Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Third-order optical nonlinearity of CsPb(Br/I)3 metal halide perovskites nano-crystals embedded chalcogenide glass

Open Access Open Access

Abstract

The nonlinear optical properties of emerging metal halide perovskites (MHP) materials are sufficiently intriguing that this topic has become the hotspots in the realm of material science. Hence, we investigate the third-order nonlinear optical properties of CsPbBrx/I3-x (x = 1, 2, 3) MHP nano-crystals (NCs) embedded chalcogenide glass (ChG) within a GeS2-Sb2S3 pseudo-binary system, by monitoring the composition, excitation wavelength and intensity dependencies via femtosecond Z-scan technique. We have found that the intrinsic large optical nonlinearity of ChG can be further enhanced because of the incorporation of MHP NCs, and that the optical nonlinearity of MHP-ChG containing pristine Br NCs is more pronounced compared to its counterparts with mixed Br/I NCs, due to a combination of multiple factors.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In addition to their outstanding linear optical properties [13], emerging metal halide perovskites (MHP) materials also exhibit striking nonlinear optical properties [47], resulting from their unique photoelectric features, such as strong spin-orbit coupling, high defect-tolerance, and direct p-p optical transitions, etc., making the advances in this research fundamental to the development of photonic devices of the next generation. Normally, both organic and inorganic MHP materials exhibit predominantly third-order optical nonlinearity (TONL) in their structural centrosymmetry, thus suitable and desirable for those advanced TONL-based devices [810] such as all-optical switches, terahertz generation, bio-imaging, up-conversion laser, etc. However, basic research with respect to the TONL mechanisms of MHP materials remains quite limited, so that this research area is still in its infancy [1113].

On the other hand, chalcogenide glasses (ChGs), are also excellent optical materials with large TONL, high infrared transmittance and easy shaping and molding etc., rendering them a high-performance platform for both nonlinear and infrared photonics [14,15]. More importantly, ChGs are friendly hosts for a variety of nano-particles which in turn enhance the various properties of their host ChGs [16], thus one can expect superior TONL performance via a hybrid of MHPs and ChGs. Furthermore, although not so robust as oxide glasses, ChG packages can still protect the MHPs from moisture and oxygen in ambient atmosphere, and a unique advantage over oxide glasses is that the vacuum melting process of ChGs can prohibit volatilization of the original metal halide during preparation. However, as far as we know, studies with respect to the TONL properties of MHPs in ChGs or even in oxide glasses are scarce. The main contribution to this specific research topic came from Xiang’s group [1720], who showed that the incorporation of inorganic MHP (CsPbX3 (X = Cl, Br, I) or mutants) nano-crystals in oxide glasses could overcome the intrinsic defects of MHP (especially the unstable nature) and improve PL and TONL performances. Recently, Liu et al [21] also characterized the nonlinear properties of CsPbBr3 NCs in borosilicate glasses and discovered strong two-photon process induced up-conversion PL from the MHP NCs.

Our latest series of work [2224] established an in-situ nanocrystallization strategy for the precipitation of inorganic MHP (CsPbX3) NCs within GeS2-Sb2S3 ChG system, demonstrating the exceptional compatibility between these MHP NCs and ChGs. In this paper, an in-depth and thorough investigation with respect to the TONL attributes of the MHP-ChG composites (including composition, excitation wavelength and intensity dependencies) is present, in order to fill the gaps in this research realm.

2. Experiment

The molar compositions of the MHP-ChGs are 79.2GeS2−15.8Sb2S3−5CsPbBrx/I3-x (x = 1, 2, 3, abbreviated as Br1, Br2, Br3 respectively) and 83.4GeS2−16.6Sb2S3 for host ChG. The detailed glass preparation and nanocrystallization processes have been described in our previous paper [24]. The batch were all polished to thickness of 1 mm and mirror-like on both sides for the follow-up optical tests.

Transmission spectra from 0.4 to 2.5 µm were recorded via Perkin-Elmer Lambda 950 spectrometer, those from 2.5 to 25 µm via Nicolet381 Fourier transform IR spectrometer. The optical nonlinear attributes in wavelength range from 700 to 920 nm for the MHP-ChGs were investigated via Z-scan technique. The laser beam was generated from Coherent Chameleon Ultra II femtosecond laser (140 ± 10 fs laser pulse and 80 MHz repetition rate), and the laser power was detected by Si photodiodes. In close-aperture (CA) Z-scan measurements, for the purpose of extract the signal from the nonlinear refraction (n2), the linear transmittance was kept at 10% through the aperture. In open-aperture (OA) Z-scan measurements, removing the aperture to access the full transmitted light and the nonlinear absorption coefficient (β). The detailed fitting formulas for the CA and OA data were given in Eq. (S1) and Eq. (S2) of Supplement 1.

The incident laser focused on the sample surface via convex lens which has focal distance of 7.5 mm and beam waist estimated at 16 ± 0.5 µm. The laser power (P) was varied from 3.65 to 43.75 mW, corresponding to a laser density at the focal point (I0) from 0.028 to 1.29 GW/cm2, aiming to explore the irradiation dependence of n2 and β as well as the corresponding mechanism. It should be noted that the laser power is so small that the cumulative thermo-optic effect is absent in the Z-scan measurements. The mentioned optical tests were performed under room temperature.

3. Results and discussion

The broad-band transmission spectra as well as a photograph of the three MHP-ChG composites with host ChG are given in Fig. 1, showing that the composites are dark red in color and a little bit visible transparent. According to their scanning electron microscope images (Fig. S1 in supplemental document), the mean diameter of the spherical MHP NCs embedded in Br1, Br2, Br3 are estimated to be 38.6 ± 4 nm, 43.4 ± 4 nm, 59.3 ± 6 nm respectively, thus the NCs in such sizes do not affect the IR transparency near the 12 µm cut-off. The overall optical transmittance of the composites reached >80% in flat region, indicating negligible optical scattering at the interfaces between the glass matrix and crystals. Given that the tested linear refractive index n0 for the GeS2-Sb2S3 ChG of 2.302 at 632.8 nm and the estimated n0 values of 2.2 ± 0.1 for CsPbBr3 and 2.4 ± 0.1 for CsPbI3 via an empirical formula [25]: n02 = 1 + 8.32/Eg, where Eg is the bandgap energy, 2.4 ± 0.1 eV and 1.8 ± 0.1 eV for CsPbBr3 and CsPbI3 respectively, the matched n0 between the host ChG and the CsPb(Br/I)3 NCs minimized the optical scattering, which resulted in high optical transparency of the MHP-ChG composites.

 figure: Fig. 1.

Fig. 1. Broad-band transmission spectra of the three MHP-ChGs and the host ChG. The inset is a photograph of the four samples.

Download Full Size | PDF

Figure 2 is the absorption spectra of the MHP-ChG samples, showing that the bandgap edges are red-shifted as bromine is substituted by iodine, with the theoretical positions (defined as the linear absorption coefficient α at 10 cm−1 [26,27]) estimated at 626, 573 and 541 nm for Br1, Br2 and Br3, corresponding to bandgap energies Eg of 1.985, 2.168 and 2.296 eV respectively. It should be noted that the Eg is a complex value mainly contributed from the ChG host and influenced by the MHP NCs, thus Eg of the sample with pristine CsPbBr3 NCs (Br3) is smaller than the those for the CsPbBr3 NCs dispersed colloidal solutions as well as in oxide glasses (given in Table 1), due to the inherent large number of lone-pair electrons from the GeS2-Sb2S3 host ChG that have occupied the conduction band minimum [26]. Besides, below the bandgap edge, only Br3 exhibits obvious exciton absorption from the MHP NCs, as a result of its diversity in crystal sizes and larger crystal number as can be judged from the strongest diffraction intensity in the X-ray diffraction patterns Fig. S2 in the Supplement 1.

 figure: Fig. 2.

Fig. 2. Absorption spectra of the three MHP-ChGs. The definition of Eg and excitation wavelengths for Z-scan measurements are noted.

Download Full Size | PDF

Tables Icon

Table 1. Host material, style, bandgap energy and nonlinear optical parameters of the CsPbBr3 MHP materials in this work and others’ reports

The broad spectral TONL properties of the MHP-ChGs were explored via Z-scan method in the wavelength range from 700 to 920 nm. The excitation wavelength region is noted in the absorption spectra, showing that all the three samples have sub-bandgap absorption from the Urbach tail (defect states) at the excitation-onset wavelengths, whereas Br3 has additional linear absorption throughout the entire excitation region due to exciton interactions, such that it may exhibit a stronger nonlinear response from both the photon-generated free carrier effects and exciton effects, as detailed below.

Figure 3(a)-(c) shows 21 representative CA Z-scan traces of the three MHP-ChGs tested at seven wavelengths under a laser power of 10 ± 1 mW. Note that all the CA profiles are valley-following-peak shaped, suggesting the self-focusing behavior namely, the positive n2 sign is a commonality for the MHP-ChGs from 700 to 920 nm. At similar excitation wavelengths, this result is in contrast to the n2 sign of CsPbBr3 colloidal quantum dots in n-hexane solvent [28] and CsPbBr3 nano-cubes/rods on glass substrates [29], but is consistent with the n2 sign of CsPb(Br/I)3 cubic NCs in anhydrous toluene [30] and hexane solvent [31], manifesting the highly variable TONL attributes of MHPs, related to composition, size, shape and substrate related. Besides, as can be judged from the naked eyes, the size of the peaks and valleys in the CA profiles, more specifically, the transmittance difference between peaks and valleys (ΔTp-v), shrinks at longer excitation wavelength. In general, a smaller ΔTp-v represents a smaller nonlinear phase shift, resulting in a smaller n2 value for the same laser intensity. Hence, a normal n2 dispersion for the MHP-ChGs can be concluded.

 figure: Fig. 3.

Fig. 3. CA and OA Z-scans under laser power of 10 ± 1 mW at wavelengths from 700 to 920 nm for Br1 (a) and (d), Br2 (b) and (e), Br3 (c) and (f), respectively. The data points are obtained from the experiments, and the solid lines represent the theoretical fits.

Download Full Size | PDF

The specific n2 values were calculated by using Kwak’s fitting procedure [32] proposed for large nonlinear phase shift (ΔTp-v > 1) in the CA traces, and all n2 values are given in Table S1 of the supplemental document. By plotting the n2 values against laser power (P) for the seven tested wavelengths as illustrated in Fig. S3-S9 in the supplemental document, it can be seen that the n2-P relations exhibit a general overall negative dependence, although in some cases rather weak and insignificant. This observation is similar to the TONL attribute of the organic-inorganic hybrid MHPs [33], demonstrating that the nonlinear refraction of the present MHP-ChGs also arises in part from a photon-induced carrier effect which is suppressed at high laser intensities, the so-called Pauli blocking effect. The calculations do confirm that Br3 exhibits larger n2 values compared to the other two samples. Therefore, as a representative, its minimum and maximum n2 values along with those of the other CsPbBr3 counterparts in various styles and hosts in previous reports are summarized in Table 1. The comparison among these n2 values shows that the n2 behavior of the present CsPbBr3 NCs embedded ChG is remarkably stronger than those of the CsPbBr3 NCs embedded borosilicate glass [19] and hexane [31], comparable to quantum dots dispersed in n-hexane [28], but weaker than those of mixed-phase NCs dispersed in anhydrous toluene [30] and shaped NC films coated on silica glass [29]. Furthermore, it is noted that the CsPbBr3 nanoparticles dispersed hexane [35] has n2 values at 1064 nm larger than ours, which can be assigned to the strong accumulated free-carrier effect due to the use of nanosecond pulse laser.

Figure 3(d)-(f) shows the 21 OA Z-scan traces corresponding to their CA counterparts. It is evident that a single valley is present in the central of each profile, signifying reverse saturation absorption (RSA, positive β) as a commonality for MHP-ChGs in this wavelength region. A continuous reduction of the valley size (transmittance depth ΔTv) at longer wavelength can be observed, which manifests β attenuation, in analogy to the positive dispersion of the n2 value. In details, the photon energy Ep of excitation wavelength is from 1.35 eV (920 nm) to 1.78 eV (700 nm), giving the range of normalized photon energy (Ep/Eg) from 0.58 to 0.9 meet the criterion for two-photon absorption (2PA, 0.5 < Ep/Eg < 1), thus the RSA can be attributed to 2PA. It should be noted that 2PA dominates under the entire excitation densities from 0.028 to 1.29 GW/cm2, without the presence of sub-bandgap absorption saturation (SAS) that observed in many other MHP materials [21,28,33,40,41] at both low power and high power ends. The absent SAS for the MHP-ChGs, especially Br3 with considerable sub-bandgap absorption is simply because it is submerged by the strong 2PA of the host ChG. An analogous observation had been reported in the CsPb(Cl/Br)3 NCs embedded oxide glasses [20] which exhibit persistent 2PA under incident laser densities from 25.5 to 332 GW/cm2 at a wavelength of 800 nm.

The concrete 2PA coefficient β2PA was calculated via a fit to the OA Z-scan trajectory by using the proven 2PA model [42], and all the values are listed in Table S1. The β2PA-P relations at each excitation wavelength are illustrated in Fig. S3-S9, again they exhibit a general overall negative correlation, confirming the presence of Pauli blocking effect. Table 1 lists the minimum and maximum β2PA value for Br3 as well as those of CsPbBr3 materials in different styles and hosts from others’ reports. Notably, the β2PA magnitude for the present CsPbBr3 embedded ChG is outstanding, it is remarkably larger than those of the majority of its counterparts, similar to those of CsPbBr3 nanosheets [38], but smaller than that of CsPbBr3 quantum dots film coated on silicon substrates [34], which might exhibit inherently large 2PA at the testing wavelength of 720 nm.

The significant larger n2 and β2PA values of the CsPbBr3 NCs embedded ChG compared to the identical NCs embedded oxide glasses and organic solvents demonstrated the strong impact of host (or substrate) upon the TONL attribute of MHP materials. In this regard, a direct comparison of the OA and CA Z-scans between the host ChG and three MHP-ChGs under similar laser power of 3.75 ± 0.1 mW at 700 nm are illustrated in Fig. 4, which clearly shows that the host ChG exhibits distinct n2 and β2PA signals represented in the CA (ΔTp-v) and OA (ΔTv) profile respectively, but is weaker than the MHP-ChGs, indicating the advance of the intrinsic large ChG TONL after the inorganic MHP NCs incorporation. Besides, in previous study [43] with respect to the TONL properties of ChGs within GeS2-Sb2S3 system, the researchers found that the n2 value of the ChGs depends on Sb2S3 content and the maximum n2 obtained from a 60GeS2−40Sb2S3 (in mol%) glass is 6.65 × 10−14 cm2/W at excitation wavelength of 800 nm. In this study, molar composition of the host ChG is 83.4GeS2−16.6Sb2S3, but the average n2 value at 800 nm for the three MHP-ChGs are 10.56 × 10−14 cm2/W (Br1), 7.77 × 10−14 cm2/W (Br2) and 12.86 × 10−14 cm2/W (Br3) respectively, all of them are remarkably larger than that of the 60GeS2−40Sb2S3 glass. A major reason for such overall augment is the enhanced light-mater interactions via irradiation of femtosecond laser induced dielectric confinement effect due to the large dielectric difference between the amorphous host ChG composed of covalent bonds and these crystalline MHP NCs composed of ionic bonds.

 figure: Fig. 4.

Fig. 4. CA (a) and OA (b) Z-scans of the three MHP-ChGs and the host ChG under laser power of 3.75 ± 0.1 mW at 700 nm. The data points are obtained from the experiments, and the solid lines represent the theoretical fits.

Download Full Size | PDF

For data analysis, the n2 and β2PA values measured at same excitation wavelength but different laser powers were averaged and plotted against the Ep/Eg in Fig. 5, represented as the n2 and β2PA spectra for the MHP-ChGs. First of all, it is clear that no general overall n2-Ep/Eg and β2PA-Ep/Eg relationships regardless of the MHP composition exists, confirming that each MHP-ChG has an individual band structure, being a mixture of direct type of valence to conduction band transitions from the MHP NCs and indirect-type transitions of the amorphous host ChG. However, both n2 and β2PA values exhibit an increasing trend with increasing Ep/Eg. For the n2 spectra, this trend demonstrates that approaching excitation wavelength to bandgap edge would remarkably enhance n2 value of the material, which confirmed that free carriers generated at the sub-bandgap (namely Urbach band or defect states) region via 1PA process is a significant contribution. In particular, Br3 is explicitly higher n2 spectrum than the other two samples which are nearly overlapped. Such profile is the almost identical to the absorption spectra (scaled by Ep/Eg) as illustrated in inset of Fig. 5(a), and this is a direct support that the existence of exciton-phonon coupling [6,44] is an additional effect leading to the n2 enhancement for Br3. For the β2PA spectra, the three β2PA-Ep/Eg profiles are almost identical, and all of them equally exhibit a rapid increase of β2PA at Ep/Eg beyond 0.7. Such trend is analogous to the relativistic effects induced two-step upward trend in the 2PA spectra of MHP materials [7], which implies that the MHP NCs in ChG also have conduction bands split due to spin-orbit coupling of the heavy metal atoms, with heavy and light electronic states (HLES) above the secondary gap. As schematically indicated in inset of Fig. 5(b), the conduction band is an intermediate state for the electron transitions from the valence band to HLES with a rapidly enhancing 2PA process in the spectral region.

 figure: Fig. 5.

Fig. 5. (a) n2 versus Ep/Eg for the three MHP-ChG samples, inset is the corresponding absorption spectra in the same Ep/Eg range; (b) β2PA versus Ep/Eg for the three MHP-ChG samples, inset is schematic showing 2PA processes in conditions of Ep1 < ∼0.7Eg < Ep2, VB and CB are abbreviations for valence band and conduction band, respectively.

Download Full Size | PDF

On the other hand, it can be clearly distinguished from Fig. 5 that both n2 and β2PA values of the three samples at same Ep/Eg follow a spectral order of Br3 > Br2 > Br1. For β2PA, such compositional dependency is consistent with that of the CsPb(Br/I)3 NCs dispersed hexane at the excitation wavelength of 1064 nm [35]. For n2, such compositional dependency is consistent with the CsPb(Br/I)3 NCs dispersed anhydrous toluene in similar excitation wavelengths from 750 to 920 nm [30], but deviated from that of the CsPb(Br/I)3 NCs dissolved hexane at an excitation wavelength of 1064 nm [35] which is attributed to the different underlying mechanism for n2 at different excitation wavelength. Excitation at short wavelengths, i.e., this study and Ref. 30, resonant free carrier nonlinearity via 1PA process at sub-bandgap is the dominant contribution to the n2 value, thus stoichiometric MHPs with fewer carrier trapping defects exhibits higher n2 as compared to their non-stoichiometric counterparts. Excitation at long wavelength, i.e., Ref. 35, non-resonant bound electronic charge nonlinearity is the dominant contribution to the n2 value, thus MHPs with higher iodine content exhibit higher n2 because polarizability of iodine is higher than bromine.

At last, it is necessary to judge the functionality of the MHP-ChGs for the TONL-based applications via figure of merit (FOM), defined as the following express:

$$\textrm{FOM} = \frac{{n{}_2}}{{{\beta _{\textrm{2PA}}}\lambda }}$$
where λ is the excitation wavelength. This simple parameter manifests the balance between the nonlinear absorption and nonlinear refraction at a specifically wavelength. In general, materials meeting criterion of FOM > 1 are suitable for n2-based applications like all-optical switching; while those with the opposite FOM are suitable for β2PA-based applications, such as optical limiting. All the FOM values of the MHP-ChGs are listed in Table S1, nearly all of which are far less than 1, indicating the material is suitable for β2PA-based applications in the spectral region from 700 to 920 nm. Meanwhile, the FOM values are all laser power irrelevant due to Pauli blocking effect occurred in both n2 and β2PA sides. Figure 6 gives the mean FOM value (calculated from the mean n2 and β2PA values) as a function of Ep/Eg, represented as the FOM spectra of the MHP-ChGs. Br3 possesses relatively higher FOM for its larger n2, as a result of various factors as mentioned above. With decrease of Ep/Eg, a notable FOM maximum emerged in each spectrum. This means that the spin-orbit coupling boosted 2PA has discontinued below these threshold Ep/Eg points which are 0.75, 0.68 and 0.62 for Br1, Br2 and Br3 respectively, very close to the threshold value of 0.7 for CH3NH3PI3 polycrystalline films and 0.67 for CsPbI3 films reported in Ref. [7], indicating that the relativistic effect occurred in those MHP materials has no significant structural dependency.

 figure: Fig. 6.

Fig. 6. FOM versus Ep/Eg for the three MHP-ChG samples, the arrows note the FOM maximum in each spectrum.

Download Full Size | PDF

4. Conclusion

In summary, we prepared stable CsPb(Br/I)3 MHP NCs by packaging them in a GeS2-Sb2S3 ChG matrix, and characterized the TONL attribute (including n2 and β2PA) of the nano-composites via femtosecond Z-scan technique at excitation wavelengths from 700 to 920 nm. The experimental results demonstrate that the optical nonlinear attributes of MHP-ChGs are highly adjustable by exploiting the flexible tunability of the MHP composition as well as the excitation conditions. Firstly, the addition of MHP NCs could advance the intrinsic large optical nonlinearity of the host ChG, which is due the dielectric confinement induced local filing enhancement between the amorphous host ChG composed of covalent bonds and these crystalline MHP NCs composed of ionic bonds, and the nonlinear performance is much higher as compared to those of the MHP NCs embedded oxide glasses. Secondly, changing the composition of the MHP NCs from Br/I mixture to pristine Br further enhances the optical nonlinearity, as a result of the reduced carrier trapping defects in the stoichiometric MHP NCs. Thirdly, in the samples with pristine Br NCs, exciton-phonon coupling effect were discovered to promote its optical nonlinearity. Fourthly, the optical nonlinearity of all MHP-ChGs exhibits an overall decreasing trend with increasing excitation wavelength and power, which is attributed to spin-orbit coupling effects and Pauli blocking effects respectively. Finally, estimates of FOM manifested that the MHP-ChGs possess stronger 2PA than nonlinear refraction in the wavelength ranges from 700 to 920 nm, suggesting that they are promising materials for β2PA-based applications, such as optical limiting.

Funding

National Natural Science Foundation of China (61935006, 62075107, 62075108); K. C. Wong Magna Fund in Ningbo University.

Disclosures

The authors declare no conflicts of interest.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. J. P. Correa Baena, M. Saliba, T. Buonassisi, M. Grätzel, A. Abate, W. Tress, and A. Hagfeldt, “Promises and challenges of perovskite solar cells,” Science 358(6364), 739–744 (2017). [CrossRef]  

2. K. B. Lin, J. Xing, L. N. Quan, F. P. G. de Arquer, X. W. Gong, J. X. Lu, L. Q. Xie, W. J. Zhao, D. Zhang, C. Z. Yan, W. Q. Li, X. Y. Liu, Y. Lu, J. Kirman, E. H. Sargent, Q. H. Xiong, and Z. H. Wei, “Perovskite light-emitting diodes with external quantum efficiency exceeding 20 per cent,” Nature 562(7726), 245–248 (2018). [CrossRef]  

3. K. Sun, D. Tan, X. Fang, X. Xia, D. Lin, J. Song, Y. Lin, Z. Liu, M. Gu, Y. Yue, and J. Qiu, “Three-dimensional direct lithography of stable perovskite nanocrystals in glass,” Science 375(6578), 307–310 (2022). [CrossRef]  

4. W. Chen, S. Bhaumik, S. A. Veldhuis, G. Xing, Q. Xu, M. Grätzel, S. Mhaisalkar, N. Mathews, and T. C. Sum, “Giant five-photon absorption from multidimensional core-shell halide perovskite colloidal nanocrystals,” Nat. Commun. 8(1), 15198 (2017). [CrossRef]  

5. F. O. Saouma, C. C. Stoumpos, J. Wong, M. G. Kanatzidis, and J. I. Jang, “Selective enhancement of optical nonlinearity in two-dimensional organic-inorganic lead iodide perovskites,” Nat. Commun. 8(1), 742 (2017). [CrossRef]  

6. A. Manzi, Y. Tong, J. Feucht, E. P. Yao, L. Polavarapu, A. S. Urban, and J. Feldmann, “Resonantly enhanced multiple exciton generation through below-band-gap multi-photon absorption in perovskite nanocrystals,” Nat. Commun. 9(1), 1518 (2018). [CrossRef]  

7. Z. Wei, D. Guo, J. Thieme, C. Katan, V. M. Caselli, J. Even, and T. J. Savenije, “The importance of relativistic effects on two-photon absorption spectra in metal halide perovskites,” Nat. Commun. 10(1), 5342 (2019). [CrossRef]  

8. A. Ferrando, J. Martinez Pastor, and I. Suarez, “Toward metal halide perovskite nonlinear photonics,” J. Phys. Chem. Lett. 9(18), 5612–5623 (2018). [CrossRef]  

9. J. Xu, X. Li, J. Xiong, C. Yuan, S. Semin, T. Rasing, and X. H. Bu, “Halide perovskites for nonlinear optics,” Adv. Mater. 32(3), 1806736 (2020). [CrossRef]  

10. Y. Zhou, Y. Huang, X. Xu, Z. Fan, J. B. Khurgin, and Q. Xiong, “Nonlinear optical properties of halide perovskites and their applications,” Appl. Phys. Rev. 7(4), 041313 (2020). [CrossRef]  

11. W. Chen, F. Zhang, C. Wang, M. Jia, X. Zhao, Z. Liu, Y. Ge, Y. Zhang, and H. Zhang, “Nonlinear photonics using low-dimensional metal-halide perovskites: Recent advances and future challenges,” Adv. Mater. 33(11), 2004446 (2021). [CrossRef]  

12. G. Wang, S. Mei, J. Liao, W. Wang, Y. Tang, Q. Zhang, Z. Tang, B. Wu, and G. Xing, “Advances of nonlinear photonics in low-dimensional halide perovskites,” Small 17(43), 2100809 (2021). [CrossRef]  

13. W. Shen, J. Chen, J. Wu, X. Li, and H. Zeng, “Nonlinear optics in lead halide perovskites: Mechanisms and applications,” ACS Photonics 8(1), 113–124 (2021). [CrossRef]  

14. B. J. Eggleton, “Chalcogenide photonics: Fabrication, devices and applications Introduction,” Opt. Express 18(25), 26632–26634 (2010). [CrossRef]  

15. L. Li, H. Lin, S. Qiao, Y. Zou, S. Danto, K. Richardson, J. D. Musgraves, N. Lu, and J. Hu, “Integrated flexible chalcogenide glass photonic devices,” Nat. Photonics 8(8), 643–649 (2014). [CrossRef]  

16. C. Lin, C. Rüssel, and S. Dai, “Chalcogenide glass-ceramics: Functional design and crystallization mechanism,” Prog. Mater. Sci. 93, 1–44 (2018). [CrossRef]  

17. G. Shao, S. Liu, L. Ding, Z. Zhang, W. Xiang, and X. Liang, “KxCs1−xPbBr3 NCs glasses possessing super optical properties and stability for white light emitting diodes,” Chem. Eng. J. 375, 122031 (2019). [CrossRef]  

18. M. Jin, S. Huang, C. Quan, X. Liang, J. Du, Z. Liu, Z. Zhang, and W. Xiang, “Blue low-threshold room-temperature stimulated emission from thermostable perovskite nanocrystals glasses through controlling crystallization,” J. Eur. Ceram. Soc. 41(2), 1579–1585 (2021). [CrossRef]  

19. M. Jin, X. Liang, H. Zhang, J. Liu, G. Shao, W. Xiang, and Y. Song, “A promising optical limiting material: Tunable third-order nonlinear optical properties of robust CsPbX3 (X = Cl/Br, Br) nanocrystals glasses,” J. Eur. Ceram. Soc. 40(12), 4140–4147 (2020). [CrossRef]  

20. C. Quan, X. Xing, S. Huang, M. Jin, T. Shi, Z. Zhang, W. Xiang, Z. Wang, and Y. Leng, “Nonlinear optical properties of CsPbClxBr3-x nanocrystals embedded glass,” Photonics Res. 9(9), 1767–1774 (2021). [CrossRef]  

21. Z. Xu, T. Chen, D. Zhang, G. Zheng, J. Wu, J. Yan, X. Liu, and J. Qiu, “Linear and nonlinear optical characteristics of CsPbBr3 perovskite quantum dots-doped borosilicate glasses,” J. Eur. Ceram. Soc. 41(1), 729–734 (2021). [CrossRef]  

22. K. Weng, N. Long, Y. Guo, Q. Jiao, S. Dai, and C. Lin, “Nanocrystallization of α-CsPbI3 perovskite nanocrystals in GeS2-Sb2S3 based chalcogenide glass,” J. Eur. Ceram. Soc. 40(12), 4148–4152 (2020). [CrossRef]  

23. K. Weng, N. Long, Q. Jiao, J. Ren, C. Liu, H. Tao, S. Dai, and C. Lin, “A modified chalcogenide flux method for confining metal halide nanocrystals into transparent glassy matrix,” J. Eur. Ceram. Soc. 40(15), 6037–6042 (2020). [CrossRef]  

24. N. Long, Y. Fu, T. Xu, D. Ding, S. Zhang, S. Sun, S. Kang, T. Xu, S. Dai, Q. Nie, and C. Lin, “Nanocrystallization and optical properties of CsPbBr3−xIx perovskites in chalcogenide glasses,” J. Eur. Ceram. Soc. 41(8), 4584–4589 (2021). [CrossRef]  

25. M. Green, J. Y. Jiang, A. Mahboubi Soufiani, and A. Ho Baillie, “Optical properties of photovoltaic organic-inorganic lead halide perovskites,” J. Phys. Chem. Lett. 6(23), 4774–4785 (2015). [CrossRef]  

26. L. Petit, N. Carlie, K. Richardson, A. Humeau, S. Cherukulappurath, and G. Boudebs, “Nonlinear optical properties of glasses in the system Ge/Ga-Sb-S/Se,” Opt. Lett. 31(10), 1495–1497 (2006). [CrossRef]  

27. K. Fedus, G. Boudebs, Q. Coulombier, J. Troles, and X. H. Zhang, “Nonlinear characterization of GeS2-Sb2S3-CsI glass system,” J. Appl. Phys. 107(2), 023108 (2010). [CrossRef]  

28. W. G. Lu, C. Chen, D. Han, L. Yao, J. Han, H. Zhong, and Y. Wang, “Nonlinear optical properties of colloidal CH3NH3PbBr3 and CsPbBr3 quantum dots: A comparison study using Z-scan technique,” Adv. Opt. Mater. 4(11), 1732–1737 (2016). [CrossRef]  

29. K. N. Krishnakanth, S. Seth, A. Samanta, and S. V. Rao, “Broadband femtosecond nonlinear optical properties of CsPbBr3 perovskite nanocrystals,” Opt. Lett. 43(3), 603–606 (2018). [CrossRef]  

30. J. Szeremeta, M. A. Antoniak, D. Wawrzyńczyk, M. Nyk, and M. Samoć, “The two-photon absorption cross-section studies of CsPbX3 (X = I, Br, Cl) nanocrystals,” Nanomaterials 10(6), 1054 (2020). [CrossRef]  

31. S. Liu, G. Chen, Y. Huang, S. Lin, Y. Zhang, M. He, W. Xiang, and X. Liang, “Tunable fluorescence and optical nonlinearities of all inorganic colloidal cesium lead halide perovskite nanocrystals,” J. Alloys Compd. 724, 889–896 (2017). [CrossRef]  

32. C. H. Kwak, Y. L. Lee, and S. G. Kim, “Analysis of asymmetric Z-scan measurement for large optical nonlinearities in an amorphous As2S3 thin film,” J. Opt. Soc. Am. B 16(4), 600–604 (1999). [CrossRef]  

33. B. S. Kalanoor, L. Gouda, R. Gottesman, S. Tirosh, E. Haltzi, A. Zaban, and Y. Tischler, “Third-order optical nonlinearities in organometallic methylammonium lead iodide perovskite thin films,” ACS Photonics 3(3), 361–370 (2016). [CrossRef]  

34. R. Li, Z. Wei, H. Zhao, H. Yu, X. Fang, D. Fang, J. Li, T. He, R. Chen, and X. Wang, “Linear and nonlinear optical characteristics of all-inorganic perovskite CsPbBr3 quantum dots modified by hydrophobic zeolites,” Nanoscale 10(48), 22766–22774 (2018). [CrossRef]  

35. I. Suárez, M. Vallés Pelarda, A. Gualdrón-Reyes, I. Mora Seró, A. Ferrando, H. Michinel, J. R. Salgueiro, and J. P. M. Pastor, “Outstanding nonlinear optical properties of methylammonium- and Cs-PbX3 (X = Br, I, and Br-I) perovskites: Polycrystalline thin films and nanoparticles,” APL Mater. 7(4), 041106 (2019). [CrossRef]  

36. J. Li, H. Dong, B. Xu, S. Zhang, Z. Cai, J. Wang, and L. Zhang, “CsPbBr3 perovskite quantum dots: Saturable absorption properties and passively Q-switched visible lasers,” Photonics Res. 5(5), 457–460 (2017). [CrossRef]  

37. Y. Wang, X. Li, X. Zhao, L. Xiao, H. Zeng, and H. Sun, “Nonlinear absorption and low-threshold multiphoton pumped stimulated emission from all-inorganic perovskite nanocrystals,” Nano Lett. 16(1), 448–453 (2016). [CrossRef]  

38. J. Zhang, T. Jiang, X. Zheng, C. Shen, and X. Cheng, “Thickness-dependent nonlinear optical properties of CsPbBr3 perovskite nanosheets,” Opt. Lett. 42(17), 3371–3374 (2017). [CrossRef]  

39. F. Saouma, C. Stoumpos, M. Kanatzidis, Y. soo Kim, and J. Jang, “Multiphoton absorption order of CsPbBr3 as determined by wavelength-dependent nonlinear optical spectroscopy,” J. Phys. Chem. Lett. 8(19), 4912–4917 (2017). [CrossRef]  

40. J. Yi, L. Miao, J. Li, W. Hu, C. Zhao, and S. Wen, “Third-order nonlinear optical response of CH3NH3PbI3 perovskite in the mid-infrared regime,” Opt. Mater. Express 7(11), 3894–3901 (2017). [CrossRef]  

41. R. Zhang, J. Fan, X. Zhang, H. Yu, H. Zhang, T. Xu, J. Wang, and H. Snaith, “Nonlinear optical response of organic-inorganic halide perovskites,” ACS Photonics 3(3), 371–377 (2016). [CrossRef]  

42. M. Sheik Bahae, A. A. Said, T. H. Wei, D. J. Hagan, and E. W. van Stryland, “Sensitive measurement of optical nonlinearities using a single beam,” IEEE J. Quantum Electron. 26(4), 760–769 (1990). [CrossRef]  

43. H. T. Guo, C. Q. Hou, F. Gao, A. X. Lin, P. F. Wang, Z. G. Zhou, M. Lu, W. Wei, and B. Peng, “Third-order nonlinear optical properties of GeS2-Sb2S3-CdS chalcogenide glasses,” Opt. Express 18(22), 23275–23284 (2010). [CrossRef]  

44. K. Ohara, T. Yamada, H. Tahara, T. Aharen, H. Hirori, H. Suzuura, and Y. Kanemitsu, “Excitonic enhancement of optical nonlinearities in perovskite CH3NH3PbCl3 single crystals,” Phys. Rev. Mater. 3(11), 111601 (2019). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       The figure of SEM, XRD and the nonlinear refractive and nonlinear absorption versus laser power.

Data Availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. Broad-band transmission spectra of the three MHP-ChGs and the host ChG. The inset is a photograph of the four samples.
Fig. 2.
Fig. 2. Absorption spectra of the three MHP-ChGs. The definition of Eg and excitation wavelengths for Z-scan measurements are noted.
Fig. 3.
Fig. 3. CA and OA Z-scans under laser power of 10 ± 1 mW at wavelengths from 700 to 920 nm for Br1 (a) and (d), Br2 (b) and (e), Br3 (c) and (f), respectively. The data points are obtained from the experiments, and the solid lines represent the theoretical fits.
Fig. 4.
Fig. 4. CA (a) and OA (b) Z-scans of the three MHP-ChGs and the host ChG under laser power of 3.75 ± 0.1 mW at 700 nm. The data points are obtained from the experiments, and the solid lines represent the theoretical fits.
Fig. 5.
Fig. 5. (a) n2 versus Ep/Eg for the three MHP-ChG samples, inset is the corresponding absorption spectra in the same Ep/Eg range; (b) β2PA versus Ep/Eg for the three MHP-ChG samples, inset is schematic showing 2PA processes in conditions of Ep1 < ∼0.7Eg < Ep2, VB and CB are abbreviations for valence band and conduction band, respectively.
Fig. 6.
Fig. 6. FOM versus Ep/Eg for the three MHP-ChG samples, the arrows note the FOM maximum in each spectrum.

Tables (1)

Tables Icon

Table 1. Host material, style, bandgap energy and nonlinear optical parameters of the CsPbBr3 MHP materials in this work and others’ reports

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

FOM = n 2 β 2PA λ
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.