2) (2) χ (2) = χ = χ = χ yxx xyx xxy yyy

6

40

0

(c)

BN

4

χ (2) = χ (2) zzy zyz

100

60

MoS2

6

0

1 Paragon 2 Plus Environment 3 4 ACS

ω (eV)

5

6

0

1

2

3

ω (eV)

4

5

6

Im χ (2) (-2 ω; ω, ω) (10 4 pm 2 /V)

Page 29 of 33

Re χ (2) (-2 ω; ω, ω) (10 4 pm 2 /V)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1000

Nano Letters

(a)

GeSe

500

total

Re χ yxx

Re χ e ( ω)

Re χ e (2 ω) Re χ i ( ω)

0

Re χ i (2 ω)

-500

Re χ i (2 ω)

Re χ m ( ω) r

v Re χ i (2 ω)

-1000 800

(b)

400

Im χ total yxx

Im χ e ( ω)

Im χ e (2 ω) Im χ i ( ω)

0

Im χ i (2 ω) Im χ m ( ω)

Im χ ri (2 ω)

-400

-800 0

Im χ vi (2 ω) 1

2

3

ACS Paragon Plus Environment

ω (eV)

4

5

6

χ v (2ω) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Im

GeSe

i

Nano Letters

MoS2

(g) 0.5

Δmn= vmm-vnn

(m=CBM; n=VBM)

(h)

Im

(i)

Re

×

Λ

(c)

y x 30 )of 33 rmn )/(ωPage -2ω ω2mn ∝ -i(r nm mn

total χ (2)

(a) Im

Λ

0

Re

×

(d) Λ -

Λ χ r (2ω) i

(b) Re ky

0.5 - 0.5

0

χ r (VBM, CBM, VBM-2) i

0.5

+ χ ri

+

(VBM, CBM, CBM+1)

χ r (VBM, CBM, CBM+2) i

Im

(j)

Im

Im

Im

Re

(k)

Re

Re

Re

(e) kx

(f) ACS Paragon Plus Environment

Page 31 of 33 (a) GeSe 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 150° 19 20 21 22 23 180° 24 25 26 27 28 210° 29 30 31 32 33 34 35 36 37 38 max 39

(b) y

E(ω)

θ

120 °

60°

0

330°

|χ |

x

300°

270°

90 °

60 °

150 °

120 °

30 °

180 °

0

210 °

330 ° 240 °

300 °

60 °

150 °

30 °

180 °

0

210 °

330 ° 240 °

300 ° 270 °

270 °

@ ω = 0.66 eV (2) 2

θ

90 °

90°

@ ω = 1.67 eV 13

= 2.66 ×10

4

pm / V

2

E(ω)

y

x

30°

240°

BN

E(ω)

y

θ x

120°

(c)

Letters MoSNano 2

ACS(2) Paragon Plus Environment 2

max |χ

|

= 9.14 ×1010 pm4 / V 2

@ ω = 3.22 eV (2) 2

max |χ

| = 4.06 ×109 pm4 / V 2

χyxx ≠0

1 χxyy =0 2 3 4 5 6 7 8 9 10 11 -12 Px , η xx 13 χyxx =0 14 -χxyy ≠0 15 16 y 17 18 z x 19 20

η xx χyxx =0 χxyy ≠0

+ Px ,

ferroelectric transition ferroelastic transition - Py , η yy

-χyxx ≠0 χxyy =0

Im χ (2) (-2 ω; ω, ω)

+ Py , η yy

400

Re χ (2) (-2 ω; ω, ω)

(a)

400

(b)

(c)

χ (2) Letters Nano yxx χ (2) xyy

0

-400

0

-400 ACS Paragon Plus Environment

0

1

ω (eV)

2

3

χ (2)

yxx

χ (2)

xyy

(d)

χ (2)

yxx χ (2) xyy

(e)

(2) Page 32 χ ofyxx 33 (2) χ xyy

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29

90°

Nano Letters

E(ω)

180 °



270 °

SHG

ω ω

2ω E (eV)

Page 33 of 33

ACS Paragon Plus Environment