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

Low amplified spontaneous emission threshold from 2-thiophenemethylammonium quasi-2D perovskites via phase engineering

Open Access Open Access

Abstract

Quasi-2D Ruddlesden-Popper perovskites attract great attention as an optical gain media in lasing applications due to their excellent optoelectronic properties. Herein, a novel quasi-2D Ruddlesden-Popper perovskite based on 2-thiophenemethylammonium (ThMA) is synthesized by a facile solution-processed method. In addition, an anti-solvent treatment method is proposed to tune the phase distribution, and preferential orientation of quasi-2D (ThMA)2Csn-1PbnBr3n+1 thin films. The large-n-dominated narrow domain distribution improves the energy transfer efficiency from small-n to large-n phases. Also, the highly oriented nanocrystals facilitate the efficient Förster energy transfer, beneficial for the carrier population transfer. Furthermore, a green amplified spontaneous emission with a low threshold of 13.92 µJ/cm2 is obtained and a single-mode vertical-cavity laser with an 0.4 nm linewidth emission is fabricated. These findings provide insights into the design of the domain distribution to realize low-threshold multicolor continuous-wave or electrically driven quasi-2D perovskites laser.

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

1. Introduction

Metal halide perovskites have emerged as prospective semiconductors for optoelectronic devices spanning solar cells, light-emitting diodes (LEDs), and lasers due to excellent properties, such as high absorption coefficient, tunable bandgap, high carrier mobility, solution processability, and low cost [13]. However, halogen ion migration and instability under the humid atmosphere have hindered the practical application of the three-dimensional (3D) perovskites [4]. Many strategies including interface engineering, solvent engineering, and polymeric doping are used to overcome these obstacles [57]. Introducing large hydrophobic cations to form quasi-2D perovskites has attracted worldwide research interest due to their enhanced high resistance to moisture and air compared to traditional 3D perovskites [8,9]. Ruddlesden-Popper perovskites (RPPs), as one of the well-known quasi-2D perovskites, has a chemical formula of T2An-1MnX3n+1, where T donates large organic cations such as phenemethylammonium (PEA+), A represents monovalent cation like Cs+ or formamidinium (FA+), M is a divalent metal cation such as Pb2+ or Sn2+, and X is a halide anion (Cl, Br, or I), and n defines the number of [MX6]4- octahedral slabs between organic cations [10]. The alternating inorganic and organic sheets with different energy bandgaps (Eg) as potential wells and barriers, respectively, generate an ordered self-assembly multiple quantum wells (QWs) structure [11]. Such QW distribution provides an energy funnel from the large-Eg domains of small-n (n ≤ 3) phases (donor domain) to the low-Eg domains of large-n phases (acceptor domain) for photogenerated carriers, which contribute to the establishment of population inversion, allowing amplified spontaneous emission (ASE) and lasing in the large-n phases of quasi-2D RPPs [12,13]. Thus, it makes them promising for LED and lasing applications [14,15]. However, it is still lagging behind its 3D counterparts in the development of quasi-2D RPPs with high optical gain and low threshold. Recently, a novel set of bulky thiophene-based ligands containing electron-rich π-conjugated functional groups, such as 2-thiophenemethylammonium (ThMA+) cations, have been introduced into quasi-2D perovskite as spacer blocks to modify the near-band edge integrity and improves charge transport within the semiconductor [16]. It is reported that the incorporation of ThMA+ can improve carrier lifetime and reduce recombination losses through structure reconstruction and defect reduction [17]. However, there are no reports of using ThMA+ in quasi-2D RPPs for LEDs and lasers. Introducing alternative spacer cations and focusing on the ligand optimization to enhance the carrier transport and reduce the van der Waals gaps between quasi-2D perovskite layers are important to promote the energy transfer and obtain a low-threshold [18].

On the other hand, the inhomogeneous phase distribution in quasi-2D RPPs firmly impresses the energy transfer between small-n with high-Eg QWs and large-n phases with low-Eg QWs [13,1921]. There are plenty of lattice defects in the small-n phases with random stacking low-D structures which tend to form due to their low formation energy [22]. The quasi-2D RPPs dominated by small-n phases are not suitable for ASE because of the high density of defect states which tend to increase the electron−phonon coupling effect and non-radiative recombination, and thus reduce the efficiency of energy transfer and light emission [23,24]. Furthermore, the excessive organic barriers in wide n-phase RPPs would lead to energy loss or interrupts the transmitted energy before reaching the acceptor domain of the large-n phases during the energy transfer process. However, it has been reported that inducing more small-n phases in quasi-2D perovskite provide a graded energy cascade pathway to facilitate more efficient energy transfer processes [25]. In addition, it is a challenge to lower the impact of the rapid Auger recombination in the carrier transfer process to enhance the ASE for quasi-2D RPPs [26,27]. Therefore, a deep understanding of the relation between the controlled phase distribution (e.g., narrow-n phases) and ASE properties including the excitonic nature is necessary to improve the laser emission in the quasi-2D RPP films [2830].

In this work, we have synthesized quasi-2D (ThMA)2Csn-1PbnBr3n+1 (<n>= 1-5) RPPs by using ThMA ligand. Narrow-n phase distribution and highly oriented crystal domains for the (ThMA)2Cs2Pb3Br10 films are achieved by antisolvent engineering. The high-quality thin films with narrow-n phase distributions and preferential orientation enable efficient and more complete energy transfer to acceptor domains for low-threshold ASE. We have fabricated a vertical-cavity surface-emitting laser based on the (ThMA)2Cs2Pb3Br10 thin films and realized the output of single-mode.

2. Material and methods

2.1 Material and methods

ThMABr, CsBr, and PbBr2 were purchased from Xi'an Concentrator Technology. The dimethylformamide (DMF, 99.999%), dimethyl sulfone (DMSO, 99.999%), and chlorobenzene (CB, 99.9%) were purchased from J & K Scientific. All chemicals were directly used without further purification.

2.2 Preparation of RPP films and devices

The (ThMA)2Csn-1PbnBr3n+1 films were fabricated by spin-coating a precursor solution where the ThMABr, CsBr, and PbBr2 were dissolved in a certain amount of DMSO: DMF (1: 1) according to the standard stoichiometric ratio of the RPP. The prepared solution was stirred at room temperature for 2 hours. Subsequently, the precursor solution (30 µL) was quickly dropped on the glass substrate and rotated at 4000 rounds per minute (rpm) for 50 s. During the rotation, the CB was added dropwise at various times. The (ThMA)2Cs2Pb3Br10 films tend to be whitish when the CB was dropped before 19 s, which can be attributed to the incomplete crystallization of the surface. Thus, 19 s was chosen as the starting time of dripping CB for the experiments. All samples were heated at 80 °C for 10 minutes. All the operations above are conducted in a glove box filled with an argon atmosphere. The Fabry-Pérot cavity was fabricated by sandwiching a CB-processed (ThMA)2Cs2Pb3Br10 quasi-2D perovskite film between two distributed Bragg reflectors (DBRs) mirrors. The DBR mirrors showed a high reflectivity of 98% around 530 nm. The perovskite film was spin-coated on a clean DBR substrate at 4000 rpm for 50 s and thermally annealed on a hot plate at 80 °C for 15 min in an argon-filled glove box. Then, another DBR mirror was bonded to it with an adhesive to form the quasi-2D-RP perovskite thin film device.

2.3 Measurements of materials and devices

The crystal structures and phases of (ThMA)2Csn-1PbnBr3n+1 films were characterized by using the X-ray diffractometer (D8 Discover, Bruker) with a scanning rate of 0.05°/s in a 2θ ranging from 2° to 50° with Cu Kα radiation. The morphology of the films was analyzed by using a field-effect scanning electron microscopy (SEM, S-4800). The absorption spectra and PL spectra of (ThMA)2Csn-1PbnBr3n+1 films were measured by UV–vis spectrophotometer (Cary-60, Agilent) and a fluorescence spectrometer (FLS980, Edinburgh Instruments), respectively. A portion of the fundamental laser pulses (800 nm, 1 kHz, 7 mJ pulse−1, 35 fs, Coherent) was directed vertically onto the quasi-2D RPP thin films via a neutral density filter (USB4000-UV-VIS, Ocean Optics), and the emission from the film's edges was vertically collected and detected with a fiber spectrometer. The TA measurements were carried out on a pump-probe system (Helios, Ultrafast Systems). White-light probe pulses (390-600 nm) were produced by focusing a portion of the laser pulse (800 nm, 1 kHz, 7 mJ pulse−1, 35 fs, Coherent) onto a calcium fluoride crystal. Pump pulses centered at 320 nm were generated from the other part pulse through the optical parametric amplifier (TOPAS, 800 fs). The instrument response function was about 100 fs.

3. Experiments and results

The (ThMA)2Csn-1PbnBr3n+1 films are fabricated by the spin-coating method, where the composition <n > is manipulated by the stoichiometric ratio of the precursors, and the detailed information is presented in Experimental Section. Figure 1(a) shows the schematics of (ThMA)2Csn-1PbnBr3n+1 (n = 1, 2, 3, and ∞). There are n sheets of PbBr6 octahedra and one ThMA layer in the n-phase RPP. X-ray diffraction (XRD) measurements (Fig. 1(b)) are conducted to verify the evolution of the n-phases for (ThMA)2Csn-1PbnBr3n+1 with various 〈n〉-composition. The XRD peaks with 2θ values of 6.4°, 12.6°, 18.7°, 24.9°, and 31.2° correspond to the facets of (002), (004), (006), (008) and (0010) for (ThMA)2Csn-1PbnBr3n+1 with n = 1 according to Bragg's Law [31]. For <n> = 2-5 RPP films, Bragg reflections at a low angle of 6.4° is a feature of the 2D layered perovskite, and their relative intensity decreases with the increasing <n > values [16]. Besides these diffraction peaks of n = 1 phase, there are peaks at 15.6° and 31.9° in <n> = 2-5 films, representing the crystallographic planes of (100) and (200) of the 3D cubic CsPbBr3, respectively, which suggests that CsPbBr3 nanocrystals appear [32]. Thus the small/intermediate-n and large-n phases coexist in the (ThMA)2Csn-1PbnBr3n+1 films although the small-n phases are decreasing with the increasing composition <n > values [33,34].

 figure: Fig. 1.

Fig. 1. Structure and optical properties of (ThMA)2Csn-1PbnBr3n+1 with different compositions. (a) Schematics for (ThMA)2Csn-1PbnBr3n+1 (n = 1, 2, 3, and ∞). (b) XRD patterns for (ThMA)2Csn-1PbnBr3n+1 films. (c) UV-vis absorption (solid line) and PL spectra (dashed line) for (ThMA)2Csn-1PbnBr3n+1 films.

Download Full Size | PDF

The optical properties of (ThMA)2Csn-1PbnBr3n+1 thin films were measured by the UV-vis optical absorption spectra and photoluminescence (PL) spectra (Fig. 1(c)). There is a sharp exciton absorption peak at 407 nm and a PL emission peak at 413 nm for pure 2D (ThMA)2PbBr4 films. Owing to the exciton-phonon scattering, the PL peak of (ThMA)2PbBr4 is red-shifted about 6 nm relative to the absorption peak [27]. The exciton absorption peaks at 407 nm, 436 nm, 464 nm, and 508 nm of the (ThMA)2Csn-1PbnBr3n+1 films with <n> = 2-5 ascribe to the layered phases with n = 1, 2, 3, and ≥ 5, respectively. The band gaps of (ThMA)2Csn-1PbnBr3n+1 with n = 2-5 are estimated from the onset of the Tauc plots to be 3.05 eV, 2.84 eV, 2.67 eV and 2.43 eV, respectively. In addition, the absorption peaks of small-n phases are progressively weakened with the increase of <n>, indicating the percentage of small-n QWs decreases gradually. The PL peaks at 413 nm 440 nm and 468 nm for the small phases of n = 1, 2, and 3, respectively, are weaker very much than that of the large-n phases for the (ThMA)2Csn-1PbnBr3n+1 films with <n> = 2-5. The signal of the large-n phase with low-Eg in films <n> = 2-5 dominates the PL spectra, indicating an energy cascade from the small-n QWs to large-n ones. The red-shift of the excitonic signals for absorption and PL (Fig. (S1)) with the increasing <n > value (from 1 to 5) is consistent with the previous reports [16].

To explore the optical gain behavior for (ThMA)2Csn-1PbnBr3n+1 films, the optically pumped lasing measurements were performed. For the <n> = 3 pristine films, there is a broad SE peak at 518 nm with a wide full width at half maximum (FWHM) of 22 nm under low pump fluence < 31.81 µJ/cm2 (Fig. (2a)). When the pump fluence exceeds a specific threshold (Pth), a sharp ASE peak develops on the lower-energy side of the SE spectra at 529 nm with an FWHM of 6 nm and finally becomes dominant when the fluence increases from 33.48 to 41.63 µJ/cm2. The pump-fluence-dependent emission intensity in Fig. (2b) indicates that the FWHM is narrowed from 22 to 6 nm as the pump fluence exceeds the Pth of 32.62 µJ/cm2, consistent with the results for (PEA)2(CsPbBr3)4PbBr4 and (NMA)2(FA)Pb2BrI6 [12,34]. The 10 nm red-shift for the ASE peak compared with the SE peak is introduced by the self-absorption effect [31]. There is an analogous transition from the SE to ASE with the increasing pump fluency in the films with <n> = 4 and 5 (Fig. (S2)). What is interesting is that the Pth gradually decreases and the peak position of the ASE signal progressively redshifts as <n > increases. For (ThMA)2PbBr4, there is no effective energy transport due to the lack of large-n phases. The coherent oscillation decaying in the TA kinetic curve is caused by the electron-optical phonon interactions due to strong quantum confinement in <n> = 1 quasi-2D RPP film with a thin inorganic layer (Fig. S3(a)) [35]. The (ThMA)2CsPb2Br7 films present a unique bimodal emission of ASE, where the Pth are 128.56 and 88.95 µJ/cm2 for the peaks at 526 nm and 537 nm, respectively (Fig. S(3b)). There is a competition between the two ASE peaks with the increase of pump fluence for (ThMA)2CsPb2Br7 films.

To further track the energy transfer process and excited-state dynamics, the ultrafast transient absorption (TA) measurements of (ThMA)2Cs2Pb3Br10 films were performed in the range of 380 nm – 600 nm. As shown in Fig. 2(c), four ground state bleaching (GSB) bands of perovskite grains agree with the peaks in the UV-vis absorption spectra (Fig. 1(c)), which are attributed to n = 1, 2, 3, and n ≥ 5 (n = ∞) phases, named as GSBn = 1, GSBn = 2, GSBn = 3, and GSBn≥5, respectively. The GSBn = 1-3 signals for the small-n phases decay rapidly while the peaks keep in the same positions with the increasing delay time from 0.19 ps to 100 ps. However, the broad GSBn≥5 band covering the whole absorption tail shifts gradually from 508 nm to 512 nm with the increasing delay time, because of the multiple QWs with a larger n value to n = ∞ [12]. Figure 2(d) displays the normalized TA kinetic decay curves at the typical probe-wavelengths of 404 nm, 432 nm, and 508 nm for GSBn = 1, GSBn = 2, and GSBn≥5, fitted with multi-exponential function, ΔA(t) = a1exp(−t/τ1) + a2exp(−t/τ2) − c1exp(−t/τet), where a1, a2, and c1 are amplitudes, τ1, and τ2 is the decay time constants, and τet is formation time constant [36]. The dynamic decay curves for GSBn = 1 and GSBn = 2 undergo fast decay progress with time constants of 0.21 ps and 0.44 ps, while the signal for GSBn≥5 shows an increasing kinetic with a rising constant (τet) of 0.84 ps in early delay time. The comparable timescale of bleaching for the small-n phases and excited-state formation at n ≥ 5 phases verify the fast energy transfer from the small-n phases to the large-n phases. Furthermore, the short lifetime in the donor domain and long lifetime in the acceptor domain illustrate that the carrier population transfer from the small-n and large-n phases is conducted by Förster resonance energy transfer (FRET) [37].

 figure: Fig. 2.

Fig. 2. ASE and TA spectra for quasi-2D (ThMA)2Cs2Pb3Br10 films (pristine films). (a) Pump-fluence dependence of PL characteristics under 400 nm excitation. (b) The dependence of FWHM and PL intensity on the pump fluence, respectively. (c) TA spectra for 0.19, 0.52, 1 and 2 ps. (d) Normalized GSB curves at 404, 432 and 508 nm (320 nm excitation, 45 µJ/cm2) corresponding to n = 1, n = 2 and n ≥ 5 QWs, respectively.

Download Full Size | PDF

The energy transfer behavior and the ASE performance are strongly influenced by the morphology and the phase distribution which could be controlled by the anti-solvent method [33,38]. Figure 3 represents the ASE and TA spectra of (ThMA)2Cs2Pb3Br10 films by the treatment of anti-solvent chlorobenzene (CB). Similar ASE characteristics pumped with 400 nm pulse are observed, but the Pth of 13.92 µJ/cm2 is much lower than 32.79 µJ/cm2 of the (ThMA)2Cs2Pb3Br10 pristine film (Fig. 3(a)–3(b)). For the GBS signals, they concentrate around 432 nm and 508 nm ascribed primarily to n = 2 and ≥ 5 phases, suggesting low content of n = 1 and n = 3 phases (Fig. 3(c)). The photoinduced carriers are mainly generated in the n = 2 phase with a fast buildup of GSBn = 2 peak at the initial delay time. As the decay time prolongs from 0.18 ps to 2.00 ps, the photobleaching peaks of small-n phases especially the GSBn = 2 peak fade with a continuous increase of the GSBn = 5 peak, elucidating that the carriers transfer from small-n phases to the large-n phases. Figure 3(d) manifests the dynamic decay curves of the TA spectra at these typical probe-wavelengths of 432 nm and 508 nm for GSBn = 2 and GSBn≥5 of the (ThMA)2Cs2Pb3Br10 films treated with CB. The dynamic decay curve related to GSBn = 2 shows an initial rapid increase in 1.0 ps. However, the absorption at 508 nm corresponding to GSBn≥5 demonstrates a fast increase within 2.0 ps, hinting that the photogenerated carriers undergo energy transfer through the QWs, concentrate at n ≥ 5 phases to endure the process of population inversion and ASE in the (ThMA)2Cs2Pb3Br10 films treated with CB. The fast decay time constant τn = 2 of GBSn = 2 for the donor domain and the τet of GBSn≥5 for the acceptor domain are 0.26 ps and 0.62 ps, respectively, significantly smaller than 0.44 ps and 0.84 ps of the pristine film (Fig. 2(d)). It indicates that the accumulation speed of the population number is faster in (ThMA)2Cs2Pb3Br10 films treated with CB which features narrow phase distribution than that in the pristine film. The fewer small-n phases increase the carrier concentration at large-n phases, making it preferred to form ASE in narrow phase distributed (ThMA)2Cs2Pb3Br10 perovskite films. In contrast, the fast band edge-to-trap process due to the electron-phonon coupling effect in pristine films with excessive small-n phases would trap some carriers and form the bound-biexcitons, leading to incomplete energy transfer to large-n phases [23,31]. The absorption intensity at 508 nm corresponding to GBSn≥5 of the (ThMA)2Cs2Pb3Br10 treated with CB is 0.33 optical density (OD), higher than 0.22 OD of the pristine film (Fig. S4), suggesting most carriers are accumulated at large-n phases and less energy loss during energy transfer in the narrow phase distribution film. The (ThMA)2Cs2Pb3Br10 with a narrow phase distribution favor generating ASE.

 figure: Fig. 3.

Fig. 3. Pump-fluence dependence of the ASE and the TA for (ThMA)2Cs2Pb3Br10 films (with CB). (a) Pump-fluence dependence of PL under 400 nm excitation. (b) The dependence of FWHM and PL intensity as a function of pump fluence, respectively. (c) TA spectra for 0.18, 0.52, 1 and 2 ps (inset: the maximum intensity of different n-phase CBS signals). (d) Normalized GSB kinetics curves at 432 and 508 nm for <n> = 3 films (320 nm excitation, 45 µJ/cm2) corresponding to n = 2 and n ≥ 5 QWs, respectively.

Download Full Size | PDF

To further investigate the effects of the CB treatment for achieving low-threshold ASE, the (ThMA)2Cs2Pb3Br10 films treated with and without CB were characterized by various techniques including SEM, UV–vis absorption, PL, and XRD. The SEM shows that both perovskite films are uniform and compact, but the one treated with CB becomes smoother and denser with the significantly shrunken grain size (Fig. S5). Thus, the CB treatment increases the film uniformity and enhances spatial confinement to promote optical gain. Compared with the prime films, the UV-Vis spectra of the CB treated (ThMA)2Cs2Pb3Br10 films exhibit a narrower-n phase distribution that the n = 1 phases and n = 3 phases mostly vanished (Fig. 4(a)). After CB treatment, the PL peaks for small-n phases nearly disappear except for the main peak at 518 nm for the large-n phases. The emission intensities of the small-n phases for the film treated with CB are two orders of magnitude smaller than the pristine film in the normalized PL spectra where all the peaks are normalized with the intensity of the signal of the large-n phase (Fig. 4(a)). Because of the small content for the small-n phases, the XRD peaks below 10°, characteristic of the 2D small-n perovskites, are not detected due to the resolution for the (ThMA)2Cs2Pb3Br10 films treated with CB (Fig. 4(b)). The FWHM of the peaks at 15.6° and 31.9° corresponding to the large-n phases increases significantly for the CB-treated sample compared with the pristine specimen. According to the Debye-Scherrer formula, the grain size is reduced from 12.8 nm to 9.7 nm after the CB treatment, similar to the results of the CB-treated RPPs based on PEA [34]. Furthermore, the <100 > planes (including (100) at 15.6°and (200) at 31.9°) dominate the XRD peaks for the (ThMA)2Cs2Pb3Br10 films treated with CB, different from the prime films with <001 > and <100 > faces. The results from the XRD patterns manifest that the addition of CB suppresses the formation of the small-n phases and concentrates the phases with a 3D-like single crystal with a <100 > plane. Therefore, the nanocrystal orientation with a <100 > plane tuned by using the CB should play a positive role in the efficient FRET process in the well-oriented thin films because of the alignment of the in-plane transition dipole moments [14].

 figure: Fig. 4.

Fig. 4. The effects of anti-solvent for (ThMA)2Cs2Pb3Br10 films. (a) Upper panel: UV-vis absorption spectra. Low panel: normalized PL spectra (dashed line: the normalized PL spectra of these films in logarithmic). The PL peak shifts from 518 to 513 nm after CB treatment. (b) XRD patterns.

Download Full Size | PDF

The integrated intensity of the GBS signal as a function of the value of n is plotted in Fig. 5(a), indicating that the phase distribution of the film with CB is narrowed and concentrated at n = 2 and n = ∞ phases. The phase distribution could be expressed by the GSB ratio η between small-n phases and large-n phases,

$$\eta \textrm{ = (}\Delta \textrm{OD}_{\textrm{MAX,}n = 2} + \Delta \textrm{OD}_{\textrm{MAX,}n\textrm = 3})/\beta$$
where ΔODMAX,n = 2, and ΔODMAX,n = 3 are the maximum absorption intensity at 430 nm, and 464 nm for the n = 2 and 3 phases, and β is the absorption intensity for large n-phase, respectively. As the CB dripping time (tCB) increases from 19 s to 25 s, the threshold Pth declines from 65 µJ/cm2 to 11.51 µJ/cm2 while the GSB ratio η rises from 21.91% to 43.44% that the signals of GBSn = 1 and GBSn = 3 drop weaken or almost vanished (Fig. 5(b)). Then the Pth rises 58.66 µJ/cm2 when the η value drops below 21.91% for tCB = 30 s, and keeps at the value for tCB = 30 s – 35 s. Thus the lowest threshold can be achieved when the population of intermediate small-n phases (n = 2) is enhanced to maximum and most excitons quickly transfer to the n = 5 phase in the (ThMA)2Cs2Pb3Br10 films treated with CB.

 figure: Fig. 5.

Fig. 5. The phase distribution and the threshold caused by anti-solvent CB for (ThMA)2Cs2Pb3Br10 films. (a) the integrated intensities of the GBS peak of the pristine films (cyan-blue) and the films with CB (purple) for n = 1, 2, 3, and ∞. (b) The Pth and η of (ThMA)2Cs2Pb3Br10 films with the various dripping time of the CB.

Download Full Size | PDF

An optically pumped green laser has been fabricated by sandwiching the CB treated (ThMA)2Cs2Pb3Br10 perovskite between two DBR mirrors. The DBR mirrors demonstrated a high reflection rate of 98% around 530 nm (Fig. 6(a)). The thickness of the gain medium is 165.2 nm measured by the SEM (Fig. S6). The single-mode lasing is obtained at 530 nm with a low FWHM of 0.4 nm. To obtain the threshold for green lasing, the pump-dependent PL spectra on ThMA-based devices were performed. Figure 6(b) exhibits the PL spectra with the typical pump fluences from 3.63 µJ/cm2 to 40.21 µJ/cm2. The dependence of PL intensity as a function of pump fluence in the inset of Fig. 6(b) show that the threshold of the green laser is 15.21 µJ/cm2. These data demonstrate that the (ThMA)2Cs2Pb3Br10 RPPs treated with CB anti-solvent is a promising optical gain material.

 figure: Fig. 6.

Fig. 6. Lasing properties. (a) Schematic diagram of the vertical-cavity surface-emitting laser. (b) Evolution of PL spectra under various pump fluences. Inset: The dependence of PL intensity as a function of pump fluence.

Download Full Size | PDF

4. Conclusion

In summary, the (ThMA)2Csn-1PbnBr3n+1 quasi-2D perovskites are firstly synthesized via the facile solution method. The phase distribution and preferential orientation of the (ThMA)2Csn-1PbnBr3n+1 films are optimized by the treatment of the anti-solvent CB. An efficient Förster energy transfer is obtained, inducing a low Pth of 13.92 µJ/cm2 at room temperature in the (ThMA)2Cs2Pb3Br10 films treated with CB. Single-mode vertical-cavity lasing with 0.4 nm linewidth is realized based on the perovskite thin film with the narrow phase distribution. The results would guide the realization of low-threshold quasi-2D perovskite laser gain media.

Funding

National Natural Science Foundation of China (12074104, 12174090, 11874011); Natural Science Foundation of Henan Province (222300420057); Young Backbone Teacher Training Program in Higher Education of Henan Province (2019GGJS065).

Disclosures

The authors declare no conflicts of interest. The authors have no conflicts to disclose.

Data availability

The data underlying the results presented herein are not publicly available currently but can be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. Z. Ni, C. Bao, Y. Liu, Q. Jiang, W.-Q. Wu, S. Chen, X. Dai, B. Chen, B. Hartweg, Z. Yu, Z. Holman, and J. Huang, “Resolving spatial and energetic distributions of trap states in metal halide perovskite solar cells,” Science 367(6484), 1352–1358 (2020). [CrossRef]  

2. 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]  

3. L. N. Quan, B. P. Rand, R. H. Friend, S. G. Mhaisalkar, T.-W. Lee, and E. H. Sargent, “Perovskites for Next-Generation Optical Sources,” Chem. Rev. 119(12), 7444–7477 (2019). [CrossRef]  

4. H. Cho, Y. H. Kim, C. Wolf, H. D. Lee, and T. W. Lee, “Improving the Stability of Metal Halide Perovskite Materials and Light-Emitting Diodes,” Adv. Mater. 30(42), 1704587 (2018). [CrossRef]  

5. S. Liu, J. Lyu, D. Zhou, X. Zhuang, Z. Shi, R. Sun, L. Liu, Y. Wu, B. Liu, D. Liu, and H. Song, “Dual Modification Engineering via Lanthanide-Based Halide Quantum Dots and Black Phosphorus Enabled Efficient Perovskite Solar Cells with High Open-Voltage of 1.235 V,” Adv. Funct. Mater. 32(19), 2112647 (2022). [CrossRef]  

6. L. Chao, T. Niu, W. Gao, C. Ran, L. Song, Y. Chen, and W. Huang, “Solvent Engineering of the Precursor Solution toward Large-Area Production of Perovskite Solar Cells,” Adv. Mater. 33(14), 2005410 (2021). [CrossRef]  

7. N. Yaghoobi Nia, M. Zendehdel, M. Abdi-Jalebi, L. A. Castriotta, F. U. Kosasih, E. Lamanna, M. M. Abolhasani, Z. Zheng, Z. Andaji-Garmaroudi, K. Asadi, G. Divitini, C. Ducati, R. H. Friend, and A. Di Carlo, “Beyond 17% stable perovskite solar module via polaron arrangement of tuned polymeric hole transport layer,” Nano Energy 82, 105685 (2021). [CrossRef]  

8. E. Amerling, H. Lu, B. W. Larson, A. E. Maughan, A. Phillips, E. Lafalce, L. Whittaker-Brooks, J. J. Berry, M. C. Beard, Z. V. Vardeny, and J. L. Blackburn, “A Multi-Dimensional Perspective on Electronic Doping in Metal Halide Perovskites,” ACS Energy Lett. 6(3), 1104–1123 (2021). [CrossRef]  

9. C. Qin, A. S. D. Sandanayaka, C. Zhao, T. Matsushima, D. Zhang, T. Fujihara, and C. Adachi, “Stable room-temperature continuous-wave lasing in quasi-2D perovskite films,” Nature 585(7823), 53–57 (2020). [CrossRef]  

10. C. Qin, L. Xu, Z. Zhou, J. Song, S. Ma, Z. Jiao, and Y. Jiang, “Carrier dynamics in two-dimensional perovskites: Dion–Jacobson vs. Ruddlesden–Popper thin films,” J. Mater. Chem. A 10(6), 3069–3076 (2022). [CrossRef]  

11. T. He, Y. Jiang, X. Xing, and M. Yuan, “Structured Perovskite Light Absorbers for Efficient and Stable Photovoltaics,” Adv. Mater. 32(26), 1903937 (2020). [CrossRef]  

12. M. Li, Q. Gao, P. Liu, Q. Liao, H. Zhang, J. Yao, W. Hu, Y. Wu, and H. Fu, “Amplified Spontaneous Emission Based on 2D Ruddlesden–Popper Perovskites,” Adv. Funct. Mater. 28(17), 1707006 (2018). [CrossRef]  

13. Y. Jiang, J. Wei, and M. Yuan, “Energy-Funneling Process in Quasi-2D Perovskite Light-Emitting Diodes,” J. Phys. Chem. Lett. 12(10), 2593–2606 (2021). [CrossRef]  

14. L. Lei, D. Seyitliyev, S. Stuard, J. Mendes, Q. Dong, X. Fu, Y.-A. Chen, S. He, X. Yi, L. Zhu, C.-H. Chang, H. Ade, K. Gundogdu, and F. So, “Efficient Energy Funneling in Quasi-2D Perovskites: From Light Emission to Lasing,” Adv. Mater. 32(16), 1906571 (2020). [CrossRef]  

15. Z. Guo, Y. Zhang, B. Wang, L. Wang, N. Zhou, Z. Qiu, N. Li, Y. Chen, C. Zhu, H. Xie, T. Song, L. Song, H. Xue, S. Tao, Q. Chen, G. Xing, L. Xiao, Z. Liu, and H. Zhou, “Promoting Energy Transfer via Manipulation of Crystallization Kinetics of Quasi-2D Perovskites for Efficient Green Light-Emitting Diodes,” Adv. Mater. 33(40), 2102246 (2021). [CrossRef]  

16. C. Ni, Y. Huang, T. Zeng, D. Chen, H. Chen, M. Wei, A. Johnston, A. H. Proppe, Z. Ning, E. H. Sargent, P. Hu, and Z. Yang, “Thiophene Cation Intercalation to Improve Band-Edge Integrity in Reduced-Dimensional Perovskites,” Angew. Chem. Int. Ed. 59(33), 13977–13983 (2020). [CrossRef]  

17. T. Zhou, H. Lai, T. Liu, D. Lu, X. Wan, X. Zhang, Y. Liu, and Y. Chen, “Highly Efficient and Stable Solar Cells Based on Crystalline Oriented 2D/3D Hybrid Perovskite,” Adv. Mater. 31(32), 1901242 (2019). [CrossRef]  

18. Y. Li, I. Allegro, M. Kaiser, A. J. Malla, B. S. Richards, U. Lemmer, U. W. Paetzold, and I. A. Howard, “Exciton versus free carrier emission: Implications for photoluminescence efficiency and amplified spontaneous emission thresholds in quasi-2D and 3D perovskites,” Mater. Today 49, 35–47 (2021). [CrossRef]  

19. D. Ma, K. Lin, Y. Dong, H. Choubisa, A. H. Proppe, D. Wu, Y.-K. Wang, B. Chen, P. Li, J. Z. Fan, F. Yuan, A. Johnston, Y. Liu, Y. Kang, Z.-H. Lu, Z. Wei, and E. H. Sargent, “Distribution control enables efficient reduced-dimensional perovskite LEDs,” Nature 599(7886), 594–598 (2021). [CrossRef]  

20. P. Pang, G. Jin, C. Liang, B. Wang, W. Xiang, D. Zhang, J. Xu, W. Hong, Z. Xiao, L. Wang, G. Xing, J. Chen, and D. Ma, “Rearranging Low-Dimensional Phase Distribution of Quasi-2D Perovskites for Efficient Sky-Blue Perovskite Light-Emitting Diodes,” ACS Nano 14(9), 11420–11430 (2020). [CrossRef]  

21. Z. Ren, L. Li, J. Yu, R. Ma, X. Xiao, R. Chen, K. Wang, X. W. Sun, W.-J. Yin, and W. C. H. Choy, “Simultaneous Low-Order Phase Suppression and Defect Passivation for Efficient and Stable Blue Light-Emitting Diodes,” ACS Energy Lett. 5(8), 2569–2579 (2020). [CrossRef]  

22. L. Kong, X. Zhang, Y. Li, H. Wang, Y. Jiang, S. Wang, M. You, C. Zhang, T. Zhang, S. V. Kershaw, W. Zheng, Y. Yang, Q. Lin, M. Yuan, A. L. Rogach, and X. Yang, “Smoothing the energy transfer pathway in quasi-2D perovskite films using methanesulfonate leads to highly efficient light-emitting devices,” Nat. Commun. 12(1), 1246 (2021). [CrossRef]  

23. X. Gong, O. Voznyy, A. Jain, W. Liu, R. Sabatini, Z. Piontkowski, G. Walters, G. Bappi, S. Nokhrin, O. Bushuyev, M. Yuan, R. Comin, D. McCamant, S. O. Kelley, and E. H. Sargent, “Electron–phonon interaction in efficient perovskite blue emitters,” Nat. Mater. 17(6), 550–556 (2018). [CrossRef]  

24. Q. Du, C. Zhu, Z. Yin, G. Na, C. Cheng, Y. Han, N. Liu, X. Niu, H. Zhou, H. Chen, L. Zhang, S. Jin, and Q. Chen, “Stacking Effects on Electron-Phonon Coupling in Layered Hybrid Perovskites via Microstrain Manipulation,” ACS Nano 14(5), 5806–5817 (2020). [CrossRef]  

25. W. Cai, M. U. Ali, P. Liu, M. He, C. Zhao, Z. Chen, Y. Zang, M.-C. Tang, H. Meng, H. Fu, G. Wei, and H.-L. Yip, “Unravelling Alkali-Metal-Assisted Domain Distribution of Quasi-2D Perovskites for Cascade Energy Transfer toward Efficient Blue Light-Emitting Diodes,” Adv. Sci. 9(20), 2200393 (2022). [CrossRef]  

26. Y. Jiang, M. Cui, S. Li, C. Sun, Y. Huang, J. Wei, L. Zhang, M. Lv, C. Qin, Y. Liu, and M. Yuan, “Reducing the impact of Auger recombination in quasi-2D perovskite light-emitting diodes,” Nat. Commun. 12(1), 336 (2021). [CrossRef]  

27. M. Cui, C. Qin, Y. Jiang, M. Yuan, L. Xu, D. Song, Y. Jiang, and Y. Liu, “Direct Observation of Competition between Amplified Spontaneous Emission and Auger Recombination in Quasi-Two-Dimensional Perovskites,” J. Phys. Chem. Lett. 11(14), 5734–5740 (2020). [CrossRef]  

28. C. Wang, G. Dai, J. Wang, M. Cui, Y. Yang, S. Yang, C. Qin, S. Chang, K. Wu, Y. Liu, and H. Zhong, “Low-Threshold Blue Quasi-2D Perovskite Laser through Domain Distribution Control,” Nano Lett. 22(3), 1338–1344 (2022). [CrossRef]  

29. D. Zhang, Y. Fu, C. Liu, C. Zhao, X. Gao, J. Zhang, W. Guo, J. Liu, C. Qin, and L. Wang, “Domain Controlling by Compound Additive toward Highly Efficient Quasi-2D Perovskite Light-Emitting Diodes,” Adv. Funct. Mater. 31(38), 2103890 (2021). [CrossRef]  

30. A. Caiazzo, K. Datta, J. Jiang, M. C. Gélvez-Rueda, J. Li, R. Ollearo, J. M. Vicent-Luna, S. Tao, F. C. Grozema, M. M. Wienk, and R. A. J. Janssen, “Effect of Co-Solvents on the Crystallization and Phase Distribution of Mixed-Dimensional Perovskites,” Adv. Energy Mater. 11(42), 2102144 (2021). [CrossRef]  

31. Y. Liang, Q. Shang, Q. Wei, L. Zhao, Z. Liu, J. Shi, Y. Zhong, J. Chen, Y. Gao, M. Li, X. Liu, G. Xing, and Q. Zhang, “Lasing from Mechanically Exfoliated 2D Homologous Ruddlesden–Popper Perovskite Engineered by Inorganic Layer Thickness,” Adv. Mater. 31(39), 1903030 (2019). [CrossRef]  

32. M. Li, Q. Shang, C. Li, S. Li, Y. Liang, W. Yu, C. Wu, L. Zhao, Y. Zhong, W. Du, X. Wu, Z. Jia, Y. Gao, H. Chen, X. Liu, S. Guo, Q. Liao, G. Xing, L. Xiao, and Q. Zhang, “High Optical Gain of Solution-Processed Mixed-Cation CsPbBr3 Thin Films towards Enhanced Amplified Spontaneous Emission,” Adv. Funct. Mater. 31(25), 2102210 (2021). [CrossRef]  

33. Y. Liang, Q. Shang, M. Li, S. Zhang, X. Liu, and Q. Zhang, “Solvent Recrystallization-Enabled Green Amplified Spontaneous Emissions with an Ultra-Low Threshold from Pinhole-Free Perovskite Films,” Adv. Funct. Mater. 31(48), 2106108 (2021). [CrossRef]  

34. J. Li, L. Zhang, Z. Chu, C. Dong, J. Jiang, Z. Yin, J. You, J. Wu, W. Lan, and X. Zhang, “Amplified Spontaneous Emission with a Low Threshold from Quasi-2D Perovskite Films via Phase Engineering and Surface Passivation,” Adv. Opt. Mater. 10(6), 2102563 (2022). [CrossRef]  

35. F. Thouin, D. A. Valverde-Chávez, C. Quarti, D. Cortecchia, I. Bargigia, D. Beljonne, A. Petrozza, C. Silva, and A. R. Srimath Kandada, “Phonon coherences reveal the polaronic character of excitons in two-dimensional lead halide perovskites,” Nat. Mater. 18(4), 349–356 (2019). [CrossRef]  

36. C. Qin, W. Liu, N. Liu, Z. Zhou, J. Song, S.-H. Ma, Z.-Y. Jiao, and S. Lei, “Observation of Hole Transfer in MoS2/WS2 Van der Waals Heterostructures,” ACS Photonics 9(5), 1709–1716 (2022). [CrossRef]  

37. S. Panuganti, L. V. Besteiro, E. S. Vasileiadou, J. M. Hoffman, A. O. Govorov, S. K. Gray, M. G. Kanatzidis, and R. D. Schaller, “Distance Dependence of Förster Resonance Energy Transfer Rates in 2D Perovskite Quantum Wells via Control of Organic Spacer Length,” J. Am. Chem. Soc. 143(11), 4244–4252 (2021). [CrossRef]  

38. K. Bruening and C. J. Tassone, “Antisolvent processing of lead halide perovskite thin films studied by in situ X-ray diffraction,” J. Mater. Chem. A 6(39), 18865–18870 (2018). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental Information

Data availability

The data underlying the results presented herein are not publicly available currently but can be obtained from the authors upon reasonable request.

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. Structure and optical properties of (ThMA)2Csn-1PbnBr3n+1 with different compositions. (a) Schematics for (ThMA)2Csn-1PbnBr3n+1 (n = 1, 2, 3, and ∞). (b) XRD patterns for (ThMA)2Csn-1PbnBr3n+1 films. (c) UV-vis absorption (solid line) and PL spectra (dashed line) for (ThMA)2Csn-1PbnBr3n+1 films.
Fig. 2.
Fig. 2. ASE and TA spectra for quasi-2D (ThMA)2Cs2Pb3Br10 films (pristine films). (a) Pump-fluence dependence of PL characteristics under 400 nm excitation. (b) The dependence of FWHM and PL intensity on the pump fluence, respectively. (c) TA spectra for 0.19, 0.52, 1 and 2 ps. (d) Normalized GSB curves at 404, 432 and 508 nm (320 nm excitation, 45 µJ/cm2) corresponding to n = 1, n = 2 and n ≥ 5 QWs, respectively.
Fig. 3.
Fig. 3. Pump-fluence dependence of the ASE and the TA for (ThMA)2Cs2Pb3Br10 films (with CB). (a) Pump-fluence dependence of PL under 400 nm excitation. (b) The dependence of FWHM and PL intensity as a function of pump fluence, respectively. (c) TA spectra for 0.18, 0.52, 1 and 2 ps (inset: the maximum intensity of different n-phase CBS signals). (d) Normalized GSB kinetics curves at 432 and 508 nm for <n> = 3 films (320 nm excitation, 45 µJ/cm2) corresponding to n = 2 and n ≥ 5 QWs, respectively.
Fig. 4.
Fig. 4. The effects of anti-solvent for (ThMA)2Cs2Pb3Br10 films. (a) Upper panel: UV-vis absorption spectra. Low panel: normalized PL spectra (dashed line: the normalized PL spectra of these films in logarithmic). The PL peak shifts from 518 to 513 nm after CB treatment. (b) XRD patterns.
Fig. 5.
Fig. 5. The phase distribution and the threshold caused by anti-solvent CB for (ThMA)2Cs2Pb3Br10 films. (a) the integrated intensities of the GBS peak of the pristine films (cyan-blue) and the films with CB (purple) for n = 1, 2, 3, and ∞. (b) The Pth and η of (ThMA)2Cs2Pb3Br10 films with the various dripping time of the CB.
Fig. 6.
Fig. 6. Lasing properties. (a) Schematic diagram of the vertical-cavity surface-emitting laser. (b) Evolution of PL spectra under various pump fluences. Inset: The dependence of PL intensity as a function of pump fluence.

Equations (1)

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

η  = ( Δ OD MAX, n = 2 + Δ OD MAX, n = 3 ) / β
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.