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

Amplified spontaneous emission properties of solution processed CsPbBr3 perovskite thin films doped with large-group ammonium cations

Open Access Open Access

Abstract

Inorganic solution processed CsPbBr3 perovskite thin films are highly promising optoelectronic materials and have been applied in solar cells, light emitting diodes, and photodetectors. They show good electroluminescence properties combined with high thermal stability. In this work, CsPbBr3 perovskite thin films doped with a series of organic cations of various chain lengths, namely, methylammonium (MA), ethylammonium (EA), butylammonium (BA) and octylammonium (OA) at optimized molar ratios of (organic cations: Cs+), were fabricated with a two-step solution process. The crystalline structure, surface morphology and photophysical properties of the films were characterized in detail. The surface morphology of the films was improved with reduced surface roughness, accompanying by doping with the organic cations. The amplified spontaneous emissions (ASE) were observed from all of the different CsPbBr3 films at room temperature. The gain coefficients for the doped CsPbBr3 films were higher than the pristine films, due to much reduced bulk defects, reduced non-radiative recombination in the films, as well as improved surface quality of the films. This work provides a simple method of preparing organic cations modified CsPbBr3 laser gain media of high performance.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

In recent years, solution-processed perovskite semiconductors in the form of thin films, nanocrystals and single crystals have attracted great attention for their potential applications in solar cells, light-emitting diodes (LEDs), photodetectors and lasers [1], because of their excellent optical and electrical properties, such as long carrier lifetimes, micrometer-scaled diffusion lengths, high charge mobilities and large optical absorption [2]. In particular, organic-inorganic hybrid perovskites (CH3NH3PbX3, X = Cl, Br, and I) have enabled solution-processed solar cells with power conversion efficiency of 23.7% [3]. Meanwhile, perovskites have shown high photoluminescence quantum yield (PLQY) and large optical gain [4]. A red hybrid perovskite LED with an external quantum efficiency (EQE) of 20.7% has been demonstrated by Wang et al. [5]. Moreover, wavelength-tunable amplified spontaneous emission (ASE) in hybrid organic-inorganic CH3NH3PbX3 films at low thresholds of photoexcitation was demonstrated [6]. Even more, low threshold single-crystal CH3NH3PbX3 nanowires lasers at room temperature (RT) were obtained [7], and also two-photon pumped CH3NH3PbBr3 microwire lasers [8]. However, hybrid organic-inorganic perovskites present poor thermal stability due to the volatile organic constituents [9]. Inorganic perovskites CsPbX3 (X = Cl, Br, and I) have been studied intensively for their applications in solar cells, photodetectors, LEDs and lasers, due to their enhanced stability and excellent optoelectronic properties compared to their hybrid organic-inorganic counterparts. They have also drawn great attention as laser gain medium. Benefiting from various processing methods, such as solution processing [10], chemical vapor deposition (CVD) [11] and atomic layer deposition (ALD) [12], many inorganic micro/nano perovskite laser devices have been developed, such as perovskites quantum dot lasers [13], wavelength tunable CsPbX3 (X = Cl, Br, and I) nanowire lasers [14], microdisk lasers fabricated using solution self-assembly method [1518] and a single-mode laser in a cesium lead halide perovskite submicron sphere [19]. Thin film perovskite laser array with two dimension (2D) photonic crystal resonator [20] and random CsPbBr3 laser have also been demonstrated [21]. Compared with their inorganic-organic hybrid counterpart CH3NH3PbX3, all-inorganic perovskites present much stronger environment stability and photo-thermal stability (thermal decomposition temperature over 500 K in air) [22]. They have also very high physical and chemical stability. Under 1.2 times threshold pumping excitation, the integrated emission intensity of CsPbBr3 nanowire lasers maintained the same after over 4.4×108 excitation cycles under atmosphere condition [23]. Besides CsPbX3 perovskite nanomaterials, polycrystalline perovskite CsPbX3 thin films used for ASE or lasing have also been reported [2426]. Compared with CsPbX3 nanomaterials, the synthesis of CsPbX3 thin films using one or two-step method is much simpler and the organic capping ligands are not needed to stabilize the nanocrystals during the preparation process. However, the performance of the ASE or lasing in CsPbX3 thin films is not as efficient as CsPbX3 nanocrystal materials since the internal quantum efficiency of CsPbX3 thin films (∼7%) is much lower than that of CsPbX3 nanocrystals (∼90%) [21] . The conventional one-step method only needs direct spin-coating of a CsPbX3 precursor solution, it is usually difficult to control perovskite crystallization and film quality due to low solubility of the cesium bromide (CsBr) precursor [22], causing voids and pinholes in the films. Polymer additive poly(ethylene oxide) (PEO) has been introduced into CsPbX3 thin films to improve the quality of the films, i.e., the film morphology and the PLQY, thus to enhance EQE of the CsPbX3 LEDs and the ASE or lasing performance of the films [27]. Inorganic solids ZnO nanoparticles (NPs) were also introduced into the CsPbX3 precursor solution to enhance the crystallization and to improve the photoluminescence (PL) intensity of the final CsPbBr3:ZnO films [21], the introduction of ZnO NPs provided an effective route for CsPbBr3 nucleation during film forming process and led to compact and smooth films with no obviously large voids and pinholes. ASE was also investigated in CsPbBr3 thin films at various temperatures and trap densities [26]. Better photophysical properties, such as PLQY, photoluminescence relaxation dynamics, and electroluminescence properties have been observed in films deposited from CsBr-rich solution, which was ascribed to a reduction of halogen vacancies [26]. Recently, You et al. incorporated a small amount of organic cation (MA) into the CsPbBr3 lattice and prepared a mixed-cation Cs0.87MA0.13PbBr3 film as the emitting layer in a green LED, the optimized perovskite composition modulation together with electron-injection interface passivation with a polymer polyvinyl pyrrolidine (PVP) minimized pinholes in the perovskite film and reduced the device leakage current, thus resulted in obtaining a high brightness of 91,000 cdm−2 and a high external quantum efficiency (EQE) of 10.4% [28]. The improved device performance was attributed to improved film morphology, reduced non-radiative recombination as well as improved charge injection balance. By far, all of the reported inorganic perovskite film waveguides were fabricated by one step method. Two step method has never been employed to fabricate inorganic perovskite ASE waveguides, especially at room temperature and atmospheric conditions. Compared with one-step method, films of better flatness and compactness can be more easily fabricated with the two-step method. The solvent volatilization and perovskite crystal crystallization occur in the same process for the one-step method, which is adverse to the formation of compact and flat films due to film shrinkage. But for the two-step method, the final films quality is mainly determined by the PbX2 films fabricated during the first step [29].

In this work, inorganic CsPbBr3 polycrystalline films were facilely prepared via a two-step solution synthesis method. During the synthesis process, we incorporated a series of organic cations of various chain lengths, i.e., methylammonium (MA), ethylammonium (EA), butylammonium (BA) and octylammonium (OA) into CsPbBr3 films to improve film quality [28,30]. By introducing ammonium halides of different chain lengths into the CsBr precursor solution, we demonstrate an effective route for CsPbBr3 nucleation during the two-step synthesis process to better control the film morphology and reduce the occurrence of pinholes or voids. The crystalline structural properties and photophysical properties, as well as the ASE properties of the resulted films were investigated in detail. X-ray diffraction (XRD) confirmed the crystalline phase of the perovskite nanocrystals in the films. Multiple spectroscopy measurements, including UV-visible optical absorption, steady-/transient-state fluorescence and power-dependent PL, indicate that notable green emission originates from excitonic radiative recombination of the CsPbBr3 nanocrystals. Owing to the improved film compactness and smoothness, the CsPbBr3 films with organic cations show enhanced PL and ASE efficiency. The ASE thresholds decreased from 404 µJ/cm2 for the pristine CsPbBr3 films, to 271 µJ/cm2, 300 µJ/cm2, 338 µJ/cm2 and 399 µJ/cm2 for MA, EA, BA and OA doped CsPbBr3 films, respectively. We expect these results can provide an effective route to the synthesis of high-quality perovskite films and promote their applications in waveguide lasers.

2. Experimental section

Lead bromide (PbBr2, 99.99%) was purchased from Sigma. Cesium bromide (CsBr, 99.99%) was purchased from Alfa Aesar. Methylammonium Bromide (MABr, CH3NH3Br) (>99.5%), Ethylammonium Bromide (EABr, CH3CH2NH3Br) (>99.5%), Butylammonium Bromide (BABr, CH3(CH2)3NH3Br)(>99.5%), Octylammonium Bromide (OABr, CH3(CH2)7NH3Br) (>99.5%) were purchased from Xi’an Polymer Light Technology Corp. Dimethyl formamide (DMF, >99.9%) and methanol were purchased from Sigma Aldrich. All solvents were used without further purification.

The CsPbBr3 thin films were fabricated by two step solution method since it is a facile and low-cost approach for large-scale production of perovskite materials (Fig. 1). Briefly speaking, 300 mg PbBr2 dissolved in 1 mL DMF solvent and 10 mg/mL CsBr in methanol were prepared as precursor solutions [31]. For films incorporated with the organic cations, MABr, EABr, BABr and OABr were introduced into CsBr solution at 1:20 molar ratio of the organic cations and Cs+ (MA, EA, BA and OA: Cs+).

 figure: Fig. 1.

Fig. 1. Schematic diagram of the fabrication process of inorganic perovskite films.

Download Full Size | PDF

The PbBr2 solution was firstly stirred at 60 °C for 5 h, then filtered by hydrophobic filter head (PTFE) (0.45 µm, Titan) for using. Quartz substrates were cleaned by sonication in detergent, deionized water, acetone, and isopropanol for 20 min successively. Subsequently, substrates were dried by nitrogen flow and cleaned with ultraviolet ozone plasma for 15 min and heated at 75 °C before use. The PbBr2 solution was heated up and kept at 75 °C, then spin-coated on the quartz substrates at 3000 rpm for 30 s, followed by annealing on a hot plate at 75 °C for 30 min. Then the PbBr2 film was dipped and immersed into the CsBr solution at 50 °C for 5 min to induce perovskite crystallization and enable the growth of the pristine CsPbBr3 films. Finally, the films were air annealed at 150 °C for 10 min to remove the organic residues and to obtain the final perovskites granular films. The whole process of film fabrication was carried out under ambient environment at room temperature.

The crystalline structure of the CsPbBr3 perovskite films were characterized by XRD with CuKa radiation (Smartlab, Rigaku). SEM images were taken by a scanning electron microscope (JSM-7800F, JEOL, Japan). Absorption spectra were recorded by a UV-vis-NIR spectrophotometer (SHIMADZU UV-1750, Shimadzu, Japan), PL spectra of the perovskite films were obtained by a fluorescence spectrophotometer (F-4600, Hitachi, Japan), and time-resolved fluorescence spectra were achieved using a fluorescence lifetime spectrofluorometer (Edinburgh Instruments Ltd., FLS980, United Kingdom). The surface roughness of the films was measured using atomic force microscope (XE-70, Park). The PLQY of the perovskite films were obtained by a home-made PLQY measurement system.

A nanosecond Nd:YAG laser system in combination with an optical parametric oscillator (OPO) system (Surelite-II-10 Continuum) was used as the pumping source for the ASE excitation. The output wavelength of 355 nm of the Nd:YAG laser (repetition rate: 10 Hz, pulse width: 5 ns) was used to pump the OPO to generate a lasing output of 400 nm (repetition rate: 10 Hz, pulse width: 5 ns), which was employed to excite ASE of the CsPbBr3 films. The laser beam passed successively through diaphragm (4 mm in diameter), slit (4 mm in width), neutral density filter and was then focused by a cylindrical lens (focal length: 10 cm) into a narrow strip (4.7 mm ×0.27 mm), which irradiated onto the sample surface. The light signals emitted from the samples were collected by an optical fiber and detected by a spectrometer in conjunction with a CCD detector (SR-500i, Andor) with a spectral resolution of 0.1 nm. All the measurements were performed in atmosphere air and at room temperature (293 K).

3. Results and discussion

CsPbBr3 perovskite thin films were fabricated by spin-coating a PbBr2 precursor in DMF onto quartz substrates, followed by annealing at 75 °C for 30 min to remove the solvent. Then the coated PbBr2 layers were dipped into CsBr precursor in methanol. The soaking time was firstly optimized in order to achieve pure phase CsPbBr3 films. Figure 8 shows the images of the obtained films prepared from various soaking time. XRD was also utilized to monitor the crystalline structure of the films during their growth. The films prepared from 5 min soaking time presented the best film quality, as verified by the XRD patterns Fig. 9). The CsPbBr3 perovskite films deposited onto quartz substrates exhibited a high density of pinholes (Fig. 2(a)). Then we added a small amount of organic additives, i.e., CH3NH3Br (MABr), CH3CH2NH3Br (EABr), CH3(CH2)3NH3Br (BABr), CH3(CH2)7NH3Br (OABr), into the CsBr precursor solution to control the crystallization kinetics of the CsPbBr3 films through molecular pinning [32,33], thus to reduce the pinholes and to improve the film morphology. The molar ratios of the organic cations and Cs+ were optimized at 1:20. Figures 2(b)–(e) show the SEM images of the pristine CsPbBr3 films and CsPbBr3 films doped with organic cations. The morphology of the CsPbBr3 films was improved considerably once we added the additives. They presented a negligible density of pinholes and smooth and compact surface. The gullies between crystalline grain boundaries were also largely reduced. The CsPbBr3 films prepared with ammonium halides presented relatively smaller crystalline grains. The size statistics of the corresponding films are given in the inserts of Fig. 2(b)-(e). We noticed that organic cations with longer chains were beneficial to achieve smaller crystal grain size, the CsPbBr3 crystalline size reduced from 528 nm for the films without additives to 447 nm, 387 nm, 334 nm and 268 nm for MA, EA, BA and OA doped films, respectively. We ascribed the size reduction of grains to the introduction of organic cations.

 figure: Fig. 2.

Fig. 2. SEM images of different organic cations doped perovskite films ((a). CsPbBr3, (b). CsPbBr3:MA, (c). CsPbBr3:EA, (d). CsPbBr3:BA, (e). CsPbBr3:OA).

Download Full Size | PDF

Surface roughness affects the optical characteristics of the films significantly. The surface roughness of the films was measured with AFM. From the AFM images shown in Fig. 10, we can see the surface of the CsPbBr3 films doped with additives were improved compared to the pristine CsPbBr3 films. The roughness of the CsPbBr3 films (scanning area 5 µm × 5 µm) was 38.46 nm. while the films doped with organic cations were 28.02 nm, 28.83 nm, 29.94 nm, 30.51 nm for films doped with MA, EA, BA and OA, respectively. The pinholes were filled, and the films surface was smoothed out. The smoother surface will decrease the scattering, reflection and deflection loss of incident pumping light at air-film interface and enhance the proportion of the pump light propagating in the films, which would be beneficial for the generation of ASE in the CsPbBr3 films with organic cations. The non-radiative transition loss was also reduced at grain boundaries due to the filling effect of the organic cations.

We characterized the crystal structure of the CsPbBr3 films and films fabricated with organic additives. XRD measurement revealed the effect of doped organic cations on the crystallization properties of the films. Figure 9 shows the X-ray diffraction patterns of the pristine films prepared at different soaking time, which were almost identical, with all of the main crystallographic signatures matching that of the pure CsPbBr3 phase. Figure 3 shows the X-ray diffraction patterns of the doped films. The diffraction peaks at 21.69° and 34.62° are from (110) and (210) planes of the CsPbBr3 crystal in Fig. 3. Besides of the pure CsPbBr3 phase, the diffraction peak located at 11.6° can be assigned to the resident tetragonal crystal of CsPb2Br5 in the films, the diffraction peaks at 12.66°, 12.94° and 22.48° belong to (012) (110) and (300) planes of Cs4PbBr6 crystal [30,31]. The generation of metastable CsPb2Br5 phase has also been observed by other groups [27,31,34]. It is attributed to the nonstoichiometric material transfer and structural rearrangement during the solution synthesis and thermal annealing process [35,36]. The transformation from CsPb2Br5 to CsPbBr3 happened when the reaction continued and there always exists an impure phase CsPb2Br5 in the final films, this secondary phase can reduce the trap state density of the CsPbBr3 and benefit to the light emission of the CsPbBr3 nanocrystals [37]. 5 min soaking time was found to be the optimized reaction time. The ASE mainly originates from CsPbBr3 rather than Cs4PbBr6 [38]. We found that the relative intensities of the peaks decreased with increasing chain length, which meant long chain organic amine ions were in favour of the formation of perovskite crystals. In addition, all of the XRD diffraction peaks were quite sharp and narrow, indicating the great crystallinity of the perovskite films. The intensity of the peaks became stronger, which implied the ratio of orthorhombic CsPbBr3 crystals increased as prolonger reaction time. We found that the XRD patterns from the CsPbBr3 films doped with different organic cations were almost identical, as the XRD diffraction peak positions did not change as the variation of the chain length of the additives. We considered that most of the organic cations did not incorporate into the perovskite lattice structure in the two-step solution process, they only cladded on top of the surface of perovskite crystalline grains similar as surface-active agent [39], thus the introduction of the organic cations did not change the intrinsic perovskite crystalline structure. On the other hand, from the SEM images, we did not find any quasi two-dimension perovskite crystalline structure as expected, when introducing longer chain cations. In addition, longer chain ammonium with larger steric hindrance prevented ions (Cs+, Pb2+) from continuously entering into the forming perovskite crystals and prevented crystal growth, so the films doped longer chain length had smaller grain size. The intensity of the XRD diffraction peaks from films doped with additives enhanced, indicating better crystallization in the additives doped films. The diffraction peak intensity of films doped with additives with shorter chain length was stronger than those of films doped with additives with longer chain length, which indicated that organic cations with shorter chain length were more beneficial to the crystallization of the CsPbBr3 nanocrystals.

 figure: Fig. 3.

Fig. 3. XRD patterns of the additives doped perovskite films.

Download Full Size | PDF

Combined with the SEM and XRD results, all of the characteristics imply that high-quality perovskite films can be obtained by introducing a certain amount of ammonium halides into the CsBr precursor. The addition of ammonium halides affects the crystal nucleation on the substrate, hindering the crystal grains to grow up. The CsPbBr3 films with organic cations have less pinholes, and the crystal grains of the CsPbBr3 films with additives became smaller than that of the pristine films. The organic chains of the ammonium halides play a similar role to the organic ligands on nanocrystals surface (or quantum dots) during the CsPbBr3 crystal growth [40].

The absorption, photoluminescence (PL) spectra and decay lifetime of the CsPbBr3 films and CsPbBr3 films doped with additives are shown in Fig. 4. The absorption and fluorescence spectra showed that the doped organic cations had no effect on the band gap structure of the CsPbBr3 films. The absorption spectra of the CsPbBr3 films with additives showed almost no change in comparison with the pristine films, with the absorption peak at around 518 nm, indicating that the addition of organic cations into CsPbBr3 precursor solution does not affect the original absorption property of CsPbBr3 films. We then carried out steady-state PL on CsPbBr3 films prepared with different additives, negligible shifts of the emission peaks’ positions were observed from the doped CsPbBr3 films relative to the pristine films. The emission peak was around 535 nm. We also observed the full width half maximum (FWHM) of the PL was around 16 nm for MA, EA, BA, and OA doped films. The absorption and PL characterization results are summarized in Table 1. In our experiments, we also studied the absorption and PL spectra of the CsPbBr3 films with organic cations prepared from the CsBr precursors doped with different amounts of ammonium halides, the molar ratios of CsBr: MABr/EABr/BABr/OABr were set at 40:1, 20:1, 10:1, 5:1, 2.5:1, 1.7:1 and 1:1 (Fig. 11). We found that the absorption spectra of them were almost identical, which suggests that during the two-step synthesis process of the CsPbBr3 films, the organic cations did not incorporate into the CsPbBr3 crystal lattice and no Ruddlesden-Popper layered perovskite structures were generated in the final films [28,40].

 figure: Fig. 4.

Fig. 4. Absorption spectra (a) and normalized PL spectra (b) of the perovskite films doped with different organic cations (Excitation wavelength: 400 nm).

Download Full Size | PDF

Tables Icon

Table 1. Absorption spectra and fluorescence spectra of the CsPbBr3 films doped with organic cations

We next acquired time-resolved PL decay spectra of the different perovskite films. The time-resolved PL curves were fit to bi-exponential decays, where the fast decay component is associated with trap-assisted recombination at grain boundaries or surfaces, and the slow decay is ascribed to radiative recombination inside the bulk perovskite phase [41,42]. Figure 5 showed the decay curves for the different perovskite films. The decay times for the pristine films are τ1=1.15 ns, τ2=3.16 ns. For the films doped with MA, EA, BA and OA, the decay times are (τ1=1.51 ns, τ2=3.99 ns), (τ1=1.93 ns, τ2=4.66 ns), (τ1=1.56 ns, τ2=4.36 ns), (τ1=2.36 ns, τ2=6.43 ns), respectively. Generally, it was found that the PL lifetime of the perovskite films increased after the addition of the organic cations. The longer the chain length of the organic cations, the longer the PL lifetime. We hypothesize that this might be a result of reduced bulk defects. In addition, stronger PL intensity was achieved from the films with organic cations. The PLQY of the CsPbBr3 films were measured to be ∼1.9% at a light intensity of 3.1 mW/cm2. Together with the improvement of film morphology, it can be proved that the ASE properties of the CsPbBr3 films with additives could be improved compared to the pristine films.

 figure: Fig. 5.

Fig. 5. Time-resolved fluorescence spectra of the perovskite films with organic cations.

Download Full Size | PDF

To evaluate the ASE properties of the CsPbBr3 films, laser excitation experiments were conducted. The laser output pulses with a pulse width of 5 ns, central wavelength of 400 nm and repetition rate of 10 Hz from an optical parametric oscillator (OPO) pumped by a Nd:YAG laser at 355 nm were focused onto the films by a cylindrical lens to form a narrow strip (4.7 mm×0.27 mm) pumping zone. Measured emission spectra of the films are given in Fig. 6. Figure 6(a) shows the power-dependent emission spectra of the films under excitation. For the CsPbBr3 films with a thickness of 210 nm, a broad spontaneous emission (SE) spectrum centred at about 533 nm with a FWHM of about 22 nm can be observed under relatively low pump excitation (< 404 µJ/cm2). When the pump intensity increased, a relatively sharp and narrow peak at about 539 nm appeared and then became the dominant spectral component of the emission spectrum. Meanwhile, the FWHM of the emission spectra gradually reduced to about 4 nm, indicating that the SE transformed into ASE. Similarly, the emission of the CsPbBr3 films doped with organic cations also changed from SE into ASE with the increasing intensity of the pump laser. The ASE emission intensity of the CsPbBr3 films doped with organic cations was stronger than that of pristine films as shown in the Fig. 6, which agrees well with the results of PL emission. The ASE thresholds for the CsPbBr3 films doped with organic cations were lower than those for the pristine films, which were 271 µJ/cm2, 300 µJ/cm2, 338 µJ/cm2 and 399 µJ/cm2, respectively. The MA doped perovskite films had the lowest ASE threshold. There are several aspects contributing to the ASE enhancement. Firstly, by introducing the organic cations, the crystal grains became smaller and the larger pinholes or voids have been effectively reduced, therefore the waveguide loss has been reduced. Secondly, the CsPbBr3 films with organic cations showed more compact and smoother surface, which makes the light in the films have higher chance to achieve total reflection and reflect back to the films, thus the escaping light at air/film interface could be decreased. As a result, the light is expected to undergo oscillation with much more gain before it outputs as ASE from the edge of the films, which can decrease the ASE threshold and enhance the ASE intensity. We also measured the gain and loss coefficients for the CsPbBr3 films using variable stripe lengths (VSL) method [43]. Figure 12 shows the stripe-length dependence of PL detected at peak wavelengths of 539 nm of the ASE spectra. The stripe width was ∼2 mm, excitation intensity was about 1200 µJ/cm2, which was nearly 3 folds above the ASE thresholds. The gain and loss coefficients were deduced from the classical VSL equation I = I0/g·{exp(g·L)-α}. The measurement results are summarized in Fig. 12 and Table 2. The gain coefficients were much higher for the CsPbBr3 films with organic cations than that for the pristine films, while the losses were lower. We found that the CsPbBr3 films with additives of longer chain lengths achieved lower gain coefficients and lower optical loss. The net gain for the CsPbBr3 films doped with MA was the highest, which was complying with the ASE measurement results. The gain and loss of the doped perovskite films showed different trends as the variation of the chain length of the organic cations. From the SEM and AFM measurement results, we found that the surface roughness of the doped films increased with the chain length. The introduction of organic amine cations could reduce the perovskite grain size, which can then much better fill the holes and voids in the films, thus resulting in lower intrinsic optical loss of the film waveguides. However, smaller grains resulted in the unevenness of the film surface (as seen in Fig. 2). Increased surface roughness of the films led to larger scattering optical loss and interface reflection loss of the incident pump beam, which reduced the optical gain of the film waveguides in the end, thus the CsPbBr3 films doped with OA achieved the lowest gain. From the lifetime measurement, we found that introducing organic cations with long chain length would significantly reduce the defects in the bulk, which is beneficial to reducing the nonradiative transition loss due to the energy transfer to defects. But the interface scattering loss increased with introducing the cations with longer chain length, the contribution from the scattering loss to the loss was less than that from the defects’ reduction. Thus, the reduction of the nonradiative transition to defects cancels out the scattering loss. Thus the OA doped films had the smallest loss. However, MA and EA doped films showed better film uniformity and flatness, thus lower scattering loss, which results in their larger net gain. In addition, from the XRD measurements, we found that with the increasing of the organic chain length, the diffraction peak intensity of Cs4PbBr6 (300) increased, especially for BA doped films, which means the increasing of Cs4PbBr6 phase. According to the previous work, Cs4PbBr6 phase has little contribution to ASE, which may be another cause of reducing the net gain.

 figure: Fig. 6.

Fig. 6. ASE spectra of the CsPbBr3 films pumped by a 400 nm lasing output with a 5 ns pulse width at a repetition rate of 10 Hz. The top shows the fluorescence spectra at different pump energies and the down shows the fluorescence intensity variation as the pump energy increased. (CsPbBr3 films (a)&(f), CsPbBr3:MA films (b)&(g), CsPbBr3:EA films (c)&(h), CsPbBr3:BA films (d)&(i), CsPbBr3:OA films (e) &(j)).

Download Full Size | PDF

Tables Icon

Table 2. Gain and loss coefficients of the CsPbBr3 films with different organic cations

We measured photostability of the CsPbBr3 films after introducing organic cations. The ASE intensities was monitored as the number of the pump pulse increased. Figure 7 shows the results. We can see the ASE intensity from the perovskite films with organic cations did not change significantly after being excited with 18k pulses at an excitation intensity of 1127 µJ/cm2, 2.8 folds of thresholds. The increment of the ASE thresholds of the films were less than 10% after keeping the films for 5 days under ambient atmosphere (Fig. 13 and Table 3 in supporting information). The ASE intensity fluctuation in the photostability test may result from the unstable pump energies and film degradation under continuous pumping.

 figure: Fig. 7.

Fig. 7. Stability of different organic cations doped perovskite films (Pump energy: 1127 µJ/cm2 above threshold).

Download Full Size | PDF

Tables Icon

Table 3. Photostability of perovskite films at room temperature in air

4. Conclusions

In summary, we have fabricated CsPbBr3 perovskite thin film waveguides doped with a series of organic cations of various chain lengths, namely, MA, EA, BA and OA with a two-step solution process. The crystalline structure, surface morphology and photophysical properties of the films were characterized in detail. Organic cations could fill pinholes, thus reduce loss from nonradiative recombination at grain boundary and surface. We achieved ASE from all of different CsPbBr3 films at RT, with the minimum thresholds of 271 µJ/cm2, 300 µJ/cm2, 338 µJ/cm2, and 399 µJ/cm2 for MA, EA, BA and OA doped CsPbBr3 films, respectively. We also measured the gain coefficients for the doped CsPbBr3 films, which were 48.2 ± 5.9 cm−1, 45.3 ± 2.4 cm−1, 27.9 ± 2.0 cm−1, and 25.4 ± 2.0 cm−1 for MA, EA, BA and OA doped CsPbBr3 thin films, respectively. The intrinsic light loss for MA, EA, BA and OA doped films were 9.6 ± 1.0 cm−1, 8.2 ± 0.6 cm−1, 7.5 ± 0.4 cm−1 and 5.2 ± 0.4 cm−1, respectively. The organic cations incorporated films presented higher gain and lower optical loss, due to reduced bulk defects and non-radiative recombination in the films as well as improved surface quality of the films.

Appendix

A. CsPbBr3 film image

 figure: Fig. 8.

Fig. 8. Images of the CsPbBr3 films prepared at different soaking time.

Download Full Size | PDF

B. XRD of CsPbBr3 film

 figure: Fig. 9.

Fig. 9. XRD patterns of CsPbBr3 films prepared at different soaking time.

Download Full Size | PDF

C. AFM image of perovskite films

 figure: Fig. 10.

Fig. 10. AFM images of different organic cations doped perovskite films.

Download Full Size | PDF

D. Absorption and PL spectra of perovskite films

 figure: Fig. 11.

Fig. 11. Absorption spectra ((a). CsPbBr3, (c). CsPbBr3:MA, (e). CsPbBr3:EA, (g). CsPbBr3:BA, (i). CsPbBr3:OA) and PL spectra ((b). CsPbBr3, (d). CsPbBr3:MA, (f). CsPbBr3:EA, (h). CsPbBr3:BA, (j). CsPbBr3:OA) of different organic cations doped perovskite films (excitation wavelength: 400 nm).

Download Full Size | PDF

E. Gain and loss of the perovskite films

 figure: Fig. 12.

Fig. 12. Gain and loss of the CsPbBr3 films (CsPbBr3 films (a)&(b), CsPbBr3:MA films (c)&(d), CsPbBr3:EA films (e)&(f), CsPbBr3:BA films (g)&(h), CsPbBr3:OA films (i) &(j)).

Download Full Size | PDF

F. Stability of perovskite film after 5 days in air

 figure: Fig. 13.

Fig. 13. Stability study of perovskite films at room temperature in air (The left column shows the initial ASE thresholds; the right column shows the ASE thresholds after 5 days) ((a),(b) CsPbBr3 film, (c),(d) CsPbBr3:MA film, (e),(f) CsPbBr3:EA film, (g),(h) CsPbBr3:BA film, i,j CsPbBr3:OA film).

Download Full Size | PDF

Funding

National Natural Science Foundation of China (2015CB932200, 61605075).

Disclosures

The authors declare no conflicts of interest.

References

1. B. R. Sutherland and E. H. Sargent, “Perovskite photonic sources,” Nat. Photonics 10(5), 295–302 (2016). [CrossRef]  

2. G. Xing, N. Mathews, S. Sun, S. S. Lim, Y. M. Lam, M. Graetzel, S. Mhaisalkar, and T. C. Sum, “Long-Range Balanced Electron- and Hole-Transport Lengths in Organic-Inorganic CH3NH3PbI3,” Science 342(6156), 344–347 (2013). [CrossRef]  

3. H. Kim, A. Hagfeldt, and N. Park, “Morphological and compositional progress in halide perovskite solar cells,” Chem. Commun. 55(9), 1192–1200 (2019). [CrossRef]  

4. S. Yakunin, L. Protesescu, F. Krieg, M. I. Bodnarchuk, G. Nedelcu, M. Humer, G. D. Luca, M. Fiebig, W. Heiss, and M. V. Kovalenko, “Low-threshold amplified spontaneous emission and lasing from colloidal nanocrystals of caesium lead halide perovskites,” Nat. Commun. 6(1), 8056 (2015). [CrossRef]  

5. Y. Cao, N. Wang, H. Tian, J. Guo, Y. Wei, H. Chen, Y. Miao, W. Zou, K. Pan, Y. He, H. Cao, Y. Ke, M. Xu, Y. Wang, M. Yang, K. Du, Z. Fu, D. Kong, D. Dai, Y. Jin, G. Li, H. Li, Q. Peng, J. Wang, and W. Huang, “Perovskite light-emitting diodes based on spontaneously formed submicrometre-scale structures,” Nature 562(7726), 249–253 (2018). [CrossRef]  

6. G. Xing, N. Mathews, S. S. Lim, N. Yantara, X. Liu, D. Sabba, M. Gratzel, S. Mhaisalkar, and T. C. Sum, “Low-temperature solution-processed wavelength-tunable perovskites for lasing,” Nat. Mater. 13(5), 476–480 (2014). [CrossRef]  

7. H. Zhu, Y. Fu, F. Meng, X. Wu, Z. Gong, Q. Ding, M. V. Gustafsson, M. T. Trinh, S. Jin, and X. Zhu, “Lead halide perovskite nanowire lasers with low lasing thresholds and high quality factors,” Nat. Mater. 14(6), 636–642 (2015). [CrossRef]  

8. Z. Gu, K. Wang, W. Sun, J. Li, S. Liu, Q. Song, and S. Xiao, “Two-Photon Pumped CH3NH3PbBr3 Perovskite Microwire Lasers,” Adv. Opt. Mater. 4(3), 472–479 (2016). [CrossRef]  

9. B. Salhi, Y. S. Wudil, M. K. Hossain, A. Al-Ahmed, and F. A. Al-Sulaiman, “Review of recent developments and persistent challenges in stability of perovskite solar cells,” Renewable Sustainable Energy Rev. 90, 210–222 (2018). [CrossRef]  

10. T. S. Kao, Y. Chou, C. Chou, F. Chen, and T. Lu, “Lasing behaviors upon phase transition in solution-processed perovskite thin films,” Appl. Phys. Lett. 105(23), 231108 (2014). [CrossRef]  

11. H. Zhou, S. Yuan, X. Wang, T. Xu, X. Wang, H. Li, W. Zheng, P. Fan, Y. Li, and L. Sun, “Vapor Growth and Tunable Lasing of Band Gap Engineered Cesium Lead Halide Perovskite Micro/Nanorods with Triangular Cross Section,” ACS Nano 11(2), 1189–1195 (2017). [CrossRef]  

12. B. R. Sutherland, S. Hoogland, M. M. Adachi, C. T. O. Wong, and E. H. Sargent, “Conformal Organohalide Perovskites Enable Lasing on Spherical Resonators,” ACS Nano 8(10), 10947–10952 (2014). [CrossRef]  

13. Y. Wang, X. Li, J. Song, L. Xiao, H. Zeng, and H. Sun, “All-Inorganic Colloidal Perovskite Quantum Dots: A New Class of Lasing Materials with Favorable Characteristics,” Adv. Mater. 27(44), 7101–7108 (2015). [CrossRef]  

14. D. Zhang, S. W. Eaton, Y. Yu, L. Dou, and P. Yang, “Solution-Phase Synthesis of Cesium Lead Halide Perovskite Nanowires,” J. Am. Chem. Soc. 137(29), 9230–9233 (2015). [CrossRef]  

15. Q. Liao, K. Hu, H. Zhang, X. Wang, J. Yao, and H. Fu, “Perovskite Microdisk Microlasers Self-Assembled from Solution,” Adv. Mater. 27(22), 3405–3410 (2015). [CrossRef]  

16. F. Sasaki, H. Mochizuki, Y. Zhou, Y. Sonoda, and R. Azumi, “Optical pumped lasing in solution processed perovskite semiconducting materials: Self-assembled microdisk lasing,” Jpn. J. Appl. Phys. 55(4S), 04ES02 (2016). [CrossRef]  

17. K. Wang, W. Sun, J. Li, Z. Gu, S. Xiao, and Q. Song, “Unidirectional Lasing Emissions from CH3NH3PbBr3 Perovskite Microdisks,” ACS Photonics 3(6), 1125–1130 (2016). [CrossRef]  

18. S. Liu, W. Sun, Z. Gu, K. Wang, N. Zhang, S. Xiao, and Q. Song, “Tailoring the lasing modes in CH3NH3PbBr3 perovskite microplates via micro-manipulation,” RSC Adv. 6(56), 50553–50558 (2016). [CrossRef]  

19. B. Tang, H. Dong, L. Sun, W. Zheng, Q. Wang, F. Sun, X. Jiang, A. Pan, and L. Zhang, “Single-Mode Lasers Based on Cesium Lead Halide Perovskite Submicron Spheres,” ACS Nano 11(11), 10681–10688 (2017). [CrossRef]  

20. S. Chen, K. Roh, J. Lee, W. K. Chong, Y. Lu, N. Mathews, T. C. Sum, and A. Nurmikko, “A Photonic Crystal Laser from Solution Based Organo-Lead Iodide Perovskite Thin Films,” ACS Nano 10(4), 3959–3967 (2016). [CrossRef]  

21. C. Li, Z. Zang, C. Han, Z. Hu, X. Tang, J. Du, Y. Leng, and K. Sun, “Highly compact CsPbBr3 perovskite thin films decorated by ZnO nanoparticles for enhanced random lasing,” Nano Energy 40, 195–202 (2017). [CrossRef]  

22. N. Yantara, S. Bhaumik, F. Yan, D. Sabba, H. A. Dewi, N. Mathews, P. P. Boix, H. V. Demir, and S. Mhaisalkar, “Inorganic Halide Perovskites for Efficient Light-Emitting Diodes,” J. Phys. Chem. Lett. 6(21), 4360–4364 (2015). [CrossRef]  

23. S. W. Eaton, M. Lai, N. A. Gibson, A. B. Wong, L. Dou, J. Ma, L. Wang, S. R. Leone, and P. Yang, “Lasing in robust cesium lead halide perovskite nanowires,” Proc. Natl. Acad. Sci. U. S. A. 113(8), 1993–1998 (2016). [CrossRef]  

24. X. Tang, Z. Hu, W. Chen, X. Xing, W. Hu, J. Qiu, Z. Zang, J. Du, Y. Leng, X. Jiang, and L. Mai, “Room temperature single-photon emission and lasing for all-inorganic colloidal perovskite quantum dots,” Nano Energy 28, 462–468 (2016). [CrossRef]  

25. D. Priante, I. Dursun, M. S. Alias, D. Shi, V. A. Melnikov, T. K. Ng, O. F. Mohammed, O. M. Bakr, and B. S. Ooi, “The recombination mechanisms leading to amplified spontaneous emission at the true-green wavelength in CH3NH3PbBr3 perovskites,” Appl. Phys. Lett. 106(8), 081902 (2015). [CrossRef]  

26. M. L. De Giorgi, A. Perulli, N. Yantara, P. P. Boix, and M. Anni, “Amplified Spontaneous Emission Properties of Solution Processed CsPbBr3 Perovskite Thin Films,” J. Phys. Chem. C 121(27), 14772–14778 (2017). [CrossRef]  

27. Y. Ling, Y. Tian, X. Wang, J. C. Wang, J. M. Knox, F. Perez-Orive, Y. Du, L. Tan, K. Hanson, B. Ma, and H. Gao, “Enhanced Optical and Electrical Properties of Polymer-Assisted All-Inorganic Perovskites for Light-Emitting Diodes,” Adv. Mater. 28(40), 8983–8989 (2016). [CrossRef]  

28. L. Zhang, X. Yang, Q. Jiang, P. Wang, Z. Yin, X. Zhang, H. Tan, Y. M. Yang, M. Wei, B. R. Sutherland, E. H. Sargent, and J. You, “Ultra-bright and highly efficient inorganic based perovskite light-emitting diodes,” Nat. Commun. 8(1), 15640 (2017). [CrossRef]  

29. Y. Zhao and K. Zhu, “Organic-inorganic hybrid lead halide perovskites for optoelectronic and electronic applications,” Chem. Soc. Rev. 45(3), 655–689 (2016). [CrossRef]  

30. Y. Wang, L. Yang, H. Chen, C. Liu, W. Liu, H. Xu, C. Zhang, Z. Wang, and Y. Liu, “White LED based on CsPbBr3 nanocrystal phosphors via a facile two-step solution synthesis route,” Mater. Res. Bull. 104, 48–52 (2018). [CrossRef]  

31. D. Liu, Z. Hu, W. Hu, P. Wangyang, K. Yu, M. Wen, Z. Zu, J. Liu, M. Wang, W. Chen, M. Zhou, X. Tang, and Z. Zang, “Two-step method for preparing all-inorganic CsPbBr3 perovskite film and its photoelectric detection application,” Mater. Lett. 186, 243–246 (2017). [CrossRef]  

32. L. Meng, J. You, T. Guo, and Y. Yang, “Recent Advances in the Inverted Planar Structure of Perovskite Solar Cells,” Acc. Chem. Res. 49(1), 155–165 (2016). [CrossRef]  

33. N. J. Jeon, J. H. Noh, W. S. Yang, Y. C. Kim, S. Ryu, J. Seo, and S. I. Seok, “Compositional engineering of perovskite materials for high-performance solar cells,” Nature 517(7535), 476–480 (2015). [CrossRef]  

34. X. Zhang, B. Xu, J. Zhang, Y. Gao, Y. Zheng, K. Wang, and X. W. Sun, “All-Inorganic Perovskite Nanocrystals for High-Efficiency Light Emitting Diodes: Dual-Phase CsPbBr3-CsPb2Br5 Composites,” Adv. Funct. Mater. 26(25), 4595–4600 (2016). [CrossRef]  

35. C. C. Stoumpos, C. D. Malliakas, and M. G. Kanatzidis, “Semiconducting tin and lead iodide perovskites with organic cations: phase transitions, high mobilities, and near-infrared photoluminescent properties,” Inorg. Chem. 52(15), 9019–9038 (2013). [CrossRef]  

36. D. M. Trots and S. V. Myagkota, “High-temperature structural evolution of caesium and rubidium triiodoplumbates,” J. Phys. Chem. Solids 69(10), 2520–2526 (2008). [CrossRef]  

37. X. Zhang, X. Bing, J. Zhang, G. Yuan, and W. S. Xiao, “All-Inorganic Perovskite Nanocrystals for High-Efficiency Light Emitting Diodes: Dual-Phase CsPbBr3 -CsPb2Br5 Composites,” Adv. Funct. Mater. 26(25), 4595–4600 (2016). [CrossRef]  

38. Y. Liu, Z. Gao, W. Zhang, X. Sun, Z. Wang, X. Wang, B. Xu, and X. Meng, “Stimulated emission from CsPbBr3 quantum dot nanoglass,” Opt. Mater. Express 9(8), 3390 (2019). [CrossRef]  

39. Z. Yuan, Y. Shu, Y. Tian, Y. Xin, and B. Ma, “A facile one-pot synthesis of deep blue luminescent lead bromide perovskite microdisks,” Chem. Commun. 51(91), 16385–16388 (2015). [CrossRef]  

40. H. Cho, S. Jeong, M. Park, Y. Kim, C. Wolf, C. Lee, J. H. Heo, A. Sadhanala, N. Myoung, S. Yoo, S. H. Im, R. H. Friend, and T. Lee, “Overcoming the electroluminescence efficiency limitations of perovskite light-emitting diodes,” Science 350(6265), 1222–1225 (2015). [CrossRef]  

41. D. Shi, V. Adinolfi, R. Comin, M. Yuan, E. Alarousu, A. Buin, Y. Chen, S. Hoogland, A. Rothenberger, K. Katsiev, Y. Losovyj, X. Zhang, P. A. Dowben, O. F. Mohammed, E. H. Sargent, and O. M. Bakr, “Low trap-state density and long carrier diffusion in organolead trihalide perovskite single crystals,” Science 347(6221), 519–522 (2015). [CrossRef]  

42. J. Valenta, I. Pelant, and J. Linnros, “Waveguiding effects in the measurement of optical gain in a layer of Si nanocrystals,” Appl. Phys. Lett. 81(8), 1396–1398 (2002). [CrossRef]  

43. B. Mitzi D, “Templating and structural engineering in organic–inorganic perovskites,” Journal Of The Chemical Society Dalton Transactions 1, 1–12 (2001).

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 (13)

Fig. 1.
Fig. 1. Schematic diagram of the fabrication process of inorganic perovskite films.
Fig. 2.
Fig. 2. SEM images of different organic cations doped perovskite films ((a). CsPbBr3, (b). CsPbBr3:MA, (c). CsPbBr3:EA, (d). CsPbBr3:BA, (e). CsPbBr3:OA).
Fig. 3.
Fig. 3. XRD patterns of the additives doped perovskite films.
Fig. 4.
Fig. 4. Absorption spectra (a) and normalized PL spectra (b) of the perovskite films doped with different organic cations (Excitation wavelength: 400 nm).
Fig. 5.
Fig. 5. Time-resolved fluorescence spectra of the perovskite films with organic cations.
Fig. 6.
Fig. 6. ASE spectra of the CsPbBr3 films pumped by a 400 nm lasing output with a 5 ns pulse width at a repetition rate of 10 Hz. The top shows the fluorescence spectra at different pump energies and the down shows the fluorescence intensity variation as the pump energy increased. (CsPbBr3 films (a)&(f), CsPbBr3:MA films (b)&(g), CsPbBr3:EA films (c)&(h), CsPbBr3:BA films (d)&(i), CsPbBr3:OA films (e) &(j)).
Fig. 7.
Fig. 7. Stability of different organic cations doped perovskite films (Pump energy: 1127 µJ/cm2 above threshold).
Fig. 8.
Fig. 8. Images of the CsPbBr3 films prepared at different soaking time.
Fig. 9.
Fig. 9. XRD patterns of CsPbBr3 films prepared at different soaking time.
Fig. 10.
Fig. 10. AFM images of different organic cations doped perovskite films.
Fig. 11.
Fig. 11. Absorption spectra ((a). CsPbBr3, (c). CsPbBr3:MA, (e). CsPbBr3:EA, (g). CsPbBr3:BA, (i). CsPbBr3:OA) and PL spectra ((b). CsPbBr3, (d). CsPbBr3:MA, (f). CsPbBr3:EA, (h). CsPbBr3:BA, (j). CsPbBr3:OA) of different organic cations doped perovskite films (excitation wavelength: 400 nm).
Fig. 12.
Fig. 12. Gain and loss of the CsPbBr3 films (CsPbBr3 films (a)&(b), CsPbBr3:MA films (c)&(d), CsPbBr3:EA films (e)&(f), CsPbBr3:BA films (g)&(h), CsPbBr3:OA films (i) &(j)).
Fig. 13.
Fig. 13. Stability study of perovskite films at room temperature in air (The left column shows the initial ASE thresholds; the right column shows the ASE thresholds after 5 days) ((a),(b) CsPbBr3 film, (c),(d) CsPbBr3:MA film, (e),(f) CsPbBr3:EA film, (g),(h) CsPbBr3:BA film, i,j CsPbBr3:OA film).

Tables (3)

Tables Icon

Table 1. Absorption spectra and fluorescence spectra of the CsPbBr3 films doped with organic cations

Tables Icon

Table 2. Gain and loss coefficients of the CsPbBr3 films with different organic cations

Tables Icon

Table 3. Photostability of perovskite films at room temperature in air

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.