• Photonics Research
  • Vol. 10, Issue 3, 834 (2022)
Feng Zhou1、2, Cacere Jelah Nieva2, Dianyuan Fan1, Shunbin Lu1、3、*, and Wei Ji1、2、4、*
Author Affiliations
  • 1SZU-NUS Collaborative Innovation Center for Optoelectronic Science & Technology, International Collaborative Laboratory of 2D Materials for Optoelectronic Science & Technology of Ministry of Education, Institute of Microscale Optoelectronics (IMO), Shenzhen University, Shenzhen 518060, China
  • 2Department of Physics, National University of Singapore, Singapore 117542, Singapore
  • 3e-mail: shunbin_lu@szu.edu.cn
  • 4e-mail: phyjiwei@nus.edu.sg
  • show less
    DOI: 10.1364/PRJ.447029 Cite this Article Set citation alerts
    Feng Zhou, Cacere Jelah Nieva, Dianyuan Fan, Shunbin Lu, Wei Ji. Superior optical Kerr effects induced by two-dimensional excitons[J]. Photonics Research, 2022, 10(3): 834 Copy Citation Text show less

    Abstract

    Materials with strong optical Kerr effects (OKEs) are crucial for a broad range of applications, such as all-optical data processing and quantum information. However, the underlying OKE mechanism is not clear in 2D materials. Here, we reveal key insights of the OKE associated with 2D excitons. An admirably succinct formalism is derived for predicting the spectra and the magnitude of the nonlinear refractive index (n2) of 2D materials. The predicted n2 spectra are consistent with reported experimental data and exhibit pronounced excitonic resonances, which is distinctively different from bulk semiconductors. The n2 value is predicted to be 3×10-10 cm2/W for a 2D layered perovskite at low temperature as 7 K, which is four orders of magnitude larger than those of bulk semiconductors. The superior OKE induced by 2D excitons would give rise to a narrow refractive index-near-zero region for intense laser light. Furthermore, we demonstrate that the 2D layered perovskite should exhibit the best OKE efficiency (WFOM=1.02, TFOM=0.14) at 1550 nm, meeting the material requirements for all-optical switching. Our findings deepen the understanding of the OKE of 2D semiconducting materials and pave the way for highly efficient all-optical excitonic devices.

    1. INTRODUCTION

    The optical Kerr effect (OKE) is one of the nonlinear optical (NLO) phenomena observed in dielectric or semiconducting materials, where the refractive index changes in response to intense laser light. The change of refractive index (or so-called nonlinear refractive index, n2) leads to Kerr lensing [1] and phase modulation [2] of the laser light. By utilizing these characteristics, materials with the OKE have shown great potential for applications, such as all-optical switching, all-optical modulation, four wave mixing, optical parametric amplification, and laser mode-locking [38]. In the above-mentioned context, an ideal OKE material should possess large nonlinear refractive index at low optical power. In addition, a short response time is also crucial. Prior to this work, n2 spectra of various 3D dielectrics and semiconductors have been experimentally measured [911] and theoretically verified via a two-parabolic band (TPB) model [12]. Such a theory provides a general guideline on predicting the n2 magnitude of bulk semiconducting materials. However, 3D materials usually show either weak nonlinear refractive change or long response time even under illumination with strong optical fields. Such limitations therefore call for alternative strategies in order to obtain OKE materials for efficient all-optical devices.

    One such strategy is to exploit excitonic effects of 2D materials (or so-called 2D excitons). The 2D exciton can be naturally formed by optical coupling excitonic transitions in layered materials where the spatial confinement and the reduced screening effect give rise to enhanced Coulomb interactions. Since the excitons are confined in a plane that is thinner than their Bohr radius in most 2D semiconductors [1315], quantum confinement enhances the exciton binding energy [16] and alters materials’ optical properties [17]. In particular, the 2D exciton layered materials manifest strong and ultrafast light–matter coupling on nonlinear optical responses in the visible to the near-infrared spectra [18,19]. For instance, a single crystal of Ruddlesden–Popper perovskite (RPP) in the 2D structure has been reported to possess an n2 dispersion with a remarkable resonant feature arising from the Pauli blocking [1] in one-photon-induced transitions to the lowest excitonic state [20], giving rise to considerable OKEs in the visible range. The temporal response for this OKE has been reported to be 100  ps, which is attributed to either an exciton-exciton annihilation (EEA) effect or a radiative recombination process of the 1s excitons [2124]. Apart from the one-photon absorption, OKEs along with the two-photon absorption (2PA) have also been reported in 2D semiconducting materials, where the photon energy (hυ) of the incident light is far below the bandgap (Eg), i.e., Eg/2<hυ<Eg. It is noted that n2 values of layered crystals of RPP, (C4H9NH3)2(CH3NH3)n1PbnI3n+1 (In=1,2,3,4), have shown more than an order of magnitude, as compared with a bulk semiconductor and the TPB model [25] at the excitation wavelength of 2.7 μm. The enhancement of the OKEs was demonstrated to arise from excitonic effects on the 2PA [26,27]. Upon the 2PA associated OKEs, “dark” excitons are created instantaneously and then relax through electron-electron interaction, which exhibits a fast temporal decay <60  fs in transition metal dichalcogenide (TMD) monolayers and RPPs [23,27]. However, a theoretical guideline and systematic understanding on the OKE of 2D layered materials are still lacking.

    In this paper, we develop a simple model on OKEs originated from 2D exciton-associated 2PA. By utilizing the second-order quantum perturbation theory for two-photon transitions among the energy states of 2D excitons and then performing the Kramers–Kronig (K-K) transformation, the nonlinear refractive index (n2) is successfully derived over a broad range of light frequency with an assumption that only 2p exciton is taken into consideration. The theoretically calculated results from our model are comparable with those reported experimental data for monolayer RPPs, TMDs, black phosphorus (BP), and hexagonal boron nitrides (h-BNs). Furthermore, our model explicitly shows that the n2 dispersions of 2D semiconductors are dominated by exciton-resonant features, in particular, the resonance with the 2p exciton; and the n2 magnitudes are closely related to both its linewidth and energy level. The n2 value of a 2D RPP is predicted to be 3×1010  cm2/W, which is four orders of magnitude more than those of bulk semiconductors (1014  cm2/W) [10], demonstrating the superior OKEs induced by 2D excitons and implying a narrow refractive index-near-zero region for intense laser light. Last, we evaluate the figures-of-merit (FOMs) of the OKEs for these 2D semiconductors, which proves them to be of great potential for all-optical excitonic devices.

    2. THEORY

    The nonlinear refractive index (n2) intrinsically accompanied by the 2PA coefficient (β) is governed by the K-K relation [1]. In the present context, we write it in the form n2(hυ)=chπ0+β(hυ,hυ)d(hυ)(hυ)2(hυ)2.

    For a novel K-K transformation, two distinct photon energies for 2PA are needed, i.e., hυ the “cause” and hυ the integration variable in Eq. (1). Here, we consider a special case for the degenerate 2PA, in which the photon energies hυ and hυ are equivalent. The degenerate β in the case of exciton-resonant 2PA of 2D semiconductors is schematized by Fig. 1 and has been calculated based on a second-order, time-dependent, quantum-mechanical perturbation theory [28,29]. Here, we simplify the expression of degenerate β further, and the detailed simplification can be found in the appendix. The simplified expression is as follows: β(hυ)=C2(n02+2)4E2p2F2,exc(hυ),where C2 is a material-related parameter (and its definition can be found in Appendix A); n0 is the linear refractive index. F2,exc(hυ)=hυ(2hυE2p)2+(hγ)2 is the dispersion function with E2p and hγ referring to the energy level of 2p excitonic state and half of its linewidth, respectively. Here, we assume that only 2p excitons make the contribution to the 2PA, and other excitonic states are ignored completely (see Fig. 1).

    Schematic of the optical Kerr effect (self-focusing or -defocusing) induced by 2PA resonant with the exciton energy (E2p). Eg is the bandgap.

    Figure 1.Schematic of the optical Kerr effect (self-focusing or -defocusing) induced by 2PA resonant with the exciton energy (E2p). Eg is the bandgap.

    By substituting Eq. (2) into Eq. (1), the n2 dispersion is derived to be n2(hυ)chC2π(n02+2)4E2p2×0E2p1(E2p2hυ)2+(hγ)2hυ(hυ)2(hυ)2d(hυ).

    Because the equivalence of hυ and hυ would lead to the integral of β(hυ) with divergences, a substitution of υ with υ+Δυ and a numerical summation were made for an approximate integral of β over the 2PA range (0<hυ<E2p). With further simplifications, see the details in Appendix B; we obtain a semiempirical expression as n2(hυ)=Z2(n02+2)4E2p2G(x),where Z2 is another material-related parameter, which is determined by the density (De), the effective Bohr radius (aB), and the linewidth (2hγ) of 2D excitons [24]. The definition of Z2 can be found in Appendix A. The 10 2D materials (see Appendix C) have an average Z2 value of 1×1014 in the SI units such that n2 is in cm2/W, and hυ, E2p, and hγ are in units of eV. G(x) is given by G(x)=0.5x(0.5x)2+0.3(hγ/E2p)2,where x=hυ/E2p.

    3. RESULTS AND DISCUSSION

    Note that Eq. (5) interprets the normalized n2-dispersion. As shown by the pink area in Fig. 2, the normalized n2 values with hγ/E2p being in the range from 0.05 to 0.15 exhibit a central-symmetrical dispersion with respect to x=0.5. For |x0.5|0.25, the normalized n2 values vary from the positive to the negative dramatically, indicating that the OKE would be on (or near) resonance with E2p. Otherwise, the normalized n2 values can be regarded as the off-resonant ones and are nearly independent on x or hγ/E2p.

    Comparison of the normalized n2 between the prediction (pink area) by Eq. (5) and the measured values (symbols) of monolayer (or few-layer) TMDs, RPP(In=1), and h-BN [25,30–34" target="_self" style="display: inline;">–34]. The envelopes of the pink area are calculated with hγ/E2p=0.05 and 0.15.

    Figure 2.Comparison of the normalized n2 between the prediction (pink area) by Eq. (5) and the measured values (symbols) of monolayer (or few-layer) TMDs, RPP(In=1), and h-BN [25,3034" target="_self" style="display: inline;">34]. The envelopes of the pink area are calculated with hγ/E2p=0.05 and 0.15.

    To validate our model, the calculated n2 dispersions are normalized and compared with experimental data of monolayer (or few-layer) TMDs, RPPs, h-BN, and BP [25,3034], where the values of E2p2/[Z2(n02+2)4] are obtained from Table 3 in Appendix A. In the off-resonance region, our model is in agreement with the experimental data within an order of magnitude for RPP (In=1) in polycrystalline structures (at 2.7 μm) [25] and h-BN (at 1.064 μm) [33]. The experimental data of RPP (In=2,3,4) powders are less than the prediction of Eq. (4). It is anticipated that surface states and/or surface scattering from the crystallites may hinder its intrinsic NLO properties of the 2D crystals. The normalized n2 value of BP [34] is 26 at x=hυ/E2p=1.29, which is comparable with the magnitude predicted by Eq. (4), but it is not shown because it is beyond the scale of the x axis in Fig. 2.

    In the near-resonance region, there is a large discrepancy with the experimental data of monolayer (or few-layer) TMDs [3032], which is attributed to the spin-split-off excitonic states. Experimental evidence has shown monolayer TMDs with the spin-orbital splitting possess a pair of generated excitons (as either the A-exciton or B-exciton [35]) transitioning at the K and K valley in the momentum space with broken inversion symmetry. This gives rise to a unique valley degree of freedom in the n2 dispersions that can directly couple to the helicity of excitation photons [17]. Under two-photon excitation, it is necessary to consider the two excitons; hence, as detailed in Appendix A, the n2 dispersion is modified to be n2(hυ)=Z2(n02+2)42E2p2[GA(x)+GB(x)].Here, E2p=E2p_A, GA(x)=0.5x(0.5x)2+0.3(hγ/E2p)2, GB(x)=Λδ2(10.5δx)(10.5δx)2+0.3δ2(Λhγ/E2p)2, δ=E2p_A/E2p_B, hγ=hγA, and Λ=hγB/hγA.

    Figure 3 shows good agreement between Eq. (6) and the experimental data. GA(x) and GB(x) result in two peaks in the near-resonance region, indicating that the spin-split-off energy of excitons should play a significant role in the OKE of monolayer TMD materials. Furthermore, the B-exciton with a higher energy level from the spin-orbital splitting leads to a broadened and blueshift of the resonance of the individual n2 dispersion, in comparing the blue curves to the red ones in Fig. 3. This is attributed to higher energy levels of B-excitons, which exhibit larger linewidths of NLO resonance. The spin-splitting effect makes a great variation of overall n2 values as displayed by the purple curves.

    Normalized n2 dispersion of monolayer MoS2, MoSe2, WS2, and WSe2 calculated by the two excitons (purple), A exciton (red), and B exciton (blue) with hγ/E2p=0.075 eV. The symbols are the experimental data from Refs. [3032" target="_self" style="display: inline;">–32] and scaled with Z2′=1×10−14.

    Figure 3.Normalized n2 dispersion of monolayer MoS2, MoSe2, WS2, and WSe2 calculated by the two excitons (purple), A exciton (red), and B exciton (blue) with hγ/E2p=0.075  eV. The symbols are the experimental data from Refs. [3032" target="_self" style="display: inline;">–32] and scaled with Z2=1×1014.

    Figure 4 compares calculated n2 magnitudes from Eq. (4) or Eq. (6) with the TPB model. It is interesting to note that the maximal n2 values by Eq. (4) or Eq. (6) are in the range from 1×1012 to 1×1010  cm2/W, which are greater than the TPB model results by at least one order of magnitude. This indicates the 2D excitonic effects greatly enhance the OKE in 2D cases when compared with their bulk counterparts. As derived in Eqs. (4) and (6), the calculated n2 values are proportional to De·aB4, implying 2D materials possessing the higher density of excitons and/or the larger effective Bohr radius would exhibit greater OKE along with degenerate 2PA in many ways analogous to plasma screening effects [36]. As a result, RPP (In=4) is predicted to show the largest peak n2 value of 1.5×1010cm2/W among the 10 materials. The calculated n2 values are of the same order of magnitude as the multilayer graphene at an excitation wavelength in the near-infrared [37]. As compared with the experimental data of monolayer (or few-layer) TMDs in thin films, or dispersion [31,3841], the measured absolute n2 values are greater than from Eq. (6) [38,40,41]; also see Table 4 in Appendix C. One reason for the discrepancy is that the presence of a large number of defects/impurities in these films would give extra transitions in either 2PA or saturable absorption (SA) processes and, thus, enhance the measured nonlinearities. It is also noted that the absolute n2 values from our model are one order of magnitude larger than the reported results of MoS2 dispersion [31] and MoSe2 nanosheets [39], where light scattering in those amorphous samples may screen a refractive index change arising from the 2D nature of excitons.

    Calculated n2 dispersions of (a) monolayer RPP (In=1,2,3,4) and (b) TMDs by Eq. (4) and Eq. (6). Calculated n2 dispersions of (c) monolayer RPP (In=1,2,3,4) and (d) TMDs by the two-parabolic-band model [12]. Parameters used in the calculation are displayed in Table 3 in Appendix A.

    Figure 4.Calculated n2 dispersions of (a) monolayer RPP (In=1,2,3,4) and (b) TMDs by Eq. (4) and Eq. (6). Calculated n2 dispersions of (c) monolayer RPP (In=1,2,3,4) and (d) TMDs by the two-parabolic-band model [12]. Parameters used in the calculation are displayed in Table 3 in Appendix A.

    With the effects of 2D excitons, further enhancement of the OKE is predicted at lower temperature. As derived from Eq. (4) or (6), n2 values on resonance are inversely proportional to the square of the linewidth (2hγ) of 2D exciton, which would narrow down (as 35  meV) for RPPs, with temperature decreasing to 7  K [42]. The temperature dependence of the OKE could enlarge n2 values at peak (or valley) by one order of magnitude [see an example of n2 dispersion of RPP (In=1) in Fig. 5(a)]. The n2 value of RPP (In=1) is predicted to be 3×1010  cm2/W at 7 K, which would lead to a change in the refractive index of 3 (i.e., around 150% of the linear refractive index) in the case of I=10  GW/cm2. This would give rise to a narrow refractive index-near-zero region of intense laser light when it is on resonance with 2D exciton. In the off-resonance region, our calculated n2 values are nearly independent of x=hυ/E2p or the linewidth (2hγ). The scaled n2 values explicitly show an E2p2 dependence, which gives a variation by two orders of magnitude for 2D semiconducting materials with E2p ranging from 1 to 10 eV. This variation is clearly illustrated in Fig. 5(b).

    (a) Calculated n2 values as a function of photon energy (x axis) and temperature (y axis) for monolayer RPP (In=1). (b) A log-log plot of the scaled n2 in the off-resonance region versus E2p. The experimental n2 values are scaled by Z2′(n02+2)4G(x) with n0 and Z2′ listed in Table 3 in Appendix A. The solid line is the theoretical result of Eq. (4) with no adjustable parameters and a slope of −2.

    Figure 5.(a) Calculated n2 values as a function of photon energy (x axis) and temperature (y axis) for monolayer RPP (In=1). (b) A log-log plot of the scaled n2 in the off-resonance region versus E2p. The experimental n2 values are scaled by Z2(n02+2)4G(x) with n0 and Z2 listed in Table 3 in Appendix A. The solid line is the theoretical result of Eq. (4) with no adjustable parameters and a slope of 2.

    Indeed, in our model, the absolute magnitude of the nonlinear refractive index is enhanced through two-photon transitions, when the photon energy approaches hυ/(2E2p). Such a resonant enhancement of n2 is basically accompanied by absorption losses, commonly given by α=α0+βI, where α0 is the linear absorption coefficient. The OKE efficiency/transparency trade-off (n2/α) at a specific wavelength λ is therefore of importance to the assessment of a material for all-optical switching devices. The material requirements are as follows: WFOM=|n2|Iα0λ1 and TFOM=βλ|n2|1 [43,44]. Table 1 lists the FOMs for selected materials operated at room temperature, λ=1550  nm, and I=10  GW/cm2. Among them, RPP (In=4) is the best: its n2 is predicted to be as large as 1×1010  cm2/W with both WFOM=1.02 and TFOM=0.14. Compared with multilayer graphene, Si, and GaAs, RPP (In=4) performs even better, as its nonlinear refractive index is enhanced by two orders of magnitude, while TFOM is comparable [45,46]. Though RPPs (In=1,2,3) exhibit better TFOM (<0.1), indicating less 2PA losses, the WFOM values are less than 1 [29]. Note that the required light irradiance (I=10  GW/cm2) would be high. To reduce it, 2D materials in microcavities or photonic structures should be considered. For example, one may integrate 2D materials on clad nanophotonic cavities, such as ring and disk resonators, to increase the effective path length for light–matter interaction [47]. Arbitrary vertical heterostructures such as intentionally designed sequences of graphene, h-BN, and TMD monolayers (or 2D perovskites) [48,49] would give also rise to higher 2D-exciton density in order to reduce the operating light intensity.

    Nonlinear Coefficients and FOMs of Materials at 1550 nm

    Materialn2 [×1012  cm2/W]β [cm/GW]WFOMTFOMRef.
    2D RPP (In=4)105.596.31.020.14This work
    2D RPP (In=3)22.99.90.270.07This work
    2D RPP (In=2)7.21.60.120.03This work
    2D RPP (In=1)1.30.30.030.04This work
    2D MoS21.55.60.0080.56This work
    2D MoSe219.1212.50.071.72This work
    2D WS22.77.70.020.45This work
    2D WSe215.879.10.060.78This work
    Multilayer graphene−8009000.201.40[37]
    Si1100.0450.790.37[45]
    GaAs0.1610.20.10[45]
    GaAs/AlAs superlattice0.151.50.87[46]
    Conjugated 3,3’-bipyridine derivative0.0046<0.01>600<0.15[43]

    4. CONCLUSION

    In summary, based on the K-K transformation of exciton-associated 2PA, we have successfully developed an admirably succinct model to predict the OKE of 2D semiconductors. In our model, all the parameters are measurable experimentally, except for the linewidth of 2p exciton. On the examining the 10 2D materials mentioned here, we find that our model can provide a general guideline on both the n2 magnitude and n2 spectra. On our analysis, two-photon resonance with 2D excitons yields a remarkable contribution to the nonlinear refractive index spectra. As compared with the TPB model for 3D semiconductors, the OKE induced by 2D excitons shows a great enhancement. Based on this simple model, we assess RPP (In=4) crystals to be the best material candidate for all-optical switching at the 1550 nm wavelength. Our findings deepen the understanding of the OKE of 2D semiconducting materials and pave the way for highly efficient all-optical excitonic devices.

    APPENDIX A: DERIVATION OF n2 DISPERSION FOR 2D SEMICONDUCTORS WITH TWO EXCITONS

    In monolayer transition metal dichalcogenides (TMDs), there are two distinctive excitons: exciton A and exciton B. For these 2D semiconductors with the two-exciton feature, by a quantum perturbation theory associated with 2D excitons, we have derived the wavelength-dependent, degenerate 2PA coefficient as [28] β(hυ)=CNhυ(ElocE)4×[|μG1s|2(E1s_Ahυ)2+(Γ1s_A/2)2+|μG1s|2(E1s_Bhυ)2+(Γ1s_B/2)2]|Wnμ1snp|2Γnp/2π(Enp2hυ)2+(Γnp/2)2,where N is the density of active unit cells; ElocE=13(n02+2) is the local-field correction factor; n0 is the refractive index; hυ is the photon energy; E1s_A, E1s_B, and Enp are the energy level of the lowest 1s excitons (A exciton and B exciton) and higher np excitons, respectively. Γ1s_A, Γ1s_B, and Γnp refer to their linewidth. μij is the corresponding transition dipole moment from an i state to a j state. Wn is the weight value of an excitonic state; and C has a value of 3.47×1045 in units such that β is in cm/MW, N is in cm3, hυ, Ei, and Γi are in units of eV, and μij is in units of esu. Both the transition dipole moments and the weight values can be obtained in Refs. [28,29].

    Here, we assume that 1s- and np-excitonic states at room temperature have the equal linewidths, respectively (Γ1s_A=Γnp_A=2hγA, Γ1s_B=Γnp_B=2hγB). To fulfill the odd-parity requirement of two-photon transitions [50], only 2p excitons make the significant contribution to the 2PA (E2p2hυ), and other np-excitonic states are neglected, as shown in Table 2.

    Normalized Transition Dipole Moments for 2PA Transitionsa

    2PA Transitions|μG1s||μ1snp|[a.u.]
    G1s2p1
    G1s2s0
    G1s3p0.41
    G1s3s0
    G1s4p0.24
    G1s5p0.16
    G1s6p0.11

    Note: Transition dipole moments are normalized by |μG1s||μ1s2p| amplitude [28].

    The 2PA coefficient from Eq. (A1) becomes β(hυ)=CN(hγ)81π|W2|2|μG1s|2|μ1s2p|2(n02+2)4×[1(E1s_Ahυ)2+(hγA)2hυ(2hυE2p_A)2+(hγA)2+1(E1s_Bhυ)2+(hγB)2hυ(2hυE2p_B)2+(hγB)2],where the energy of 2p-excitonic state (E2p) is replaced with E2p_A and E2p_B. With approximations that |E1shυ|hγ and E1sE2phυ, Eq. (A2) is simplified to be β(hυ)=CN(hγ)81π|W2|2|μG1s|2|μ1s2p|2(n02+2)4×[1E2p_A2hυ(2hυE2p_A)2+(hγA)2+1E2p_B2hυ(2hυE2p_B)2+(hγB)2].

    By substituting Eq. (A3) into the K-K transformation, the n2 dispersion is derived to be n2(hυ)=CN(ch2γ)81π2|W2|2|μG1s|2|μ1s2p|2(n02+2)4×+[1/E2p_A2(2hυE2p_A)2+(hγA)2+1/E2p_B2(2hυE2p_B)2+(hγB)2]×hυ(hυ)2(hυ)2d(hυ).

    Because the equivalence of hυ and hυ would lead to the integral of β(hυ) with infrared divergences, a substitution of υ with υ+Δυ and a numerical summation were made for an approximate integral of β over the 2PA range (0<hυ<E2p). With further simplifications, we obtain a semi-empirical expression for n2 as n2(hυ)=Z2(n02+2)4E2p2[GA(x)+GB(x)],GA(x)=0.5x(0.5x)2+0.3(hγ/E2p)2,and GB(x)=Λδ2(10.5δx)(10.5δx)2+0.3δ2(Λhγ/E2p)2.

    Here, x=hυE2p. δ=E2p_A/E2p_B denotes an energy deviation of the spin-orbital splitting. Λ=hγB/hγA is the ratio of linewidths of the split excitons [51,52] with hγ=hγA being the best fitting parameter of A excitons. Z2=CN|μG1s|2·|μ1s2p|2(hγ)DeaB4(hγ) refers to a material-related parameter, where De and aB are the density and the Bohr radius of the 2D exciton, respectively. As for the 10 2D materials listed in Table 3, one may calculate the Z2 values and find that they are in the range from 0.98×1015 to 1.67×1015 with an average value of 1×1014, in the units such that n2 is in cm2/W; hυ, E2p, and hγ are in units of eV. E2p in Eq. (A5) interprets the energy of the A exciton at the 2p state, which is given by E2pEgEb/n020.1. Eb is the exciton binding energy. All parameters used in the calculation of n2 dispersions are listed in Table 3.

    Parameters Used in the Calculation of n2 Dispersionsa

    n0aB [Å]DeEg [eV]Eb [eV]E1s [eV]E2p [eV]Z2 [×1015]hγ [eV]δΛ
    RPP (In=1)2.1115.30.60×10202.740.352.392.590.9790.15N.A.N.A.
    RPP (In=2)2.2117.11.20×10202.510.262.252.373.050.15N.A.N.A.
    RPP (In=3)2.2717.91.80×10202.270.162.112.145.500.15N.A.N.A.
    RPP (In=4)2.3222.02.40×10202.140.151.992.0016.70.15N.A.N.A.
    MoS21.849.35.00×10202.700.801.902.501.670.0750.9392
    MoSe22.1010.11.81×10212.430.661.772.218.400.0750.9092
    WS21.8210.36.28×10202.920.822.102.733.170.0750.8674
    WSe21.8410.52.28×10212.570.791.782.3512.40.0750.8484
    h-BN2.006.12.40×10215.810.925.505.691.000.15N.A.N.A.
    BP2.5020.23.80×10201.600.501.101.2020.00.15N.A.N.A.

    Note that no fitting parameter was used in plotting the theoretical curves, except for the linewidth, and the experimental data of n2 values [14,15,28,29,5355] are scaled with the average value of Z2=1×1014.

    APPENDIX B: SIMPLIFICATION OF n2 DISPERSION OF 2D SEMICONDUCTORS WITH ONE EXCITON

    For 2D materials without the spin-orbital splitting, the 2PA coefficient from Eq. (A3) becomes β(hυ)=CN(ch2γ)81π2|W2|2|μG1s|2|μ1s2p|2×(n02+2)4E2p2hυ(2hυE2p)2+(hγ)2.

    Here, we define a parameter C2=CN(hγ)81π|W2|2|μG1s|2·|μ1s2p|2 for the simplification of the 2PA coefficient. By substituting Eq. (B1) into the K-K transformation, the n2 dispersion of 2D semiconductors with one exciton can be obtained. Furthermore, we have δ=E2p_A/E2p_B=1 when there is no spin-orbital splitting of 2D materials. Equation (A5) can be simplified to be n2(hυ)=Z2(n02+2)4E2p2G(x),where GA(x)=GB(x)=G(x) becomes G(x)=0.5x(0.5x)2+0.3(hγ/E2p)2.

    Equation (B2) is used to calculate the n2 dispersions of 2D RPPs, h-BN, and BP.

    APPENDIX C: EXPERIMENTAL DATA OF n2 VALUES

    Extracted n2 Values from Experimental Data

     n2[×1012cm2/W]n2E2p2Z2(n02+2)4λ [nm]n0Ref.
    RPP (In=1)0.410.1627002.11[25]
    RPP (In=2)0.450.1127002.21[25]
    RPP (In=3)0.390.04827002.41[25]
    RPP (In=4)0.400.05327002.32[25]
    MoS2−1.96−1.458001.84[30]
    350260.028001.84[38]
    1.881.4010641.84[38]
    −0.21−0.1510641.84[31]
    MoSe2−0.12−0.03410642.10[31]
    0.200.05810002.10[39]
    WS2−1.10−1.028001.82.[30]
    0.810.768001.82[32]
    58.3054.5510641.82[40]
    128119.7710401.82[41]
    WSe2−18.70−12.2710401.84[41]
    h-BN0.120.3010642.0[33]
    BP86026.738002.5[34]

    APPENDIX D: TEMPERATURE DEPENDENCE OF NONLINEAR REFRACTIVE INDEX

    According to Eq. (4) or (6), n2 values are predicted to have the linewidth dependence in the near-resonance region in our model. Previous work [42] has demonstrated experimentally that the temperature-dependent linewidth of exciton absorption in perovskites can be expressed empirically by γ(T)=Γ0+σT+ΓLO(eωLO/kBT1)12h,where Γ0 is a temperature-independent inhomogeneous broadening term, which is determined by the material size, shape, and composition; σ represents the coupling strength of acoustic phonon scattering; and ΓLO represents the coupling strength of optical (LO) phonon scattering, which has a Bose–Einstein distribution for occupation numbers of the respective LO phonons. Here, for computation ease, the linewidths of the RPPs (In=1,2,3,4) or TMDs are assumed to be the same, as the variations among these 2D materials are insignificant. As a result, temperature dependence of n2 dispersion of RPPs (In=1,2,3,4) and TMDs can be calculated and displayed in Figs. 5(a), 6, and 7.

    Nonlinear refractive index, n2, as a function of photon energy (x axis) and temperature (y axis) for monolayer RPP: (a) In=2, (b) In=3, and (c) In=4. These n2 values are calculated with the averaged Z2′=1×10−14.

    Figure 6.Nonlinear refractive index, n2, as a function of photon energy (x axis) and temperature (y axis) for monolayer RPP: (a) In=2, (b) In=3, and (c) In=4. These n2 values are calculated with the averaged Z2=1×1014.

    Nonlinear refractive index, n2, as a function of photon energy (x axis) and temperature (y axis) for TMD monolayers: (a) MoS2, (b) MoSe2, (c) WS2, and (d) WSe2. These n2 values are calculated with the averaged Z2′=1×10−14.

    Figure 7.Nonlinear refractive index, n2, as a function of photon energy (x axis) and temperature (y axis) for TMD monolayers: (a) MoS2, (b) MoSe2, (c) WS2, and (d) WSe2. These n2 values are calculated with the averaged Z2=1×1014.

    APPENDIX E: FIGURES OF MERIT OF MONOLAYER RPPS AND TMDS

    To evaluate the OKE efficiency for all-optical switching devices, we have calculated the figures of merit (FOMs) for monolayer RPPs and TMDs in the wavelength range from 600 to 1700 nm. Results are displayed in Fig. 8. The WFOM values were calculated with I0=10  W/cm2, the maximum peak intensity before the onset of irreversible change or damage in the material [28,29], which represents the maximum nonlinear phase shift that can be achieved with the material. In the transparent spectral region, α0 is estimated to be 2×106  cm1 for RPPs [56]; and 1.4×105  cm1 for monolayer TMDs [57], respectively.

    (a) WFOM of monolayer TMDs, (b) WFOM of monolayer RPPs, (c) TFOM of monolayer TMDs, and (d) TFOM of monolayer RPPs. The gray area corresponds to the wavelength range for optical communications.

    Figure 8.(a) WFOM of monolayer TMDs, (b) WFOM of monolayer RPPs, (c) TFOM of monolayer TMDs, and (d) TFOM of monolayer RPPs. The gray area corresponds to the wavelength range for optical communications.

    References

    [1] R. W. Boyd. Nonlinear Optics(2020).

    [2] A. Newell. Nonlinear Optics(2018).

    [3] X. Hu, P. Jiang, C. Ding, H. Yang, Q. Gong. Picosecond and low-power all-optical switching based on an organic photonic-bandgap microcavity. Nat. Photonics, 2, 185-189(2008).

    [4] L. Deng, E. W. Hagley, J. Wen, M. Trippenbach, Y. Band, P. S. Julienne, J. E. Simsarian, K. Helmerson, S. L. Rolston, W. D. Phillips. Four-wave mixing with matter waves. Nature, 398, 218-220(1999).

    [5] K. Inoue, T. Mukai, T. Saitoh. Nearly degenerate four-wave mixing in a traveling-wave semiconductor laser amplifier. Appl. Phys. Lett., 51, 1051-1053(1987).

    [6] K. J. A. Ooi, D. K. T. Ng, T. Wang, A. K. L. Chee, S. K. Ng, Q. Wang, L. K. Ang, A. M. Agarwal, L. C. Kimerling, D. T. H. Tan. Pushing the limits of CMOS optical parametric amplifiers with USRN: Si7N3 above the two-photon absorption edge. Nat. Commun., 8, 13878(2017).

    [7] T. Brabec, C. Spielmann, P. F. Curley, F. Krausz. Kerr lens mode locking. Opt. Lett., 17, 1292-1294(1992).

    [8] X. Liu, D. Popa, N. Akhmediev. Revealing the transition dynamics from Q switching to mode locking in a soliton laser. Phys. Rev. Lett., 123, 093901(2019).

    [9] M. J. Weber, D. Milam, W. L. Smith. Nonlinear refractive index of glasses and crystals. Opt. Eng., 17, 175463(1978).

    [10] R. Adair, L. L. Chase, S. A. Payne. Nonlinear refractive index of optical crystals. Phys. Rev. B, 39, 3337-3350(1989).

    [11] X. J. Zhang, W. Ji, S. H. Tang. Determination of optical nonlinearities and carrier lifetime in ZnO. J. Opt. Soc. Am. B, 14, 1951-1955(1997).

    [12] M. Sheik-Bahae, D. J. Hagan, E. W. Van Stryland. Dispersion and band-gap scaling of the electronic Kerr effect in solids associated with two-photon absorption. Phys. Rev. Lett., 65, 96-99(1990).

    [13] T. Olsen, S. Latini, F. Rasmussen, K. S. Thygesen. Simple screened hydrogen model of excitons in two-dimensional materials. Phys. Rev. Lett., 116, 056401(2016).

    [14] G. Zhang, A. Chaves, S. Huang, F. Wang, Q. Xing, T. Low, H. Yan. Determination of layer-dependent exciton binding energies in few-layer black phosphorus. Sci. Adv., 4, 9977(2018).

    [15] B. Arnaud, S. Lebègue, P. Rabiller, M. Alouani. Huge excitonic effects in layered hexagonal boron nitride. Phys. Rev. Lett., 96, 026402(2006).

    [16] Z. Jiang, Z. Liu, Y. Li, W. Duan. Scaling universality between band gap and exciton binding energy of two-dimensional semiconductors. Phys. Rev. Lett., 118, 266401(2017).

    [17] K. F. Mak, D. Xiao, J. Shan. Light–valley interactions in 2D semiconductors. Nat. Photonics, 12, 451-460(2018).

    [18] C. Trovatello, F. Katsch, N. J. Borys, M. Selig, K. Yao, R. Borrego-Varillas, F. Scotognella, I. Kriegel, A. Yan, A. Zettl, P. J. Schuck, A. Knorr, G. Cerullo, S. Dal Conte. The ultrafast onset of exciton formation in 2D semiconductors. Nat. Commun., 11, 5277(2020).

    [19] A. Autere, H. Jussila, Y. Dai, Y. Wang, H. Lipsanen, Z. Sun. Nonlinear optics: nonlinear optics with 2D layered materials. Adv. Mater., 30, 1870172(2018).

    [20] I. Abdelwahab, P. Dichtl, G. Grinblat, K. Leng, X. Chi, I.-H. Park, M. P. Nielsen, R. F. Oulton, K. P. Loh, S. A. Maier. Giant and tunable optical nonlinearity in single-crystalline 2D Perovskites due to excitonic and plasma effects. Adv. Mater., 31, 1902685(2019).

    [21] N. Dong, Y. Li, S. Zhang, N. McEvoy, R. Gatensby, G. S. Duesberg, J. Wang. Saturation of two-photon absorption in layered transition metal dichalcogenides: experiment and theory. ACS Photon., 5, 1558-1565(2018).

    [22] Y. Yu, Y. Yu, C. Xu, A. Barrette, K. Gundogdu, L. Cao. Fundamental limits of exciton-exciton annihilation for light emission in transition metal dichalcogenide monolayers. Phys. Rev. B, 93, 201111(2016).

    [23] A. Tanaka, N. J. Watkins, Y. Gao. Hot-electron relaxation in the layered semiconductor 2H-MoS2 studied by time-resolved two-photon photoemission spectroscopy. Phys. Rev. B, 67, 113315(2003).

    [24] H. H. Fang, J. Yang, S. Adjokatse, E. Tekelenburg, M. E. Kamminga, H. Duim, J. Ye, G. R. Blake, J. Even, M. A. Loi. Band-edge exciton fine structure and exciton recombination dynamics in single crystals of layered hybrid perovskites. Adv. Funct. Mater., 30, 1907979(2020).

    [25] F. O. Saouma, C. C. Stoumpos, J. Wong, M. G. Kanatzidis, J. I. Jang. Selective enhancement of optical nonlinearity in two-dimensional organic-inorganic lead iodide perovskites. Nat. Commun., 8, 742(2017).

    [26] B. Guo, Q. L. Xiao, S. H. Wang, H. Zhang. 2D layered materials: synthesis, nonlinear optical properties, and device applications. Laser Photon. Rev., 13, 1800327(2019).

    [27] G. Grinblat, I. Abdelwahab, M. P. Nielsen, P. Dichtl, K. Leng, R. F. Oulton, K. P. Loh, S. A. Maier. Ultrafast all-optical modulation in 2D hybrid perovskites. ACS Nano, 13, 9504-9510(2019).

    [28] F. Zhou, J. H. Kua, S. Lu, W. Ji. Two-photon absorption arises from two-dimensional excitons. Opt. Express, 26, 16093-16101(2018).

    [29] F. Zhou, I. Abdelwahab, K. Leng, K. P. Loh, W. Ji. 2D perovskites with giant excitonic optical nonlinearities for high-performance sub-bandgap photodetection. Adv. Mater., 31, 1904155(2019).

    [30] T. Neupane, B. Tabibi, F. J. Seo. Spatial self-phase modulation in WS2 and MoS2 atomic layers. Opt. Mater. Express, 10, 831-842(2020).

    [31] K. Wang, Y. Feng, C. Chang, J. Zhan, C. Wang, Q. Zhao, J. N. Coleman, L. Zhang, W. J. Blau, J. Wang. Broadband ultrafast nonlinear absorption and nonlinear refraction of layered molybdenum dichalcogenide semiconductors. Nanoscale, 6, 10530-10535(2014).

    [32] X. Zheng, Y. Zhang, R. Chen, Z. Xu, T. Jiang. Z-scan measurement of the nonlinear refractive index of monolayer WS2. Opt. Express, 23, 15616-15623(2015).

    [33] P. Kumbhakar, A. K. Kole, C. S. Tiwary, S. Biswas, S. Vinod, J. Taha-Tijerina, U. Chatterjee, P. M. Ajayan. Nonlinear optical properties and temperature-dependent UV-vis absorption and photoluminescence emission in 2D hexagonal boron nitride nanosheets. Adv. Opt. Mater., 3, 828-835(2015).

    [34] X. Zheng, R. Chen, G. Shi, J. Zhang, Z. Xu, T. Jiang. Characterization of nonlinear properties of black phosphorus nanoplatelets with femtosecond pulsed Z-scan measurements. Opt. Lett., 40, 3480-3483(2015).

    [35] T. C. Berkelbach, M. S. Hybertsen, D. R. Reichman. Theory of neutral and charged excitons in monolayer transition metal dichalcogenides. Phys. Rev. B, 88, 045318(2013).

    [36] J. Wang, A. Coillet, O. Demichel, Z. Wang, D. Rego, A. Bouhelier, P. Grelu, B. Cluzel. Saturable plasmonic metasurfaces for laser mode locking. Light Sci. Appl., 9, 50(2020).

    [37] G. Demetriou, H. T. Bookey, F. Biancalana, E. Abraham, Y. Wang, W. Ji, A. K. Kar. Nonlinear optical properties of multilayer graphene in the infrared. Opt. Express, 24, 13033-13043(2016).

    [38] F. Liu, X. Zhao, X. Q. Yan, X. Xin, Z. B. Liu, J. G. Tian. Measuring third-order susceptibility tensor elements of monolayer MoS2 using the optical Kerr effect method. Appl. Phys. Lett., 113, 051901(2018).

    [39] H. Pan, H. Chu, Y. Li, S. Zhao, D. Li. Comprehensive study on the nonlinear optical properties of few-layered MoSe2 nanosheets at 1 μm. J. Alloys Compd., 806, 52-57(2019).

    [40] S. Bikorimana, P. Lama, A. Walser, R. Dorsinville, S. Anghel, A. Mitioglu, A. Micu, L. Kulyuk. Nonlinear optical responses in two-dimensional transition metal dichalcogenide multilayer: WS2, WSe2, MoS2 and Mo0.5W0.5S2. Opt. Express, 24, 20685-20695(2016).

    [41] N. Dong, Y. Li, S. Zhang, N. L. McEvoy, X. Zhang, Y. Cui, L. Zhang, G. S. Duesberg, J. Wang. Dispersion of nonlinear refractive index in layered WS2 and WSe2 semiconductor films induced by two-photon absorption. Opt. Lett., 41, 3936-3939(2016).

    [42] H. C. Woo, J. W. Choi, J. Shin, S. H. Chin, M. H. Ann, C. L. Lee. Temperature-dependent photoluminescence of CH3NH3PbBr3 perovskite quantum dots and bulk counterparts. J. Phys. Chem. Lett., 9, 4066-4074(2018).

    [43] Q. Chen, E. H. Sargent, N. Leclerc, A. J. Attias. Wavelength dependence and figures of merit of ultrafast third-order optical nonlinearity of a conjugated 3, 3′-bipyridine derivative. Appl. Opt., 42, 7235-7241(2003).

    [44] F. Chérioux, A. J. Attias, H. Maillotte. Symmetric and asymmetric conjugated 3, 3′-bipyridine derivatives as a new class of third-order NLO chromophores with an enhanced non-resonant, nonlinear refractive index in the picosecond range. Adv. Funct. Mater., 12, 203-208(2002).

    [45] M. Dinu, F. Quochi, H. Garcia. Third-order nonlinearities in silicon at telecom wavelength. Appl. Phys. Lett., 82, 2954-2956(2003).

    [46] S. J. Wagner, J. Meier, A. S. Helmy, J. S. Aitchison, M. Sorel, D. C. Hutchings. Polarization-dependent nonlinear refraction and two-photon absorption in GaAs/AlAs superlattice waveguides below the half-bandgap. J. Opt. Soc. Am. B, 24, 1557-1563(2007).

    [47] T. K. Fryett, A. Zhan, A. Majumdar. Phase-matched nonlinear optics via patterning layered materials. Opt. Lett., 42, 3586-3589(2017).

    [48] D. Pan, Y. Fu, N. Spitha, Y. Zhao, C. R. Roy, D. J. Morrow, D. D. Kohler, J. C. Wright, S. Jin. Deterministic fabrication of arbitrary vertical heterostructures of two-dimensional Ruddlesden-Popper halide perovskites. Nat. Nanotechnol., 16, 159-165(2021).

    [49] F. Withers, O. Del Pozo-Zamudio, A. Mishchenko, A. P. Rooney, A. Gholinia, K. Watanabe, T. Taniguchi, S. J. Haigh, A. K. Geim, A. I. Tartakovskii, K. S. Novoselov. Light-emitting diodes by band-structure engineering in van der Waals heterostructures. Nat. Mater., 14, 301-306(2015).

    [50] S. Uryu, H. Ajiki, T. Ando. Excitonic two-photon absorption in semiconducting carbon nanotubes within an effective-mass approximation. Phys. Rev. B, 78, 115414(2008).

    [51] O. B. Aslan, M. Deng, T. F. Heinz. Strain tuning of excitons in monolayer WSe2. Phys. Rev. B, 98, 115308(2018).

    [52] E. Courtade, B. Han, S. Nakhaie, C. Robert, X. Marie, P. Renucci, T. Taniguchi, K. Watanabe, L. Geelhaar, J. M. J. Lopes, B. Urbaszek. Spectrally narrow exciton luminescence from monolayer MoS2 and MoSe2 exfoliated onto epitaxially grown hexagonal BN. Appl. Phys. Lett., 113, 032106(2018).

    [53] A. Segura, L. Artús, R. Cuscó, T. Taniguchi, G. Cassabois, B. Gil. Natural optical anisotropy of h-BN: highest giant birefringence in a bulk crystal through the mid-infrared to ultraviolet range. Phys. Rev. Mater., 2, 024001(2018).

    [54] T. C. Doan, J. Li, J. Y. Lin, H. X. Jiang. Bandgap and exciton binding energies of hexagonal boron nitride probed by photocurrent excitation spectroscopy. Appl. Phys. Lett., 109, 122101(2016).

    [55] X. Wang, S. Lan. Optical properties of black phosphorus. Adv. Opt. Photon., 8, 618-655(2016).

    [56] I. Abdelwahab, G. Grinblat, K. Leng, Y. Li, X. Chi, A. Rusydi, S. A. Maier, K. Ploh. Highly enhanced third-harmonic generation in 2D perovskites at excitonic resonances. ACS Nano, 12, 644-650(2018).

    [57] K. F. Mak, C. Lee, J. Hone, J. Shan, T. F. Heinz. Atomically thin MoS2: a new direct-gap semiconductor. Phys. Rev. Lett., 105, 136805(2010).

    Feng Zhou, Cacere Jelah Nieva, Dianyuan Fan, Shunbin Lu, Wei Ji. Superior optical Kerr effects induced by two-dimensional excitons[J]. Photonics Research, 2022, 10(3): 834
    Download Citation