Nanowire Lasers of Formamidinium Lead Halide Perovskites and

Jan 4, 2016 - Department of Chemistry, Columbia University, New York, New York 10027, United States ... Yang ZhouJie ChenOsman M. BakrHong-Tao Sun...
5 downloads 0 Views 7MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Letter pubs.acs.org/NanoLett

Nanowire Lasers of Formamidinium Lead Halide Perovskites and Their Stabilized Alloys with Improved Stability Yongping Fu,† Haiming Zhu,‡ Alex W. Schrader,† Dong Liang,† Qi Ding,† Prakriti Joshi,‡ Leekyoung Hwang,† X-Y. Zhu,*,‡ and Song Jin*,† †

Department of Chemistry, University of Wisconsin-Madison, 1101 University Avenue, Madison, Wisconsin 53706, United States Department of Chemistry, Columbia University, New York, New York 10027, United States



S Supporting Information *

ABSTRACT: The excellent intrinsic optoelectronic properties of methylammonium lead halide perovskites (MAPbX3, X = Br, I), such as high photoluminescence quantum efficiency, long carrier lifetime, and high gain coupled with the facile solution growth of nanowires make them promising new materials for ultralow-threshold nanowire lasers. However, their photo and thermal stabilities need to be improved for practical applications. Herein, we report a low-temperature solution growth of single crystal nanowires of formamidinium lead halide perovskites (FAPbX3) that feature red-shifted emission and better thermal stability compared to MAPbX3. We demonstrate optically pumped room-temperature near-infrared (∼820 nm) and green lasing (∼560 nm) from FAPbI3 (and MABr-stabilized FAPbI3) and FAPbBr3 nanowires with low lasing thresholds of several microjoules per square centimeter and high quality factors of about 1500−2300. More remarkably, the FAPbI3 and MABr-stabilized FAPbI3 nanowires display durable roomtemperature lasing under ∼108 shots of sustained illumination of 402 nm pulsed laser excitation (150 fs, 250 kHz), substantially exceeding the stability of MAPbI3 (∼107 laser shots). We further demonstrate tunable nanowire lasers in wider wavelength region from FA-based lead halide perovskite alloys (FA,MA)PbI3 and (FA,MA)Pb(I,Br)3 through cation and anion substitutions. The results suggest that formamidinium lead halide perovskite nanostructures could be more promising and stable materials for the development of light-emitting diodes and continuous-wave lasers. KEYWORDS: Formamidinium lead trihalide perovskite, nanowires, semiconductor nanowire lasers, photostability, tunable lasers and optoelectronic applications.12,13 The past five years have witnessed tremendous advances of highly efficient perovskite solar cells that recently achieved a certified power conversion efficiency of 20.1%.14−20 In parallel with this rapid development in solar conversion efficiency, the high photoluminescence (PL) quantum efficiency suggests lead halide perovskite should also be a good light emitter.21,22 Owing to the large absorption coefficient, long carrier lifetimes and high gain, efficient lasing or amplified spontaneous emission from thin films21−24 or nanostructures25−29 of methylammonium lead halide perovskite (MAPbX3) and quantum dots of cesium lead halide perovskite (CsPbX3)30 has already been demonstrated. In particular, solution-grown single-crystal MAPbX3 NWs showed lasing thresholds as low as a few hundred nanojoules per centimeter squared and a high estimated PL quantum efficiency of ∼87% at room temperature, making them an outstanding new member among semiconductor NW lasers.2,25 Such high-lasing performance can be attributed to the remarkably long carrier

S

emiconductor nanowires (NWs) are promising candidates for the realization of miniaturized lasers in advanced photonic circuits.1−4 Each NW provides both gain medium (that it is made of) and cavity for a laser. The one-dimensional (1D) geometry and large refractive index allow strong confinement of photonic modes guided along the axial direction, and the two smooth end facets serve as a Fabry− Perot cavity for optical amplification to trigger lasing action.5,6 Over the past ∼15 years, optically pumped lasing has been demonstrated from a number of classic inorganic semiconductor NWs, such as ZnO, GaN, CdS, and GaAs,3,5,7−9 with emissions from the ultraviolet to near-infrared region. However, these inorganic NWs are usually grown via vapor phase deposition processes that require high temperatures. Moreover, NW lasers in the near-infrared (NIR) spectral region mainly based on GaAs are particularly interesting for optical communications, but it is difficult to realize efficient lasing in this material due to large surface recombination velocity at room temperature.5,10,11 To achieve room-temperature lasing, additional core−shell design to passivate the surface is required. Recently, organic−inorganic lead halide perovskites have emerged as a new class of earth-abundant semiconductor materials that have exceptional promise for both photovoltaic © XXXX American Chemical Society

Received: October 5, 2015 Revised: December 19, 2015

A

DOI: 10.1021/acs.nanolett.5b04053 Nano Lett. XXXX, XXX, XXX−XXX

Letter

Nano Letters

Figure 1. Material characterizations of FAPbI3 and MABr-stabilized FAPbI3 perovskite nanostructures. (a) The schematic crystal structure of lead halide perovskites APbX3, A = FA or MA, X = I, Br, or Cl. (b) Thermogravimetric curves of FAPbI3 and MAPbI3 powders under a N2 or O2 atmosphere with a heating rate of 2 °C/min from room temperature to ∼350 °C. (c) SEM image of as-converted perovskite phase FAPbI3 NWs. The inset is a top view of a NW, showing a hexagonal cross-section and flat end facet. (d) Low-resolution and (e) high-resolution TEM image of a converted perovskite FAPbI3 NW; the inset in panel e is the corresponding FFT pattern along the [001] ZA of the pseudocubic structure. (f) PXRD patterns of as-prepared hexagonal phase FAPbI3, perovskite phase FAPbI3, and perovskite phase MABr-stabilized FAPbI3. (g) SEM image of asgrown MABr-stabilized FAPbI3 perovskite nanowires. The inset is a top view of a NW, showing a rectangular cross-section and flat end facet. (h) EDS mapping of a MABr-stabilized FAPbI3 NW, showing the uniform elemental distribution of Pb, I, and Br. (i) 1H NMR spectra of redissolved MABr-stabilized FAPbI3 in comparison with free MAI and FAI reagents in methanol-d4 solvent, confirming the alloying of MA and FA in MABrstabilized FAPbI3.

temperature that might be achieved in an operating optoelectronic device, and the phase transition could further modify the emission proprieties of MAPbI3.27 Three-dimensional (3D) organic−inorganic hybrid perovskites generally adopt the formula of ABX3 (see Figure 1a for crystal structure) in which A is an organic cation, B is a metal ion (such as Pb2+, Sn2+, and other metal ions), and X is a halide anion. A major advantage of this family for optoelectronic applications is the wide compositional substitution toward A, B, and X sites for the tailoring of optical and physical properties.41,42 For example, exchanging the organic cation of the MAPbI3 perovskite from methylammonium to formamidinium (CH(NH2)2+, FA) leads to a semiconductor with a slightly lower bandgap of 1.47 eV, as well as better temperature and moisture stability.40,43 To compare the thermal stability of FAPbI3 and MAPbI3 quantitatively, we carried out thermogra-

lifetimes and low trap densities in MAPbX3, particularly in single crystal forms.31,32 Considering the long-range balanced ambipolar charge properties of MAPbX333,34 and recent realization of the bright light-emitting diodes (LEDs),35,36 these perovskite NWs promise the possibility to realize electrically driven lasing.4 Despite such excellent performance achieved in MAPbX3 for optically pumped lasers and LEDs, one major obstacle for practical applications is its instability that stems from poor thermal stability and high moisture sensitivity.13 In particular, the thermal stability may impose technical limits for the realization of continuous-wave (CW) pumped lasers37 or electrically driven lasers,4 because the thermal heating from prolonged CW pumping or high current injection could potentially cause the degradation of perovskite materials. Moreover, MAPbI3 undergoes a reversible phase transition between tetragonal and cubic phase at ∼57 °C,38−40 a B

DOI: 10.1021/acs.nanolett.5b04053 Nano Lett. XXXX, XXX, XXX−XXX

Letter

Nano Letters

preserved (Figure 1c and the inset). Further TEM analysis reveals that the phase transition is not single crystal to single crystal for all NWs (Supporting Information Figure S2a−c). Most NWs after conversion were polycrystalline with singlecrystal domains inside. We suspect that the NWs with small diameter could be more likely to preserve single crystalline nature after conversion based on our previous observation in other NWs.49 Figure 1d,e shows the low-resolution and highresolution TEM image together with the corresponding fast Fourier transform (FFT) pattern (inset) of an example of the converted NW. The sharp diffraction spots in the FFT pattern can be indexed to the pseudocubic perovskite structure with a zone axis (ZA) of [001]. Quantitative elemental analysis from energy dispersive X-ray spectroscopy (EDS) on individual NW revealed a I/Pb ratio of ∼3, which is in good agreement with the stoichiometry of FAPbI3 (Supporting Information Figure S2d,e). However, we found some of the black FAPbI3 gradually turned back to yellow FAPbI3 over a few weeks at room temperature even in a desiccator. The phase stability issue still needs to be addressed before it can be practically used in an optoelectronic device. It is well-known that the match of radii of A, B, and X ions plays a critical role in the formation and stability of perovskite structures. A tolerance factor [α = (rA + rX)/√2(rB + rX), where r is the effective ion radius] has been used to evaluate the ionic size mismatches that the perovskite structure can tolerate until a different type of crystal structure is formed.41,50 Simultaneous and synergic compositional modification of A and/or X sites may lead to the stabilization of perovskite structure of FAPbI3 at room temperature.19,40 Recently, Jeon et al. mixed FAPbI3 thin film with MAPbBr3 to stabilize the perovskite phase of FAPbI3 to achieve the highest perovskite solar cell efficiency of 20.1%.19,20 We further discovered that stabilized single-crystal NWs of perovskite phase FAPbI3 alloys can be directly grown by the addition of a small amount of MABr into the FAI solution. Figure 1g shows the SEM images of MABr-stabilized FAPbI3 NWs grown using a mixed solution of 20 mg/mL FAI and 5 mg/mL MABr. The length of these NWs varies from a few micrometers to ∼10 μm, while the width of most NWs is around several hundred nanometers (see Supporting Information Figure S3 for more SEM images). We employed a “seeding growth” method in order to grow the NWs in higher density (Supporting Information Figure S4). The inset in Figure 1g highlights the rectangular cross section of these NWs, which are in contrast to the hexagonal ones seen in Figure 1c for the converted FAPbI3 NWs. PXRD (Figure 1f, blue curve) shows a group of strong diffraction peaks at 13.95, 19.80, 24.33, and 28.17° that can be assigned to (100), (110), (111), and (200) lattice planes of the cubic perovskite phase, confirming the perovskite structure of as-grown products. Energy-dispersive X-ray spectroscopy mapping of a single MABr-stabilized FAPbI3 NW shows uniform spatial distribution of Pb, I, and Br elements (Figure 1h). Quantitative elemental analysis of EDS yields a I/Br ratio ∼2.7/0.3. We further determined the ratio of FA/MA to be ∼0.75/0.25 by using 1H NMR spectroscopy (Figure 1i, see Supporting Information for experimental details). A representative confocal PL spectrum of a single MABr-stabilized FAPbI3 NW at room temperature shows an emission peak centered at 786 nm, which is slightly blue shifted from that of FAPbI3 NWs due to the incorporation of MABr (Supporting Information Figure S5). We measured optically pumped lasing on the converted perovskite phase FAPbI3 NWs on a home-built inverted

vimetric analysis (TGA) in a N2 or O2 atmosphere (Figure 1b), which shows the onset of decomposition temperature of FAPbI3 is significantly higher than that of MAPbI3, especially in the presence of O2. The better thermal stability of FAPbI3 can be attributed to the enhanced hydrogen bonding between FA cations and the iodide ions of Pb−I octahedral.40,44 Because of the enhanced stability under elevated temperature and the smaller bandgap, FAPbI3 perovskite is receiving increasing attention in the photovoltaic research community,19,20,43,45,46 yet there have been no reports of light-emitting devices such as LEDs and lasers from FAPbI3 or other FA-based perovskites with the exception for the recent demonstration of LED based on FAPbBr3.47 In this Letter, we show low-temperature solution growth of high-quality single-crystal hexagonal phase FAPbI3 NWs and their conversion to perovskite phase NWs and also a direct solution-growth of stabilized FAPbI3 alloy NWs in the cubic perovskite phase by incorporating a small amount of MABr into FAPbI3. We demonstrate optically pumped lasing from the FAPbI3 perovskite NWs and MABr-stabilized FAPbI3 perovskite NWs at room temperature with near-infrared (NIR) emission at ∼800 nm, low lasing thresholds of a few microjoules per square centimeter and high quality factors of ∼1500. More importantly, both types of NWs show significantly improved lasing stability than MAPbI3 NW due to the enhanced thermal stability. In addition, we also report the solution growth of NWs of FAPbBr3, mixed cation alloys (FA,MA)PbI3, and double alloys (FA,MA)Pb(Br,I)3 through both cations and halides substitutions. Owing to the better photostability and cation-induced bandgap tuning effect, another advantage of these NWs of FA-based perovskites over MA-based perovskites is that the lasing emissions are more widely tunable from visible to near-infrared wavelength. Following our understanding of MAPbI3 nanostructure growth,25,48 we successfully synthesized single-crystal FAPbI3 NWs by reacting a lead acetate thin film deposited on glass slide immersed with a FAI solution in isopropanol (see Supporting Information for experimental details). However, an important difference is that the originally formed product of FAPbI3 is in the hexagonal phase (yellow color, nonperovskite structure) by powder X-ray diffraction (PXRD, Figure 1f, and black curve). Unlike MAPbI3, the larger radius of FA cation favors the formation of a more stable hexagonal phase instead of perovskite structure at room temperature. Supporting Information Figure S1a−c shows the scanning electron microscopy (SEM) images of hexagonal FAPbI3 NWs grown on a glass substrate at 50 °C for ∼20 h. These FAPbI3 NWs typically have lengths from several to tens of micrometers with flat hexagonal end facets. The diameter varies from a few hundred nanometers to several micrometers. Transmission electron microscopy (TEM) was further used to confirm the single-crystal nature of the as-grown NWs and the growth direction along the c-axis (Supporting Information Figure S1d,e). We also found these NWs were more stable under TEM electron beam than MAPbI3 NWs. The hexagonal phase FAPbI3 is an indirect bandgap semiconductor with a nonperovskite type structure, which is not suitable for photovoltaic and light-emitting applications.39,40 However, the yellow products could be turned into the black perovskite phase by heating in air at 170 °C for 5 min. The corresponding PXRD pattern (Figure 1f, red curve) confirmed the products after conversion were trigonal phase (a perovskite-type structure) and the NW morphology and hexagonal cross section were C

DOI: 10.1021/acs.nanolett.5b04053 Nano Lett. XXXX, XXX, XXX−XXX

Letter

Nano Letters

Figure 2. Near-infrared lasing from converted perovskite phase FAPbI3 NWs. (a) The 2D pseudocolor plot of the emission spectra under different pump fluences (P). Note the intensity is in logarithm scale. (b) NW emission spectra around the lasing threshold. Inset: Integrated emission intensity and fwhm of emission peak as a function of P showing the lasing threshold ∼6.2 μJ cm−2. (c) Time-resolved PL decay kinetics of NW below (blue circles) and above (red triangles) lasing threshold. Also shown in gray is the instrument response function (IRF). (d) Optical image (upper panel) and fluorescence images of a single NW below (middle panel) and above (lower panel) lasing threshold (scale bar, 10 μm).

Because the cross-sectional dimensions of the NWs are smaller than light wavelength, the coherent lasing emission from the two ends of a NW diffract into nearly spherical distributions that form interference pattern on fluorescence image (Figure 2d, lower).52 Among all the converted FAPbI3 NWs examined (23 in total with length ranging from 6 to 30 μm), we observed either single or multiple lasing modes from the NWs (Supporting Information Figure S9a−c). For multiple mode lasing with equally spaced peaks, the mode spacing decreases with NW length. The plot of mode spacing versus reciprocal NW length conforms to a straight line intersecting the origin (Supporting Information Figure S9d), confirming Fabry−Perot longitudinal cavity modes. The lasing threshold of these FAPbI3 NWs varies from NW to NW from 6 to 30 μJ cm−2 without clear dependence on NW length (Supporting Information Figure S10a). The limited spatial resolution of our microscope setup (∼1.3 μm) prevents us from directly obtaining the width of a specific NW measured. Instead, we estimated the width of NWs for lasing measurements (ranging from 0.5 to 1 μm) by a deconvolution process (see Supporting Information Figure S11 for an example). We noticed these perovskite phase FAPbI3 NWs generally show higher lasing thresholds than our previously reported MAPbI3 NWs,25 which could be due to its worse crystalline quality, rough surface, and imperfect end facets after thermal conversion. Note that the refractive index can also impact the lasing thresholds. Unfortunately, there have been no reports on the refractive index of FA-based perovskites so far, but we suspect the values might be close to those of MAbased perovskites (n = ∼2.5 at lasing wavength)53 due to their similar crystal structure and composition. We also characterized the lasing performance of MABrstabilized FAPbI3 perovskite NWs (Figure 3). Similar to

microscope system with 402 nm pulsed laser excitation (see Supporting Information for experimental details).25 The NW samples were dispersed on to quartz window of a N2 gas filled cell by dry contact transfer and kept in nitrogen atmosphere during measurements and excited evenly by 402 nm pulsed laser excitation. Figure 2a shows the 2D pseudocolor plot of PL spectra of a representative FAPbI3 NW (with a length of ∼11 μm and a width of ∼0.7 μm Figure 2d) with pump fluence between 4.1 and 7.8 μJ cm−2 and Figure 2b shows four PL spectra around lasing threshold. The inset in Figure 2b shows the integrated PL intensity and the fwhm of PL peak as a function of pump fluence. Below the lasing threshold (PTh) ∼ 6.2 μJ cm−2, FAPbI3 NW shows a broad PL spectra centered at ∼813 nm and a full width at half-maximum (fwhm) ∼ 40 nm (see Supporting Information Figure S6 for full range PL spectra) and the integrated PL intensity grows linearly with pump fluence (Figure 2b, inset). Above PTh, a sharp lasing peak at ∼824 nm emerges and the slope of the emission intensity versus pump fluence plot increases (Figure 2b), while the intensity of the spontaneous emission region approaches saturation (Supporting Information Figure S7), indicating the occurrence of lasing. The log−log plot further shows the expected S-shaped behavior for stimulated emission51 (Supporting Information Figure S8). The fwhm of the lasing peak (δλ) is ∼0.53 nm, corresponding to a quality factor (Q = λ/δλ) of ∼1554. Further evidence for lasing comes from the timeresolved PL decay kinetics (Figure 2c) and fluorescence images below and above lasing threshold (Figure 2d). Below PTh, the FAPbI3 NW shows a PL lifetime ∼800 ps (Figure 2c, blue symbols) and a uniform spontaneous emission image (Figure 2d middle). Above PTh, an instrument-limited ultrafast PL decay component ( x > 0.2, where the emission peaks gradually changed under continuous laser illumination (see Supporting Information Figures S17 and S18). In a previous work, a reversible, light-induced transformation of PL spectra in MAPbBr3−xIx films has been observed.55 It was proposed that photoexcitation may cause phase segregation between Br-rich and Br-poor phases in MAPbBr3−xIx films due to light-induced ion migration.55,56 Interestingly, we found that the FA-based perovskite double alloys display much better photostability compared to MAbased alloys under the same laser excitation condition (see Supporting Information Figures S19 and S20). We point out that it is difficult to synthesize NWs of FAPb(Br,I)3 alloys from a mixed FABr and FAI solution but adding a small amount of MA can promote the NW growth. We grow NWs of perovskite double alloys of (MA,FA)Pb(Br,I)3 by reacting PbAc2 film with a mixed solution of FABr (fixed at 7 mg/mL) and MAI in isopropanol. Figure 6a shows the SEM image of an example of single-crystal double alloy NWs grown with a mixed solution of 7 mg/mL FABr and 3 mg/mL MAI. The PXRD pattern confirms the cubic perovskite phase of these NWs. EDS mapping analysis on a single NW (Figure 6b) and 1H NMR spectra (Figure 6c) further determine a stoichiometry of (FA0.67MA0.33)Pb(Br2.69I0.31). The PL peak of the (MA,FA)Pb(Br,I)3 NWs can continuously red shift until 620 nm (Supporting Information Figure S20) upon increasing the MAI concentration in the mixed precursor solution from 1 to 3 mg/mL. A series of optical images of the (MA,FA)Pb(Br,I)3 NWs with increasing MAI content excited by a 442 nm CW laser (Figure 6d) clearly demonstrate colorful emissions and strong waveguiding effect along the NW axis. We further observed room-temperature lasing from (FA0.71MA0.29)Pb(Br2.78I0.22) and (FA0.67MA0.33)Pb(Br2.69I0.31) NWs, where the lasing peaks are at 595 and 621 nm, respectively. The 2D pseudocolor plot of the emission spectra from these NWs as a function of pump fluence are shown in Supporting Information Figure S21. With these new NW lasers based on the double alloys of (MA,FA)Pb(Br,I)3, we can fill in the gap of lasing wavelength previously unavailable with MA-based perovskites. There can be two factors for the improved photostability of these FA-based double alloys over MA-based alloys. First, the FA-based perovskites have better thermal stability (Figure 1b), which could further contribute to improved photostability under sustained laser illumination. Second, recent firstprinciples calculations showed the activation energies of the ion migration in FAPbI3 are slightly higher than those in MAPbI3.57 It is possible that the light-induced ion migration effect55 might be mitigated in FA-based perovskites due to the larger cation size and the enhanced interaction between FA cations and halide ions. In addition, we also demonstrated efficient lasing from (FA0.48MA0.54)PbI3 NW alloy with a lasing peak at 805 nm. The detailed structure characterizations and

Figure 5. Structure and optical properties of single-crystal FAPbBr3 NWs. (a) SEM image of as-grown FAPbBr3 NWs. The inset shows a rectangular cross-section and flat end facet of a NW. (b) Emission spectra of a NW (with a length of 12 μm) below and above lasing threshold. The insets are optical images of the NW below and above lasing threshold.

FAPbBr3 NWs with flat rectangular end facet (see Supporting Information Figure S14 for more SEM images). Unlike the FAPbI3 with two structural polymorphs, FAPbBr3 exists only as a single perovskite phase at room temperature. The corresponding PXRD (Supporting Information Figure S15, red curve) shows a set of strong diffraction peaks at 14.69, 20.88, 29.72, and 33.33° that could be well assigned to the (100), (110), (111), and (120) planes of pseudocubic structure (space group Pm3̅m). Compared to MAPbBr3, these diffraction peaks shift to smaller angles due to the larger size of FA cation (Supporting Information Figure S15). The cation-induced tuning of optical bandgap leads to ∼10 nm red shift of PL peak when moving from MA to FA cation in lead bromide perovskite.44,54 We also observed efficient lasing in the green spectral region from FAPbBr3 NWs at room temperature. Figure 5b shows a broad spontaneous emission spectra centered at ∼548 nm with a fwhm ∼23 nm below PTh and narrow lasing peaks at ∼560 nm with fwhm ∼0.24 nm above PTh (corresponding to a quality factor of ∼2300). The inset in F

DOI: 10.1021/acs.nanolett.5b04053 Nano Lett. XXXX, XXX, XXX−XXX

Letter

Nano Letters

Figure 6. Wavelength-tunable lasing from NWs of cation and anion alloys of perovskites. (a) SEM image of as-grown NWs of double alloys using (FA0.67MA0.33)Pb(Br2.69I0.31) as an example. (b) EDS mapping of a (FA,MA)Pb(Br,I)3 NW, showing the uniform elemental distribution of Pb, I, and Br. (f) 1H NMR spectrum confirms the alloying of MA and FA in (FA,MA)Pb(Br,I)3. (d) Optical images of a series of (FAxMA1−x)Pb(Br3−yIy) NWs excited by a 442 nm laser, showing colorful emission and strong waveguiding effect along the NW axis. (e) Broad wavelength-tunable lasing from single-crystal lead perovskite NWs. The rectangular boxes highlight the new wavelength range of emissions achieved by cation alloying (MA,FA)PbI3 NWs or both cation and anion alloying in (FA,MA)Pb(Br,I)3 NWs, which could not be realized in MA-based perovskite alloys. The data outside of the boxes were adapted from ref 25.

hybrid perovskite materials with different cations and anions41,42,50 to exploit their diverse physical properties.

lasing results of NWs of cation-mixed (FA,MA)PbI3 alloys can be found in Supporting Information Figures S22−24. Therefore, it is clear that the alloying of both cation and anion widens the wavelength tunability of lead halide perovskite nanowire lasers to achieve continuously tunable lasing wavelength from 490 to 824 nm (Figure 6e). In summary, we have developed the solution synthesis of high-quality single-crystal NWs of FAPbI3, MABr-stabilized FAPbI3, FAPbBr3, (FA,MA)PbI3 alloys, and (FA,MA)Pb(Br,I)3 double alloys for the first time. We further show roomtemperature lasing in the visible and NIR spectral region from these NWs with a low lasing threshold (∼several μJ cm−2) and high quality factor (∼2000). These FA-based perovskite NWs display much better photostability and wider wavelength tunability over MA-based perovskite NWs. These results suggest the FA-based pervoskites could be more promising and stable candidates for the future development of lightemitting diodes and lasers based on perovskite materials. These results also demonstrate the generality of the solution synthesis of nanostructures for various families of organic−inorganic



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.nanolett.5b04053. Experimental methods, additional SEM images, PXRD patterns, and optical characterizations of various products. (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. Author Contributions

Y.F. and H.Z. contributed equally to this work. G

DOI: 10.1021/acs.nanolett.5b04053 Nano Lett. XXXX, XXX, XXX−XXX

Letter

Nano Letters Notes

(23) Stranks, S. D.; Wood, S. M.; Wojciechowski, K.; Deschler, F.; Saliba, M.; Khandelwal, H.; Patel, J. B.; Elston, S.; Herz, L. M.; Johnston, M. B. Nano Lett. 2015, 15, 4935−4941. (24) Suárez, I.; Juárez-Pérez, E. J.; Bisquert, J.; Mora-Seró, I.; Martínez-Pastor, J. P. Adv. Mater. 2015, 27, 6157. (25) Zhu, H.; Fu, Y.; Meng, F.; Wu, X.; Gong, Z.; Ding, Q.; Gustafsson, M. V.; Trinh, M. T.; Jin, S.; Zhu, X. Nat. Mater. 2015, 14, 636−642. (26) Zhang, Q.; Ha, S. T.; Liu, X.; Sum, T. C.; Xiong, Q. Nano Lett. 2014, 14, 5995−6001. (27) Xing, J.; Liu, X.; Zhang, Q.; Ha, S. T.; Yuan, Y.; Shen, C.; Sum, T. C.; Xiong, Q. Nano Lett. 2015, 15, 4571−4577. (28) Sutherland, B. R.; Hoogland, S.; Adachi, M. M.; Wong, C. T.; Sargent, E. H. ACS Nano 2014, 8, 10947−10952. (29) Liao, Q.; Hu, K.; Zhang, H.; Wang, X.; Yao, J.; Fu, H. Adv. Mater. 2015, 27, 3405−3410. (30) Yakunin, S.; Protesescu, L.; Krieg, F.; Bodnarchuk, M. I.; Nedelcu, G.; Humer, M.; De Luca, G.; Fiebig, M.; Heiss, W.; Kovalenko, M. V. Nat. Commun. 2015, 6, 8056. (31) Dong, Q.; Fang, Y.; Shao, Y.; Mulligan, P.; Qiu, J.; Cao, L.; Huang, J. Science 2015, 347, 967−70. (32) Shi, D.; Adinolfi, V.; Comin, R.; Yuan, M. J.; Alarousu, E.; Buin, A.; Chen, Y.; Hoogland, S.; Rothenberger, A.; Katsiev, K.; Losovyj, Y.; Zhang, X.; Dowben, P. A.; Mohammed, O. F.; Sargent, E. H.; Bakr, O. M. Science 2015, 347, 519−522. (33) Stranks, S. D.; Eperon, G. E.; Grancini, G.; Menelaou, C.; Alcocer, M. J.; Leijtens, T.; Herz, L. M.; Petrozza, A.; Snaith, H. J. Science 2013, 342, 341−344. (34) Xing, G. C.; Mathews, N.; Sun, S. Y.; Lim, S. S.; Lam, Y. M.; Gratzel, M.; Mhaisalkar, S.; Sum, T. C. Science 2013, 342, 344−347. (35) Tan, Z. K.; Moghaddam, R. S.; Lai, M. L.; Docampo, P.; Higler, R.; Deschler, F.; Price, M.; Sadhanala, A.; Pazos, L. M.; Credgington, D.; Hanusch, F.; Bein, T.; Snaith, H. J.; Friend, R. H. Nat. Nanotechnol. 2014, 9, 687−692. (36) Wong, A. B.; Lai, M.; Eaton, S. W.; Yu, Y.; Lin, E.; Dou, L.; Fu, A.; Yang, P. Nano Lett. 2015, 15, 5519−5524. (37) Cadelano, M.; Sarritzu, V.; Sestu, N.; Marongiu, D.; Chen, F.; Piras, R.; Corpino, R.; Carbonaro, C. M.; Quochi, F.; Saba, M. Adv. Opt. Mater. 2015, 3, 1557. (38) Kawamura, Y.; Mashiyama, H.; Hasebe, K. J. Phys. Soc. Jpn. 2002, 71, 1694−1697. (39) Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G. Inorg. Chem. 2013, 52, 9019−9038. (40) Binek, A.; Hanusch, F. C.; Docampo, P.; Bein, T. J. Phys. Chem. Lett. 2015, 6, 1249−1253. (41) Kieslich, G.; Sun, S.; Cheetham, A. K. Chem. Sci. 2014, 5, 4712− 4715. (42) Boix, P. P.; Agarwala, S.; Koh, T. M.; Mathews, N.; Mhaisalkar, S. G. J. Phys. Chem. Lett. 2015, 6, 898−907. (43) Eperon, G. E.; Stranks, S. D.; Menelaou, C.; Johnston, M. B.; Herz, L. M.; Snaith, H. J. Energy Environ. Sci. 2014, 7, 982−988. (44) Amat, A.; Mosconi, E.; Ronca, E.; Quarti, C.; Umari, P.; Nazeeruddin, M. K.; Grätzel, M.; De Angelis, F. Nano Lett. 2014, 14, 3608−3616. (45) Koh, T. M.; Fu, K.; Fang, Y.; Chen, S.; Sum, T.; Mathews, N.; Mhaisalkar, S. G.; Boix, P. P.; Baikie, T. J. Phys. Chem. C 2014, 118, 16458−16462. (46) Pellet, N.; Gao, P.; Gregori, G.; Yang, T. Y.; Nazeeruddin, M. K.; Maier, J.; Gratzel, M. Angew. Chem., Int. Ed. 2014, 53, 3151−3157. (47) Aygüler, M. F.; Weber, M. D.; Puscher, B. M.; Medina, D. D.; Docampo, P.; Costa, R. D. J. Phys. Chem. C 2015, 119, 12047−12054. (48) Fu, Y.; Meng, F.; Rowley, M. B.; Thompson, B. J.; Shearer, M. J.; Ma, D.; Hamers, R. J.; Wright, J. C.; Jin, S. J. Am. Chem. Soc. 2015, 137, 5810−5818. (49) Li, L.; Yu, Y.; Meng, F.; Tan, Y.; Hamers, R. J.; Jin, S. Nano Lett. 2012, 12, 724−731. (50) Kieslich, G.; Sun, S.; Cheetham, A. K. Chem. Sci. 2015, 6, 3430− 3433.

The authors declare no competing financial interest.



ACKNOWLEDGMENTS S.J. acknowledges support by the Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering, under Award DE-FG02-09ER46664. X.Y.Z acknowledges support by the Department of Energy under Grant DE-SC0010692-02 for lasing and photophysical measurements. H.Z. is a research fellow in the Columbia University Materials Research Science and Engineering Center, supported by NSF grant DMR-1420634. A.W.S. thanks support of his REU by a generous gift from the University of WisconsinMadison Graduate School and National Science Foundation through the University of Wisconsin-Madison Materials Research Science and Engineering Center (DMR-0520527) and Nanoscale Science and Engineering Center (DMR0425880).



REFERENCES

(1) Yan, R.; Gargas, D.; Yang, P. Nat. Photonics 2009, 3, 569−576. (2) Fu, A.; Yang, P. Nat. Mater. 2015, 14, 557−558. (3) Huang, M. H.; Mao, S.; Feick, H.; Yan, H.; Wu, Y.; Kind, H.; Weber, E.; Russo, R.; Yang, P. Science 2001, 292, 1897−1899. (4) Duan, X.; Huang, Y.; Agarwal, R.; Lieber, C. M. Nature 2003, 421, 241−245. (5) Saxena, D.; Mokkapati, S.; Parkinson, P.; Jiang, N.; Gao, Q.; Tan, H. H.; Jagadish, C. Nat. Photonics 2013, 7, 963−968. (6) Ning, C. Phys. Status Solidi B 2010, 247, 774−788. (7) Johnson, J. C.; Choi, H.-J.; Knutsen, K. P.; Schaller, R. D.; Yang, P.; Saykally, R. J. Nat. Mater. 2002, 1, 106−110. (8) Agarwal, R.; Barrelet, C. J.; Lieber, C. M. Nano Lett. 2005, 5, 917−920. (9) Fan, F.; Turkdogan, S.; Liu, Z.; Shelhammer, D.; Ning, C. Nat. Nanotechnol. 2015, 10, 796−803. (10) Hua, B.; Motohisa, J.; Kobayashi, Y.; Hara, S.; Fukui, T. Nano Lett. 2009, 9, 112−116. (11) Mayer, B.; Rudolph, D.; Schnell, J.; Morkötter, S.; Winnerl, J.; Treu, J.; Müller, K.; Bracher, G.; Abstreiter, G.; Koblmüller, G.; Finley, J. J. Nat. Commun. 2013, 4, 2931. (12) Stranks, S. D.; Snaith, H. J. Nat. Nanotechnol. 2015, 10, 391− 402. (13) Sum, T. C.; Chen, S.; Xing, G.; Liu, X.; Wu, B. Nanotechnology 2015, 26, 342001. (14) Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. J. Am. Chem. Soc. 2009, 131, 6050−6051. (15) Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J. E. Sci. Rep. 2012, 2, 591. (16) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J. Science 2012, 338, 643−647. (17) Zhou, H. P.; Chen, Q.; Li, G.; Luo, S.; Song, T. B.; Duan, H. S.; Hong, Z. R.; You, J. B.; Liu, Y. S.; Yang, Y. Science 2014, 345, 542− 546. (18) Burschka, J.; Pellet, N.; Moon, S. J.; Humphry-Baker, R.; Gao, P.; Nazeeruddin, M. K.; Gratzel, M. Nature 2013, 499, 316−319. (19) Jeon, N. J.; Noh, J. H.; Yang, W. S.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I. Nature 2015, 517, 476−80. (20) Yang, W. S.; Noh, J. H.; Jeon, N. J.; Kim, Y. C.; Ryu, S.; Seo, J.; Seok, S. I. Science 2015, 348, 1234−1237. (21) Deschler, F.; Price, M.; Pathak, S.; Klintberg, L. E.; Jarausch, D. D.; Higler, R.; Huttner, S.; Leijtens, T.; Stranks, S. D.; Snaith, H. J.; Atature, M.; Phillips, R. T.; Friend, R. H. J. Phys. Chem. Lett. 2014, 5, 1421−1426. (22) Xing, G. C.; Mathews, N.; Lim, S. S.; Yantara, N.; Liu, X. F.; Sabba, D.; Gratzel, M.; Mhaisalkar, S.; Sum, T. C. Nat. Mater. 2014, 13, 476−480. H

DOI: 10.1021/acs.nanolett.5b04053 Nano Lett. XXXX, XXX, XXX−XXX

Letter

Nano Letters (51) Zimmler, M. A.; Bao, J.; Capasso, F.; Müller, S.; Ronning, C. Appl. Phys. Lett. 2008, 93, 051101. (52) Van Vugt, L. K.; Rühle, S.; Vanmaekelbergh, D. Nano Lett. 2006, 6, 2707−2711. (53) Green, M. A.; Jiang, Y.; Soufiani, A. M.; Ho-Baillie, A. J. Phys. Chem. Lett. 2015, 6, 4774−4785. (54) Hanusch, F. C.; Wiesenmayer, E.; Mankel, E.; Binek, A.; Angloher, P.; Fraunhofer, C.; Giesbrecht, N.; Feckl, J. M.; Jaegermann, W.; Johrendt, D. J. Phys. Chem. Lett. 2014, 5, 2791−2795. (55) Hoke, E. T.; Slotcavage, D. J.; Dohner, E. R.; Bowring, A. R.; Karunadasa, H. I.; McGehee, M. D. Chem. Sci. 2015, 6, 613−617. (56) Sadhanala, A.; Deschler, F.; Thomas, T. H.; Dutton, S. n. E.; Goedel, K. C.; Hanusch, F. C.; Lai, M. L.; Steiner, U.; Bein, T.; Docampo, P. J. Phys. Chem. Lett. 2014, 5, 2501−2505. (57) Haruyama, J.; Sodeyama, K.; Han, L.; Tateyama, Y. J. Am. Chem. Soc. 2015, 137, 10048−10051.

I

DOI: 10.1021/acs.nanolett.5b04053 Nano Lett. XXXX, XXX, XXX−XXX