• Photonics Research
  • Vol. 13, Issue 2, 286 (2025)
Longxing Su1、2、*, Bingheng Meng3, Heng Li2, Zhuo Yu2, Yuan Zhu2、5, and Rui Chen4、6
Author Affiliations
  • 1International School of Microelectronics, Dongguan University of Technology, Dongguan 523808, China
  • 2School of Microelectronics, Southern University of Science and Technology, Shenzhen 518055, China
  • 3State Key Laboratory of High Power Semiconductor Laser, School of Physics, Changchun University of Science and Technology, Changchun 130022, China
  • 4Department of Electrical and Electronic Engineering, Southern University of Science and Technology, Shenzhen 518055, China
  • 5e-mail: zhuy3@sustech.edu.cn
  • 6e-mail: chenr@sustech.edu.cn
  • show less
    DOI: 10.1364/PRJ.539352 Cite this Article Set citation alerts
    Longxing Su, Bingheng Meng, Heng Li, Zhuo Yu, Yuan Zhu, Rui Chen, "Amplified spontaneous emission and photoresponse characteristics in highly defect tolerant CsPbClxBr3−x crystal," Photonics Res. 13, 286 (2025) Copy Citation Text show less

    Abstract

    All inorganic perovskite CsPbX3 with excellent optical properties and a tunable bandgap is a potential candidate for optoelectronic applications, and the amplified spontaneous emission (ASE) is normally reported in low-dimensional structures where the quantum confinement enhances ASE. Herein, we not only demonstrate the ASE in millimeter size CsPbClxBr3-x crystal with a high defect concentration, but also tune the emission wavelength from the green band to blue band through the ion exchange of Br with Cl. The ASE centered at 456 nm is probed at 50 K with a threshold of 106 μJ/cm2. Furthermore, a metal-semiconductor-metal (MSM) structure CsPbClxBr3-x photodetector is fabricated and shows a distinct response to lights from UV to the blue band; the response spectrum range is quite different from the narrow band (30 nm) response of the CsPbBr3 photodetector induced by a charge collection narrowing (CCN) mechanism. The CsPbClxBr3-x photodetector also exhibits fast response speeds with a rise time of 96 μs and a decay time of 34 μs, indicating the defects have limited influence on the transportation speed of the photo-generated carriers.

    1. INTRODUCTION

    Recently, all inorganic lead halide perovskites (CsPbX3, X=Cl, Br, I) have received tremendous attention in optoelectronic devices because of their advantages including a large light absorption coefficient, low fabrication cost, tunable bandgap, and high carrier mobility [13]. The most interesting feature of this group of material is the ability to obtain high optical gain in the whole visible range, making them highly desirable in tunable lasers, broad band amplifiers, and response band adjustable photodetectors [46]. Generally, mixed halide perovskites (CsPbClxBr3x or CsPbBrxI3x) can be prepared by adding or mixing appropriate PbX2 salts as precursors; thus the bandgap is correspondingly modulated from near ultraviolet (UV) to the red spectrum through fast anion exchange [7]. Huang et al. have synthesized a series of CsPbClxBr3x (x=02.5) samples with emission wavelength from 525 nm to 430 nm through a solution method [8]. A vapor-assisted post-synthesis chlorination procedure is another efficient strategy to convert the green emitting CsPbBr3 into blue emitting CsPbClxBr3x through controlling the conversion temperature and time [9]. Typically, the low-dimensional structures like quantum dots, nanocrystals, and microplates are promising in realizing the low-threshold amplified spontaneous emission (ASE) because of the quantum confinement effect [10]. CsPbClxBr3x quantum dots and nanofilm have proved their ASE behavior with pumping thresholds of 2056  μJ/cm2, and the ASE peak centers are tuned to the blue band (440–495 nm) [1113]. Compared with the low-dimensional structures, a large sized crystal is more difficult in the realization of ASE but has more potential in practical application. Kim et al. report the one-photon pumped ASE in CsPbBr3 single crystals, while the limited penetration depth and reabsorption of the crystal induce a large threshold of 1.38  mJ/cm2 [14]. Zhao et al. reduce the ASE threshold (0.65  mJ/cm2) in millimeter sized CsPbBr3 crystal using two-photon excitation, but the full width at half maximum (FWHM) is as large as 7 nm [15]. To date, studies regrading blue band ASE in millimeter sized lead halide perovskites are still rare. On the other side, as another kind of optoelectronic device that can convert the light into electrical signal, an all inorganic lead halide perovskite (CsPbX3, X=Cl, Br, I) photodetector plays a key role in optical imaging, environmental monitoring, forest fire alarms, and space secure communication [1618]. However, a strong and narrow response band is normally observed in single crystal perovskites [19,20], which can be attributed to the CCN mechanism [21]. Therefore, it is highly desired if a low-threshold ASE and broadband photoresponse can be simultaneously realized within a large sized CsPbX3 crystal.

    In this work, millimeter sized CsPbClxBr3x crystals are synthesized through a simple anti-solvent process at room temperature. High surface defect density is observed and determined in the as-prepared sample with low photoluminescence quantum yield (PLQY) of only 0.35%. Nevertheless, ASE behavior still can be observed with a relatively low threshold of 106  μJ/cm2 at 50 K, which verifies the high defect tolerance of the all inorganic perovskite. In addition, compared with the emission center of CsPbBr3 crystal, the ASE emission peak blue shifts from 529 to 456  nm because of the anion exchange of Br by Cl. Subsequently, a symmetric MSM type photodetector based on CsPbClxBr3x is prepared with InGa as the Ohmic contact. At 10 V, the photodetector demonstrates good performances with a responsivity of 3.95 mA/W, a detectivity of 7×1010 Jones, and fast response speeds of 96 μs (rise time) and 34 μs (decay time).

    2. EXPERIMENTAL SECTION

    A. Chemicals and Reagents

    Cesium bromide (CsBr, 99.9%), cesium chloride (CsCl, 99.9%), and lead bromide (PbBr2, 99.9%) were purchased from Alfa Aesar (China) Co., Ltd. Dimethyl sulfoxide (DMSO, 99.5%), methanol (99.5%), and ethanol (99.5%) were purchased from Sinopharm Chemical Reagent Co., Ltd. All reagents and solvents were used without any purification.

    B. Synthesis of CsPbClxBr3−x Crystals

    The CsPbClxBr3x sample was synthesized by an infiltrating inverse solvent re-crystallization method. 0.93 g CsBr, 0.93 g CsCl, and 0.93 g PbBr2 were simultaneously loaded in 8 mL DMSO solution and then stirred for 1 h. After precipitating for 1 h, the supernatant was transferred into a vial, and then the vial was loaded inside a beaker filled with methanol. Subsequently, the beaker was sealed by paraffin film in order to maintain the methanol atmosphere. The growth temperature is at room temperature, and yellow CsPbClxBr3x crystals were obtained after 5 days. Finally, the CsPbClxBr3x crystals were washed by ethanol and dried in the oven at 60°C for 30 min. It is worth noting that the repeatability of the method at a small-batch scale (gram scale) is good. CsPbClxBr3x crystals with the similar size and the same component ratio can be collected within different batches under the same growth parameters. According to a previous study on the production of kilogram-scale PbI2 crystals using the similar solution method [22], large-scale production of CsPbClxBr3x crystals for practical applications is expected with the upgrade and expansion of the synthesized facilities.

    C. Characterizations

    The XRD pattern of the CsPbClxBr3x sample was collected using a multifunctional X-ray diffractometer (XRD, Bruker, D8 Advance) with a CuKα line (1.54 Å). The Raman spectrum of the CsPbClxBr3x sample was excited by a 532 nm laser and collected through a back scattering configuration. The room temperature photoluminescence (PL) spectrum was measured to study the luminescence behavior of the CsPbClxBr3x sample. The photoluminescence quantum yield (PLQY) was determined (Edinburgh Instruments, FLS-1000) to study the PL efficiency of CsPbClxBr3x. X-ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific, ESCALAB XI) and ultraviolet photoelectron spectroscopy (UPS) were used to investigate the chemical state and energy band structure.

    D. ASE Measurement

    A 355 nm laser (100 fs, 1 kHz) with different densities was used as the excitation source to excite the CsPbClxBr3x sample. The measurements were conducted at both room temperature and 50 K. Temperature-dependent PL spectra under excitation density of 5.9  μJ/cm2 were also collected. During the measurement, the emission was collected by a spectrometer with resolution of 0.09 nm (Princeton Instrument, SpectraPro HRS-300).

    E. Device Fabrication and Measurement

    InGa electrodes were employed as the contact electrodes for the CsPbClxBr3x MSM photodetector. The response characteristics of the device were investigated by using an electrochemical workstation (Ivium Vertex One). A Xe-lamp equipped with a monochromator was utilized as the excitation source. The photoresponse measurements were performed in an air environment with a relative humidity (RH) of 57%. The time-resolved photoresponse curve was measured by using a continuous-wave (CW) laser with 442 nm wavelength as the excited source and an oscilloscope as the data collector. The CW laser was chopped and modulated by a programmable module.

    3. RESULTS AND DISCUSSION

    The XRD pattern of the as-prepared CsPbClxBr3x crystal is shown in Fig. 1(a). Eight distinct peaks located at 15.59°, 15.7°, 21.17°, 31.38°, 31.6°, 35.37°, 35.56°, and 38.7° are detected. By comparing with the standard cards of CsPbBr3 (PDF #18-0364) and CsPbCl3 (PDF #18-0366), these peaks can be ascribed to the diffraction from the (001), (100), (101), (002), (200), (102), (201), and (112) facets of CsPbClxBr3x. The crystal constants of CsPbBr3 (5.827  ×5.827  ×5.891  ) are slightly larger than that of CsPbCl3 (5.584  ×5.584  ×5.623  ). Thus, compared with the standard XRD peaks of pure CsPbBr3, these eight XRD peaks shift to the larger angle side, and compared with the XRD peaks of pure CsPbCl3, they locate at the left angle side. Compared with the extremely high-crystal-quality hybrid perovskite MAPbX3 single crystal [23], the growth parameters including synthesized temperature, nucleation rate, reagent ratio, and additives need to be optimized in order to further improve the quality of CsPbClxBr3x crystal in the future. The optimized crystal quality can greatly improve the performance of the CsPbClxBr3x based optoelectronic devices. Figure 1(b) presents the Raman spectrum of the as-prepared CsPbClxBr3x sample, from which only two distinct Raman peaks centered at 93.6  cm1 and 167.1  cm1 are probed. According to previous Raman studies on CsPbCl3 [24,25], the first vibrational peak is relevant to the Pb(ClxBr1x)6 octahedron and the second one is associated with the motion of Cs+ cations. Compared with the Raman peaks on CsPbBr3, blue shifts can be determined owing to the substitution of a Br atom by a Cl atom [26,27]. The room temperature PL spectrum of the as-prepared CsPbClxBr3x crystal is exhibited in Fig. 1(c), from which a broad blue band emission is probed. Two distinct peaks denoted as peak A and peak B are detected; the former one is relevant to the bandgap emission and the latter one is induced by the donor defects. The optical images of the as-synthesized CsPbClxBr3x crystal are presented in Fig. 7 (Appendix A), suggesting a great number of crystal boundaries located within the surface of the sample. This will introduce a high concentration of surface defects, which supports the result from the PL measurement. It is worth noting that one key advantage in this work is the simple and low-cost equipment; only two beakers with different capacities are required, which makes the equipment easily upgraded. Then the CsPbClxBr3x crystals can be mass produced once the synthesized processes are mature. In addition, different from previous efforts on the high-quality perovskite crystal, defects and grain boundaries are intentionally introduced in order to study the defect tolerance of the optical pumping property, which will be discussed later. The bandgap of CsPbClxBr3x at room temperature is estimated as 2.76  eV (448  nm) by calculating the wavelength center of the bandgap emission, which is 0.45  eV larger than the bandgap of CsPbBr3 [28] and 0.31  eV smaller than the bandgap of CsPbCl3 [29,30]. Quantum efficiency is one important parameter to measure the performance of emitters and the corresponding PLQY result is exhibited in Fig. 1(d). The PLQY of the sample is as low as 0.35%, which is quite similar to previous work reported by Gong et al. [31]; they only detect a low PLQY of 1% in a CsPbBr3 single crystal. Typically, different from CsPbX3 nanocrystals or quantum dots, bulk single crystal usually displays very low PLQY (typically 1% or even lower). One reason is owing to the presence of indirect tail states below the direct transition edge caused by Rashba splitting in bulk CsPbBr3 single crystal [32]. The other reason of the low PLQY is the high defect concentration in our CsPbClxBr3x crystal. Theoretically, the improvement of crystal quality is one effective strategy for improving the PLQY of CsPbClxBr3x crystal. A lower defect concentration can naturally decrease the nonradiative recommendation centers, which brings the positive influence on the PLQY. Another strategy for improving the PLQY is decreasing the dimension of CsPbClxBr3x to quantum size (few nanometers). However, our aim is focusing on the large sized crystal; thus improving the crystal quality of CsPbClxBr3x crystals will be performed. Four strategies are considered in the future: (i) the synthesized temperature is enhanced from room temperature to 60°C–80°C, which provides sufficient thermal energy for the chemical reaction; (ii) a post-annealing process in the inert or halogen/lead atmosphere is considered; (iii) a specific organic molecule is selected and acts as the passivator of defects; (iv) the Bridgman method operated at high temperature is considered for growing high-quality and large sized CsPbClxBr3x crystal, and then the sample is cut into the targeted size and polished. The high crystal quality will improve the PLQY of the crystal, and then decrease the threshold of the ASE. For photodetectors, undoubtedly, high crystal quality can lower the dark current of the device, and enhance the quantum efficiency and responsivity of the device. Especially, high crystal quality can naturally avoid the effect of trap states on the carrier dynamic, which contributes to the rapid response speed of a photodetector.

    The room temperature (a) XRD pattern, (b) Raman spectrum, (c) PL spectrum, and (d) PLQY of the as-synthesized CsPbClxBr3−x crystal.

    Figure 1.The room temperature (a) XRD pattern, (b) Raman spectrum, (c) PL spectrum, and (d) PLQY of the as-synthesized CsPbClxBr3x crystal.

    Figures 2(a)–2(d) present XPS spectra of Cs-3d, Pb-4f, Cl-2p, and Br-3d core electrons, in which C-1s peak (284.8 eV) is used as calibration. Apparently, all the orbitals can be well fitted with two standard Gaussian peaks, indicating the high-purity phase of CsPbClxBr3x crystal. The spin-orbital splittings (Cs-3d5/2 and Cs-3d3/2; Pd-4f7/2 and Pd-4f5/2; Cl-2p3/2 and Cl-2p1/2; Br-3d5/2 and Br-3d3/2) are calculated to be 14 eV, 4.87 eV, 1.6 eV, and 1.04 eV, agreeing well with the results from the standard XPS database. The energy band diagram is the key factor for semiconductor devices. The valence band scanning [Fig. 2(e)] and UPS measurement [Fig. 2(f)] are employed to determine the band structure of CsPbClxBr3x. Deriving from the cutoff and onset of the UPS spectrum, the work function of CsPbClxBr3x is calculated as 3.8  eV and the Fermi energy level locates 2.22  eV above the VBM. The UPS study aligns well with the Fermi energy level (EFEVBM=2.3  eV) confirmed by the valence band scanning spectrum. Therefore, the energy band diagram of CsPbClxBr3x can be summarized in Fig. 8 (Appendix A). The energy difference between VBM and Pb-5d7/2 orbital is 17 eV, and the Fermi energy level locates close to CBM.

    The XPS spectra of (a) Cs-3d, (b) Pb-4f, (c) Cl-2p, and (d) Br-3d core electrons; (e) the valence band scanning spectrum of CsPbClxBr3−x relative to the XPS spectrum of Pb-5d7/2 core electron; (f) the UPS spectrum of CsPbClxBr3−x.

    Figure 2.The XPS spectra of (a) Cs-3d, (b) Pb-4f, (c) Cl-2p, and (d) Br-3d core electrons; (e) the valence band scanning spectrum of CsPbClxBr3x relative to the XPS spectrum of Pb-5d7/2 core electron; (f) the UPS spectrum of CsPbClxBr3x.

    Figure 3(a) displays the PL spectra under various excitation densities at room temperature, in which two distinct emissions (peak A and peak B) can be clearly observed. As further summarized in Fig. 3(b), the PL intensity at room temperature can be expressed by a power law (IPα), where I is the PL intensity, P is the excitation density, and α is the exponent factor. The α factors are calculated to be 0.34 (peak A) and 0.45 (peak B), which are smaller than one. Theoretically, when 1<α<2, excitonic recombination dominates; when α is less than one, an impurity or defect (donor and acceptor) is involved in the transitions [3336]. Thus, the small exponent factors indicate a large proportion of radiative recombination induced by defects. Theoretically, three factors including growth temperature, the ratio of the reaction solutes, and the growth rate (decided by the evaporation rate of methanol) have direct influence on the crystal quality. Low temperature (RT) cannot supply sufficient thermal energy for the migration of atoms, which will easily introduce defects in the crystal. Additionally, the methanol atmosphere can also affect the nucleation rate of CsPbClxBr3x crystals. The large diffusion rate of methanol atmosphere at a higher temperature will lead to the reduction of the growth time and also affect the crystal quality of the CsPbClxBr3x crystals. Nevertheless, in this work, CsPbClxBr3x crystals with a high defect concentration are intentionally synthesized, which is demonstrated by the low α value. As previously mentioned, a high concentration of defects is probed in the as-synthesized CsPbClxBr3x crystal, agreeing well with the emission spectra at room temperature. Figure 3(c) presents the excitation-intensity-dependent photon energies of peak A and peak B; both peak energies monotonically decrease with the increase of the excitation density. Typically, thermally induced lattice expansion can naturally reduce the bandgap of semiconductors [37,38]; thus the accumulated heat under higher excitation density is responsible for the red shift of the emission peaks at room temperature. Figure 3(d) shows the temperature-dependent PL spectra of the CsPbClxBr3x crystal at a low excitation (5.9  μJ/cm2). With the increase of temperature, a blue shift of peak A and red shift of peak B are observed, which agrees well with previous reports [39,40]. Figure 3(e) summarizes the PL intensity at 5.9  μJ/cm2 as a function of the temperature. The PL intensity of peak B becomes larger than that of peak A when the temperature is below 220 K. Figure 3(f) presents the temperature-dependent photon energies of peak A and peak B under the excitation density of 5.9  μJ/cm2. A slight blue shift on peak A and distinct red shift on peak B can be probed. The blue shift of peak A (bandgap emission) is unusual by comparing with conventional semiconductors like CdSe [41], ZnO [42], and GaN [43], but it is quite normal in perovskites [4446]. A blue shift of 10  meV (peak A) is determined with temperature increasing from 50°C to 295°C. Generally, the temperature-dependent bandgap can be well fitted by a Varshni model [45,47]: Eg(T)=E(0)(γ×T2)/(β+T),where Eg(T) and E(0) are the photon energies at T and 0 K; γ and β are shift parameters. The Varshni fit yields a γ parameter of 3.74×105  K1. The negative value indicates the slight blue shift characteristic, and the β parameter at 300 K is corresponding to the Debye temperature. This behavior can be dominated by the bandgap renormalization effect in perovskite, which is completed by the thermal expansion and the interaction with phonons [48,49]. Normally, thermal expansion in perovskite materials will lead to the blue shift of the bandgap, while the interaction with phonons will induce the red shift of the bandgap [50,51]. Both volume dilation and excitation of the lattice vibration (phonon) are relevant to the temperature, which can produce the shifts in energy levels and then change the energy bandgap. The electron-phonon interaction can induce the shift of the valence band and conduction band, resulting in a quadratic variation of the energy gap at low temperature and a linear shift at high temperature [50]. The conduction band is estimated from the electron and the filled band should be decided by the hole with a negative sign in the effective mass. For perovskite materials, volume dilation has a higher effect on producing the temperature variation of the bandgap than the lattice vibrations. Therefore, a blue shift induced by thermal expansion dominates over the bandgap renormalization, thus leading to a blue shift of the bandgap as the temperature increases. Undoubtedly, the bandgap blue shift with the increase of temperature has been expected in a great number of perovskite materials.

    (a) The room temperature PL spectra of the CsPbClxBr3−x crystal under different excitation densities; (b) the relationship between the peak intensity and the excitation density; (c) the photon energy as a function of the excitation density; (d) temperature-dependent PL spectra of the CsPbClxBr3−x crystal under the excitation density of 5.9 μJ/cm2; (e) the emission intensity as a function of the tested temperature; (f) temperature-dependent photon energies of peak A and peak B.

    Figure 3.(a) The room temperature PL spectra of the CsPbClxBr3x crystal under different excitation densities; (b) the relationship between the peak intensity and the excitation density; (c) the photon energy as a function of the excitation density; (d) temperature-dependent PL spectra of the CsPbClxBr3x crystal under the excitation density of 5.9  μJ/cm2; (e) the emission intensity as a function of the tested temperature; (f) temperature-dependent photon energies of peak A and peak B.

    (a) The ASE from the CsPbClxBr3−x crystal under various pumping densities; (b) the PL intensity (black) and FWHM (blue) as a function of the pumping density; (c) the ASE peak energy as a function of the pumping density.

    Figure 4.(a) The ASE from the CsPbClxBr3x crystal under various pumping densities; (b) the PL intensity (black) and FWHM (blue) as a function of the pumping density; (c) the ASE peak energy as a function of the pumping density.

    The schematic diagram of the CsPbClxBr3x photodetector is presented in Fig. 5(a), from which InGa metal is employed as the bottom electrodes. Figure 5(b) shows the I-V curves of the photodetector under dark and illumination by 400 nm light with different intensities. The nearly linear relationship between applied voltage and dark current indicates the Ohmic contact between InGa and CsPbClxBr3x. According to previous UPS measurement, the work function of CsPbClxBr3x is determined to be 3.8  eV, which is close to that of the InGa electrode (4.1  eV) [62]. Combining the surface defect states in the sample, the Ohmic contact property can be naturally expected between the metal and semiconductor. Under irradiation of the 400 nm light [Fig. 5(b)], the photocurrent prominently increases, and it is enhanced with the increase of the light intensity. The response speed is also an important parameter for the photodetector, which represents its signal tracking ability. As presented in Fig. 5(c), the response curve is asymmetrical with a rise time of 96 μs and a decay time of 34 μs, which reveals a fast response performance. The 3 dB bandwidth can be further calculated as 140 kHz. Additionally, the decay of the response curve can be well fitted with a second order exponential equation [63]: y=y0+A1eτ/τ1+A2eτ/τ2,where y0 is the maximum voltage, A1 and A2 are fitting constants, and τ1 and τ2 are the first order decay time and second order decay time. As exhibited in Fig. 5(d), τ1 and τ2 are well fitted as 25.6 μs and 31.3 μs, which means two physical processes are responsible for the recovery time once the illumination is turned off. Nevertheless, both processes show fast response behavior. As discussed above, the defect energy level of VBr is shallow in CsPbBr3 while that of VCl is relatively deep in CsPbCl3; the deep level of VCl may bring a negative effect on the recovery time of the photodetector. However, in the CsPbClxBr3x crystal, the deep-level property of VCl will be partly suppressed. Therefore, the fast response speed of the device can still be obtained in the low-quality CsPbClxBr3x crystals. Typically, three factors including the RC time constant, the transit time, and the excess life time of the trap carriers are responsible for the decay trace [64]. Herein, the RC time constant is originated from the large capacitance from the PN junction or Schottky junction, which can be ruled out in our photoconductive device. The transit time is determined by the space of two electrodes, the applied voltage, and the carrier mobility [65]. According to the parameters of the device, the transit time is estimated as 27.9 μs, which may be the most possible reason for the τ1 value. Additionally, the trapped carriers induced by the defects in the surface or the grain boundaries will lead to the excess life time as the incident light is turned off; this mean life time can be indexed to the τ2 time constant. Therefore, the decay time of the photodetector is contributed by the transit time and the excess life time of the trap carriers induced by defects.

    (a) The schematic diagram of the CsPbClxBr3−x photodetector; (b) the I-V curves of the CsPbClxBr3−x photodetector under dark and 400 nm light illumination; (c) the fast response V-t measurement and (d) the fitting of the decay trace.

    Figure 5.(a) The schematic diagram of the CsPbClxBr3x photodetector; (b) the I-V curves of the CsPbClxBr3x photodetector under dark and 400 nm light illumination; (c) the fast response V-t measurement and (d) the fitting of the decay trace.

    The energy band diagram of the symmetric structure InGa/CsPbClxBr3x/InGa photodetector can be observed in Fig. 6(a). The Fermi energy level locates 0.54  eV below the CBM, implying the n-type conductivity of the perovskite. The energy band offset between the work functions of InGa and CsPbClxBr3x is as small as 0.3  eV. Considering the energy level broadening and surface state or defect on the surface of CsPbClxBr3x, the shallow barrier can be easily ignored. So, the Ohmic contact property is probed in Fig. 5(b). Figure 6(b) exhibits the comparison of response spectra between the CsPbBr3 photodetector and the CsPbClxBr3x photodetector. Apparently, substitution of Br by Cl can widen the bandgap of CsPbBr3; the response spectrum range of CsPbClxBr3x is correspondingly modulated toward the shorter wavelength region. Additionally, the CsPbBr3 single crystal normally reveals a narrow band (30  nm) response characteristic, which is explained by the CCN mechanism [21]. Nevertheless, this phenomenon is not probed in the CsPbClxBr3x photodetector, which harvests broad spectra from UV to the blue band. This may be ascribed to the change of band structure as Br is substituted by Cl. The responsivity under different applied voltages is provided in Fig. 6(c) and the responsivity illuminated by 450 nm light is extracted in Fig. 6(d). At 10 V, the responsivity at 450 nm is calculated as 3.95 mA/W and it linearly depends on the working voltage. Correspondingly, the detectivities of the device under different voltages are further studied and displayed in Figs. 6(e) and 6(f). Similarly, the voltage-dependent detectivity also follows the linear tendency, and the detectivity at 10 V (@450 nm) is 7×1010 Jones. The linear dependency indicates the carrier extracted efficiency is linearly depending on the external bias voltage. Figure 9 (Appendix A) exhibits the external quantum efficiency (EQE) of the device at different bias voltages. At 10 V, the EQE at 450 nm is as small as 1.1%, indicating the low collection efficiency of the photo-generated carriers. In addition, the dark current of the device is relatively high and the light-dark current ratio is low. As mentioned above, a high defect concentration and numerous grain boundaries are probed on the as-synthesized CsPbClxBr3x crystals. The defects in the crystal normally act as donors, which leads to the high background carrier concentration, and the scattering and nonradiative recombination induced by the defects and grain boundaries significantly reduce the EQE of the photodetector. Therefore, the crystal quality of the CsPbClxBr3x crystal needs to be further improved in order to enhance the photodetection performance. Synthesized parameters such as growth temperature, reagent ratio, and nucleation rate should be carefully considered. During the photoelectronic measurement, the device is exposed in air conditioning with RH of 57%. The photodetector CsPbClxBr3x shows good stability at the air atmosphere (room temperature, RH<57%) and the photoresponse performance can be well maintained for over 6 months. However, once the RH of the environment is >60%, the photocurrent of the device slowly attenuates because of the hydrolysis of the perovskite. Owing to the intrinsic ionic feature, the interaction between perovskite and water is the main reason of the degradation of the perovskite devices [66], which is named as the well known instability in halide perovskites. For example, though a capping layer is introduced for isolating the water, the power conversion efficiency of the perovskite is reduced to 23.5% at 80% RH, which is lower than that at 20%–60% RH [67]. In addition, the oxidation of the surface is another unfavorable factor that can also degrade the photoelectronic performance of the perovskite devices, while the oxidation process typically needs an extremely long time at room temperature in air atmosphere. Therefore, water is the most unstable factor for perovskite devices. Industrial grade packaging with resin or a polymer layer can effectively protect the perovskite active layer and prolong the life time of the perovskite devices.

    (a) The energy band diagram of the InGa/CsPbClxBr3−x/InGa device versus vacuum energy level; (b) the response spectra of CsPbClxBr3−x and CsPbBr3 photodetectors; (c) the wavelength-dependent responsivity of the CsPbClxBr3−x photodetector under different bias voltages; (d) the responsivity at 450 nm as a linear function of the applied voltage; (e) the wavelength-dependent detectivity of the CsPbClxBr3−x photodetector under different bias voltages; (f) the detectivity at 450 nm as a linear function of the applied voltage.

    Figure 6.(a) The energy band diagram of the InGa/CsPbClxBr3x/InGa device versus vacuum energy level; (b) the response spectra of CsPbClxBr3x and CsPbBr3 photodetectors; (c) the wavelength-dependent responsivity of the CsPbClxBr3x photodetector under different bias voltages; (d) the responsivity at 450 nm as a linear function of the applied voltage; (e) the wavelength-dependent detectivity of the CsPbClxBr3x photodetector under different bias voltages; (f) the detectivity at 450 nm as a linear function of the applied voltage.

    4. CONCLUSIONS

    Herein, the highly defect tolerant CsPbClxBr3x crystals are synthesized via a room temperature anti-solvent precipitation process. Spontaneous and amplified spontaneous emissions are both studied under different temperatures and pumping densities. The blue band ASE emissions centered at 456  nm are probed with a relatively low threshold of 106  μJ/cm2 at 50 K. Though the as-prepared CsPbClxBr3x crystal reveals a high defect concentration, ASE is still realized between the competition of radiative recombination and defect-related nonradiative recombination, indicating the highly defect tolerant CsPbClxBr3x as a promising material for optoelectronic devices. Furthermore, an MSM structure CsPbClxBr3x photodetector is prepared and shows a fast response performance with response speeds of 96 μs (rise time) and 34 μs (decay time). Our findings enhance the understanding of the light amplification properties in all inorganic perovskite, as well as the fabrication of a next generation high-performance photodetector.

    Acknowledgment

    Acknowledgment. The XRD, PLQY, and XPS data were obtained using the equipment maintained by Dongguan University of Technology Analytical and Testing Center. The authors thank Dr. F. Yi for the helpful discussion on this manuscript.

    APPENDIX A

    Optical images and the energy band diagram of the CsPbClxBr3x crystal are shown in Figs. 7 and 8, respectively. Figure 9 illustrates the external quantum efficiency of the CsPbClxBr3x photodetector.

    The bright field (a), (c), (e) and dark field (b), (d), (f) optical images of the CsPbClxBr3−x crystal at different magnifications.

    Figure 7.The bright field (a), (c), (e) and dark field (b), (d), (f) optical images of the CsPbClxBr3x crystal at different magnifications.

    The energy band diagram of the CsPbClxBr3−x crystal.

    Figure 8.The energy band diagram of the CsPbClxBr3x crystal.

    The external quantum efficiency (EQE) of the CsPbClxBr3−x photodetector.

    Figure 9.The external quantum efficiency (EQE) of the CsPbClxBr3x photodetector.

    References

    [1] X. B. Wang, X. Y. Zhang, H. Liu. Self-assembling nanoarchitectonics of low dimensional semiconductors for circularly polarized luminescence. J. Materiomics, 9, 683-700(2023).

    [2] Y. L. Lia, H. X. Sun, Z. Lia. Electrospun perovskite nano-network for flexible, near-room temperature, environmentally friendly and ultrastable light regulation. J. Mater. Sci. Technol., 130, 35-43(2022).

    [3] M. V. Kovalenko, L. Protesescu, M. I. Bodnarchuk. Properties and potential optoelectronic applications of lead halide perovskite nanocrystals. Science, 358, 745-750(2017).

    [4] C. Y. Wu, Y. X. Le, L. Y. Liang. Non-ultrawide bandgap CsPbBr3 nanosheet for sensitive deep ultraviolet photodetection. J. Mater. Sci. Technol., 159, 251-257(2023).

    [5] X. X. He, P. Liu, H. H. Zhang. Patterning multicolored microdisk laser arrays of cesium lead halide perovskite. Adv. Mater., 29, 1604510(2017).

    [6] B. Liu, Y. Q. Wang, Y. J. Wu. Novel broad spectral response perovskite solar cells: a review of the current status and advanced strategies for breaking the theoretical limit efficiency. J. Mater. Sci. Technol., 140, 33-57(2023).

    [7] Q. A. Akkerman, V. D’Innocenzo, S. Accornero. Tuning the optical properties of cesium lead halide perovskite nanocrystals by anion exchange reactions. J. Am. Chem. Soc., 137, 10276-10281(2015).

    [8] Y. P. Huang, Z. W. Lai, J. W. Jin. Ultrasensitive temperature sensing based on ligand-free alloyed CsPbClxBr3−x perovskite nanocrystals confined in hollow mesoporous silica with high density of halide vacancies. Small, 17, 2103425(2021).

    [9] J. M. Pina, D. H. Parmar, G. Bappi. Deep-blue perovskite single-mode lasing through efficient vapor-assisted chlorination. Adv. Mater., 33, 2006697(2021).

    [10] L. X. Su. Room temperature amplified spontaneous emissions in a sub-centimeter sized CsPbBr3 bulk single crystal. Opt. Express, 31, 39020-39029(2023).

    [11] S. M. H. Qaid, H. M. Ghaithan, B. A. Al-Asbahi. Tuning of amplified spontaneous emission wavelength for green and blue light emission through the tunable composition of CsPb(Br1−xClx)3 inorganic perovskite quantum dots. J. Phys. Chem. C, 125, 9441-9452(2021).

    [12] H. H. Zhang, J. N. Yao, H. B. Fu. Patterning rainbow like amplified spontaneous emission arrays for full-color CsPbX3 quantum dot film displays. Chem. Mater., 32, 9602-9608(2020).

    [13] N. Liu, H. Y. Luo, X. Y. Wei. Linearly manipulating color emission via anion exchange technology for high performance amplified spontaneous emission of perovskites. Adv. Mater., 36, 2308672(2024).

    [14] D. Kim, H. Ryu, S. Y. Lim. On the origin of room-temperature amplified spontaneous emission in CsPbBr3 single crystals. Chem. Mater., 33, 7185-7193(2021).

    [15] C. Y. Zhao, W. M. Tian, J. X. Liu. Stable two-photon pumped amplified spontaneous emission from millimeter-sized CsPbBr3 single crystals. J. Phys. Chem. Lett., 10, 2357-2362(2019).

    [16] L. X. Su, T. F. Li, Y. Zhu. A vertical CsPbBr3/ZnO heterojunction for photo-sensing lights from UV to green band. Opt. Express, 30, 23330-23340(2022).

    [17] C. Fu, Z.-Y. Li, J. Wang. A simple-structured perovskite wavelength sensor for full-color imaging. Nano Lett., 23, 533-540(2023).

    [18] Y. Li, Z. F. Shi, L. Z. Lei. Highly stable perovskite photodetector based on vapor-processed micrometer-scale CsPbBr3 microplatelets. Chem. Mater., 30, 6744-6755(2018).

    [19] J. Z. Song, Q. Z. Cui, J. H. Li. Ultralarge all-inorganic perovskite bulk single crystal for high-performance visible-infrared dual-modal photodetectors. Adv. Opt. Mater., 5, 1700157(2017).

    [20] Y. Zhang, S. Y. Li, W. Yang. Millimeter-sized single-crystal CsPbBr3/CuI heterojunction for high-performance self-powered photodetector. J. Phys. Chem. Lett., 10, 2400-2407(2019).

    [21] F. M. Guo, J. Wang, Y. Li. Postripening fabrication and self-driven narrowband photoresponse of large-grain, phase-pure CsPbBr3 films. Sol. RRL, 6, 2200828(2022).

    [22] H. R. Sun, L. X. Su, Q. Zeng. Kilogram-scale high-yield production of PbI2 microcrystals for optimized photodetectors. J. Mater. Chem. C, 12, 6433-6442(2024).

    [23] S. Mahato, M. T. Szwajkowska, S. Singh. Surface-engineered methylammonium lead bromide single crystals: a platform for fluorescent security tags and photodetector applications. Adv. Opt. Mater., 12, 2302257(2024).

    [24] M. L. Liao, B. B. Shan, M. Li. In situ Raman spectroscopic studies of thermal stability of all-inorganic cesium lead halide (CsPbX3, X = Cl, Br, I) perovskite nanocrystals. J. Phys. Chem. Lett., 10, 1217-1225(2019).

    [25] D. M. Calistru, L. Mihut, S. Lefrant. Identification of the symmetry of phonon modes in CsPbCl3 in phase IV by Raman and resonance-Raman scattering. J. Appl. Phys., 82, 5391-5395(1997).

    [26] L. X. Su. Room temperature growth of CsPbBr3 single crystal for asymmetric MSM structure photodetector. J. Mater. Sci. Technol., 187, 113-122(2024).

    [27] L. X. Su. Growth of a sub-centimeter-sized CsPbBr3 bulk single crystal using an anti-solvent precipitation method. Symmetry, 16, 332(2024).

    [28] L. X. Su, Y. Zhang, J. Xie. All-inorganic CsPbBr3/GaN hetero-structure for near UV to green band photodetector. J. Mater. Chem. C, 10, 1349-1356(2022).

    [29] M. Baranowski, P. Plochocka, R. Su. Exciton binding energy and effective mass of CsPbCl3: a magneto-optical study. Photon. Res., 8, A50-A55(2020).

    [30] F. Xu, H. M. Wei, Y. Q. Wu. Nonmonotonic temperature-dependent bandgap change of CsPbCl3 films induced by optical phonon scattering. J. Lumin., 257, 119736(2023).

    [31] J. B. Gong, H. X. Zhong, C. Gao. Pressure-induced indirect-direct bandgap transition of CsPbBr3 single crystal and its effect on photoluminescence quantum yield. Adv. Sci., 9, 2201554(2022).

    [32] B. Wu, H. F. Yuan, Q. Xu. Indirect tail states formation by thermal-induced polar fluctuations in halide perovskites. Nat. Commun., 10, 484(2019).

    [33] S. Jin, Y. L. Zheng, A. Z. Li. Characterization of photoluminescence intensity and efficiency of free excitons in semiconductor quantum well structures. J. Appl. Phys., 82, 3870-3873(1997).

    [34] D. E. Cooper, J. Bajaj, P. R. Newmann. Photoluminescence spectroscopy of excitons for evaluation of high-quality CdTe crystals. J. Cryst. Growth, 86, 544-551(1988).

    [35] Z. C. Feng, A. Mascarenhas, W. J. Choyke. Low temperature photoluminescence spectra of (001) CdTe films grown by molecular beam epitaxy at different substrate temperatures. J. Lumin., 35, 329-341(1986).

    [36] L. Bergman, X. B. Chen, J. L. Morrison. Photoluminescence dynamics in ensembles of wide-band-gap nanocrystallites and powders. J. Appl. Phys., 96, 675-682(2004).

    [37] H. P. He, H. P. Tang, Z. Z. Ye. Temperature-dependent photoluminescence of quasialigned Al-doped ZnO nanorods. Appl. Phys. Lett., 90, 023104(2007).

    [38] C. X. Shan, Z. Liu, S. K. Hark. Temperature dependent photoluminescence study on phosphorus doped ZnO nanowires. Appl. Phys. Lett., 92, 073103(2008).

    [39] S. H. Yu, J. Xu, X. Y. Shang. Unusual temperature dependence of bandgap in 2D inorganic lead-halide perovskite nanoplatelets. Adv. Sci., 8, 2100084(2021).

    [40] K. Mukhuti, A. Mandal, B. Roy. Carrier thermalization and zero-point bandgap renormalization in halide perovskites from the Urbach tails of the emission spectrum. Appl. Phys. Lett., 121, 182104(2022).

    [41] S. Bose, S. Shendre, Z. G. Song. Temperature-dependent optoelectronic properties of quasi-2D colloidal cadmium selenide nanoplatelets. Nanoscale, 9, 6595-6605(2017).

    [42] R. Chen, Q.-L. Ye, T. C. He. Uniaxial tensile strain and exciton–phonon coupling in bent ZnO nanowires. Appl. Phys. Lett., 98, 241916(2011).

    [43] A. J. Fischer, W. Shan, J. J. Song. Temperature-dependent absorption measurements of excitons in GaN epilayers. Appl. Phys. Lett., 71, 1981-1983(1997).

    [44] B. T. Diroll, G. Nedelcu, M. V. Kovalenko. High-temperature photoluminescence of CsPbX3 (X = Cl, Br, I) nanocrystals. Adv. Funct. Mater., 27, 1606750(2017).

    [45] J. Li, X. Yuan, P. Jing. Temperature-dependent photoluminescence of inorganic perovskite nanocrystal films. RSC Adv., 6, 78311-78316(2016).

    [46] M. I. Dar, G. Jacopin, S. Meloni. Origin of unusual bandgap shift and dual emission in organic-inorganic lead halide perovskites. Sci. Adv., 2, e1601156(2016).

    [47] Y. P. Varshni. Temperature dependence of the energy gap in semiconductors. Physica, 34, 149-154(1967).

    [48] J. Bardeen, W. Shockley. Deformation potentials and mobilities in non-polar crystals. Phys. Rev., 80, 72-80(1950).

    [49] H. Y. Fan. Temperature dependence of the energy gap in monatomic semiconductors. Phys. Rev., 78, 808-809(1950).

    [50] B. Liu, R. Chen, X. L. Xu. Exciton-related photoluminescence and lasing in CdS nanobelts. J. Phys. Chem. C, 115, 12826(2011).

    [51] Y. Q. Shi, R. X. Li, G. X. Yin. Laser-induced secondary crystallization of CsPbBr3 perovskite film for robust and low threshold amplified spontaneous emission. Adv. Funct. Mater., 32, 2207206(2022).

    [52] J. Kang, L. W. Wang. High defect tolerance in lead halide perovskite CsPbBr3. J. Phys. Chem. Lett., 8, 489-493(2017).

    [53] D. P. Nenon, K. Pressler, J. Kang. Design principles for trap-free CsPbX3 nanocrystals: enumerating and eliminating surface halide vacancies with softer Lewis bases. J. Am. Chem. Soc., 140, 17760-17772(2018).

    [54] G. H. Ahmed, J. K. El-Demellawi, J. Yin. Giant photoluminescence enhancement in CsPbCl3 perovskite nanocrystals by simultaneous dual-surface passivation. ACS Energy Lett., 3, 2301-2307(2018).

    [55] D. Kim, H. Ryu, S. Y. Lim. On the origin of room-temperature amplified spontaneous emission in CsPbBr3 single crystals. Chem. Mater., 33, 7185-7193(2021).

    [56] M. B. Price, J. Butkus, T. C. Jellicoe. Hot-carrier cooling and photoinduced refractive index changes in organic-inorganic lead halide perovskites. Nat. Commun., 6, 8420(2015).

    [57] S. Cheng, Q. Chang, Z. Wang. Observation of net stimulated emission in CsPbBr3 thin films prepared by pulsed laser deposition. Adv. Opt. Mater., 9, 2100564(2021).

    [58] Y. Wang, M. Zhi, Y.-Q. Chang. Stable, ultralow threshold amplified spontaneous emission from CsPbBr3 nanoparticles exhibiting trion gain. Nano Lett., 18, 4976-4984(2018).

    [59] A. Balena, A. Perulli, M. Fernandez. Temperature dependence of the amplified spontaneous emission from CsPbBr3 nanocrystal thin films. J. Phys. Chem. C, 122, 5813-5819(2018).

    [60] H. L. Zhang, L. Yuan, Y. Chen. Amplified spontaneous emission and random lasing using CsPbBr3 quantum dot glass through controlling crystallization. Chem. Commun., 56, 2853(2020).

    [61] H. J. He, E. Ma, X. Y. Chen. Single crystal perovskite microplate for high-order multiphoton excitation. Small Methods, 3, 1900396(2019).

    [62] W. Yang, Y. Zhang, Y. J. Zhang. Transparent Schottky photodiode based on AgNi NWs/SrTiO3 contact with an ultrafast photoresponse to short-wavelength blue light and UV-shielding effect. Adv. Funct. Mater., 29, 1905923(2019).

    [63] L. X. Su, Y. Q. Zuo, J. Xie. Scalable manufacture of vertical p-GaN/n-SnO2 heterostructure for self-powered ultraviolet photodetector, solar cell and dual-color light emitting diode. InfoMat, 3, 598(2021).

    [64] K. W. Liu, J. G. Ma, J. Y. Zhang. Ultraviolet photoconductive detector with high visible rejection and fast photoresponse based on ZnO thin film. Solid State Electron., 51, 757-761(2007).

    [65] H. J. Zhang, X. Liu, J. P. Dong. Centimeter-sized inorganic lead halide perovskite CsPbBr3 crystals grown by an improved solution method. Cryst. Growth Des., 17, 6426-6431(2017).

    [66] S. J. Cheng, H. Z. Zhong. What happens when halide perovskites meet with water?. J. Phys. Chem. Lett., 13, 2281-2290(2022).

    [67] Y. Zou, W. J. Yu, H. Q. Guo. A crystal capping layer for formation of black-phase FAPbI3 perovskite in humid air. Science, 385, 161-167(2024).

    Longxing Su, Bingheng Meng, Heng Li, Zhuo Yu, Yuan Zhu, Rui Chen, "Amplified spontaneous emission and photoresponse characteristics in highly defect tolerant CsPbClxBr3−x crystal," Photonics Res. 13, 286 (2025)
    Download Citation