Open access peer-reviewed chapter

Single Crystal Hybrid Perovskite Optoelectronics: Progress and Perspectives

Written By

Feng Li

Submitted: 15 September 2020 Reviewed: 17 November 2020 Published: 09 December 2020

DOI: 10.5772/intechopen.95046

From the Edited Volume

Optoelectronics

Edited by Mike Haidar Shahine

Chapter metrics overview

731 Chapter Downloads

View Full Metrics

Abstract

Organic–inorganic hybrid perovskites, which combine the superior optical and electronic properties and solution-processed manufacturing, have emerged as a new class of revolutionary optoelectronic devices with the potential for various practical applications. Encouraged by the advantages of longer carrier diffusion length, higher carrier mobility and lower trap densities as compared to the polycrystalline counterparts’, increasing research attention has focused on preparation and optimization of perovskite crystal candidates, via using various facile growth techniques, for the development of a wide range of optoelectronic applications. This chapter presents a comprehensive review of recent advances in the field of optoelectronic technologies based on different forms of single crystals, including bulk crystals and thin ones, with emphasis placed on the optimization of crystals and the relationship among the charge-carrier transport, operation mechanism, device architecture, and device performance. First, we introduce the main methods used to prepare bulk and thin single crystals, and analyze several aspects of their properties. Thereafter, the applications of single crystals into solar cells, photodetectors, light-emitting diodes, and lasers, are discussed in depth. Finally, we summarize the challenges of perovskite single crystals and propose further improvements in the synthesis approaches and device applications.

Keywords

  • hybrid perovskite crystals
  • optoelectronics
  • solar cells
  • photodetectors
  • LEDs

1. Introduction

Perovskite-structured materials have received increasing attention, since being discovered in the 1830s, because of their rich physical properties [1]. As shown in Figure 1a [2], the general chemical formula for such compounds is ABX3, in which A and B are different cations, and X is an anion that bonds to both the A and B cations. Owing to the flexibility of bond angles inherent in the perovskite structure, there are many different distortions that can occur from the ideal structure. Importantly, A can be organic cations, like methylammonium (MA+) or formamidinium (FA+) [4, 5, 6, 7, 8], B can be metal ions, such as Pb2+ and Sn2+ [9, 10, 11, 12], and X is usually halide ions [13], and such a class of materials is known as organic–inorganic hybrid perovskites. It was reported that a stable structure of hybrid perovskites can form where 0.81 < T.F. (tolerance factor) < 1.1 and 0.44 < O.F. (octahedral factor) < 0.90 [14]. X-ray diffraction (XRD) measurements were widely used to characterize their structures. As for MAPbBr3 and MAPbI3 crystals, XRD measurements displayed the excellent single crystal properties [15]. Transmission electron microscopy (TEM) measurements were performed to provide a more intuitive picture of perovskite crystals structures ( Figure 1b ), via using contrast-transfer-function corrected method to overcome their electron beam-sensitive property [3]. After the first attempt to employ hybrid perovskite films as active sensitizers into photovoltaic devices [16], hybrid perovskite solar cells have continued to set new efficiency benchmarks [17, 18, 19, 20, 21, 22, 23], due to the excellent properties, such as ease of processing, tunable optical band gaps [24, 25], long carrier diffusion length [26], and low trap density [15], as well as large absorption coefficients and high photoluminescence (PL) efficiency [27, 28], and their relatively high power conversion efficiency (PCE) has been increased to as high as 25.2% [29]. Moreover, Leveraging their promising features, hybrid perovskites also have the potential for employment in other optoelectronic applications, including photodetectors [30], transistors [31], phototransistors [32], light-emitting diodes (LEDs) [33], and lasers [34].

Figure 1.

1a, perovskite crystal structure. Nature Photonics [2], copyright 2014. 2b, CTF-corrected high-revolution TEM image. Science [3], copyright 2018.

However, a vast array of prior research on perovskite optoelectronic devices has been centered on polycrystalline films. The polycrystalline samples usually suffer from grain boundaries, relatively higher trap densities and defects, and low stability, which would obviously obscure their potential in applications [35, 36, 37]. More recently, researchers have paid more attention to perovskite single crystals, which possess promising characteristics of no grain boundaries [15], relatively low trap density [38], large charge carrier mobility, and long carrier diffusion length [39, 40, 41]. In this regard, extensive efforts are being devoted to developing effective methods to improve the perovskite crystal quality and optimize the device performance. Existing in the forms of bulk or thin crystals, perovskite crystal samples have been widely applied in various optoelectronic applications [39, 42], and have made rapid and great strides in research progress [43, 44, 45, 46].

In this chapter, we aim to summarize the recent achievements, ongoing progress, and the challenges to date in the area of hybrid perovskite single crystals, practically MA-based ones (MAPbX3, X = Cl, Br, and I), from the perspective of both materials and devices with an emphasis placed on the optimization of crystal quality, and provide an outlook on the opportunities offered by this emerging family of materials in field of optoelectronic applications.

Advertisement

2. Growth of hybrid perovskite single crystals

2.1 Bulk single crystals

2.1.1 Solution temperature-lowering (STL) method

According to the lower solubility of MAPbX3 in HX (X = Cl, Br, and I) solution as the temperature decreases, Tao’s group introduced the STL method to synthesize a MAPbI3 bulk single crystal ( Figure 2a ) [47]. After the reaction between methylamine (CH3NH2) and hydro-iodic acid (HI) in a cold atmosphere, the obtained white microcrystal MAI was reacted with Pb(CH3COOH)2∙3H2O in aqueous HI, and the solution was then cooled to 40°C. A 10 mm × 10 mm × 8 mm black MAPbI3 single crystal was grown in about one month ( Figure 2b ). Lin’s group discovered a more efficient way, and they synthesized the single crystals with a size of 5 mm in just around 10 days [48]. Lin et al. selected high-quality seeds and dropped them back into fresh solution and obtained single crystals sized up to 1 cm ( Figure 2c ). Furthermore, MAPbBr3 − xClx and MAPbI3 − xBrx mixed-halide perovskite crystals were studied using such method [49]. Hydro-bromic acid with hydrochloric acid or hydro-iodic acid were mixed in different molar ratios into methylamine and lead (II) acetate solution to fabricate single-halide and mixed-halide perovskite crystals ( Figure 2d ). The time-consuming factor is the biggest drawback of this method, which has indirectly led to the domination of other crystallization methods.

Figure 2.

2a, schematic of STL method. 2b, image of MAPbI3 with {100} and {112} facets. CrystEngComm [47], copyright 2015. 2c, MAPbBr3 crystals from STL method. J. Cryst. Growth [48], copyright 2015. 2d, photographs of perovskite crystals with different halide ratio. Nature Photonics [49], copyright 2015. 2e, MAPbI3 and MAPbBr3 crystals growth at different time intervals. Nature Commun. [50], Copyright 2015. 2f, schematic of crystals growth. J. Mater. Chem. C [51], copyright 2016. 2 g, schematic of AVC method. Science [15], copyright 2015.

2.1.2 Inverse temperature crystallization (ITC) method

As a radically faster perovskite crystal synthesis approach, the ITC method has widely been applied in recent years. It was observed that the exhibited crystals from such method can be shape-controlled, higher quality, and obtained quicker compared with other growth techniques. Bakr et al. introduced this method to rapidly grow high-quality bulk crystals [50]. As shown in Figure 2e , an orange MAPbBr3 crystal and a black MAPbI3 crystal were grown within 3 hours. Chen’s group further studied the effect of molar ratio of MAX and PbX2 in the precursor solutions on the crystal quality [52], e.g., perovskite crystals with different sizes and shapes were obtained after a 6-hour ITC crystallization process when changing the MAX: PbX2 ratios from 1:1 to 2:1.

With an aim of growing a large-sized bulk perovskite crystal, such ITC method was further modified. Using such technique, the strategy of incorporation of seed crystal growth has been proven to be favorable for single crystals as large as convenient. Liu’s group reported various large-sized perovskite crystals via using the modified ITC method, from which a number of larger-sized crystal (7 mm) were obtained through choosing good-quality seed crystals and repeating and carefully controlling the ITC process several times ( Figure 2f ). Moreover, Liu’s group also successfully grew MAPb(BrxI1 − x)3 single crystals with a finely-tuned bandgap [51]. The application of the different solubility of different perovskite single crystals at varying temperatures contributes to the time-saving feature of such ITC method.

2.1.3 Anti-solvent vapor-assisted crystallization (AVC) method

Another main method to grow perovskite crystals is the AVC method ( Figure 2g ), which was first introduced from Bakr’s group [15]. In this method, the solvent plays a significant role because two or more solvents should be selected, of which one should be a good solvent that is less volatile, and the other is a bad solvent that is more volatile. The principle of this method can be described as follows: when the bad solvent slowly diffuses into the precursor solution, the proficiency of the crystal formation increases at the bottom of the sample vial owing to the insolubility of the material in the bad solvent. Other groups, like Loi’s group and Cao’s group, also applied this method to obtain the high-quality crystals [38, 53]. Although the AVC method costs more time than the ITC method, its temperature-irrelevant characteristic is appealing to its widespread use.

2.2 Thin single crystals

Bulk perovskite single crystals with thick sizes may cause the increase of charge recombination, which would lead to the degradation of their device performance and impede the practical applications. In this regard, growing thin perovskite crystals with a large area represents an effective approach to overcome the above obstacle and thus advances the further practical applications. Bakr et al. introduced a cavitation-triggered asymmetrical crystallization strategy, in which a very short ultrasonic pulse (≈1 s) was applied in the solution to reach a low supersaturation level with anti-solvent vapor diffusion and a thin crystal with several-micrometers grew on the substrates within hours ( Figure 3a ) [54]. Liu’s group synthesized perovskite crystal wafers with a much thinner thickness using a dynamic flow micro-reactor system [55]. They put two thin glass slides in parallel into a container with a predefined separation to grow single crystals within the slit channel, as shown in Figure 3b . Su’s group further used a space-limited ITC method and grew a 120-cm2 single crystal on fluorine-doped tin oxide (FTO)-coated glass, of which the operation and the obtained 0.4-mm-thin single crystal are shown in Figure 3c [56]. Meanwhile, Wan et al. reported a space-confined solution-processed method to grow the perovskite single-crystalline films with adjustable thickness from nanometers to micrometers ( Figure 3d ) [57]. Benefitting from the capillary pressure, the perovskite precursor solution filled the whole space between two clean flat substrates, which were clipped together and dipped in the solution.

Figure 3.

3a, schematic of cavitation-triggered asymmetrical method. Adv. Mater. [54], Copyright 2016. 3b, schematic of ultrathin crystal wafer growth. Adv. Mater. [55], Copyright 2016. 3c, schematic of the laminar MAPbBr3 crystal films preparation. Adv. Mater. [56], Copyright 2017. 3d, schematic for the growth of perovskite thin crystals. J. Am. Chem. Soc. [57], copyright 2016. 3e, schematic of geometrically-confined lateral crystal growth method. Nature Commun. [58], Copyright 2017. 3f, schematic of the scalable growth for perovskite crystal films using an inkjet printing method. Sci. Adv. [59], Copyright 2018.

Currently, more promising approaches have been employed to grow thin single crystals with high quality and large scale. A one-step printing geometrically-confined lateral crystal growth method ( Figure 3e ) was introduced by Sung’s group to obtain a large-scaled single crystal [58]. During the process, a cylindrical metal roller with a flexible poly-(dimethyl-siloxane) (PDMS) mold was wrapped and then rolled on a preheated SiO2 substrate (180°C) with an ink supplier filled with the precursor solution. Alternatively, millimeter-sized single crystals were synthesized by Song’s group by a facile seed-inkjet-printing approach ( Figure 3f ) [59]. Perovskite precursor solution was injected onto a silicon wafer, and then the ordered seeds were formed on the substrate with the evaporation of the droplets. Thereafter, the substrate with a saturated perovskite solution was covered and the single crystals can be grew as the solvent dried at room temperature. Seeds were used to inhibit the random nucleation and trigger the growth of single crystals.

As discussed above, some optimized space-limited approaches have been introduced and developed to synthesize perovskite thin crystals in recent years. Especially, size−/thickness-controlled thin crystals have also been widely used in various optoelectronic devices. With the aim to growing large-scaled and thickness-controlled thin crystals with longer carrier diffusion lengths, fewer defects, and higher efficiency, more promising strategies will be rewarding in the future.

Advertisement

3. Optoelectronic characterizations of perovskite single crystals

3.1 Optical properties

There are two normal ways to study the optical properties of hybrid perovskite crystals: absorption and PL measurements. Bakr et al. characterized the steady-state absorption and PL properties for MAPbBr3 and MAPbI3 crystals, as shown in Figure 4a and b [50]. Sharp band edges were observed in the absorption plots and the band gap values were determined to be 2.18 eV for MAPbBr3 crystals and 1.51 eV for MAPbI3 crystals; while the PL intensity peaks are located at 574 nm for MAPbBr3 and 820 nm for MAPbI3. As for the MAPbCl3 one, absorption measurement result revealed an edge at 435 nm ( Figure 4c ) [60]. Clearly, the optical absorption of perovskite crystals exhibited a clear-cut sharp band edge, which indicated that the single crystals are predominantly free from grain boundaries and have relatively low structural defects and trap densities.

Figure 4.

Steady-state absorption (4a) and PL spectra (4b) of MAPbBr3 and MAPbI3 crystals, respectively. Nature Commun. [50], Copyright 2015. 4c, steady-state absorption and PL spectra of MAPbCl3 crystal. Insets: Band gap for the above single crystals. J. Phys. Chem. Lett. [60], Copyright 2015, 4d, normalized PL decays for MAPbBr3 film (red) and crystal (blue) excited at 447 nm. Inset shows the zoom on the shorter time scale. Nature Commun. [61], Copyright 2017.

More recently, there have been more broad publications on the apparent disparity in optical properties (i.e., absorption and PL) between perovskite single crystals and thin films, which can be attributed to the incorrect measurements as a result of reabsorption effects. Snaith’s group performed a detailed investigation of the optical properties of MAPbBr3 crystals as compared to those of the polycrystalline films by employing light transmission spectroscopy, ellipsometry, and spatially resolved and time-resolved PL spectroscopy [61]. They showed that the optical properties of the perovskite crystals were almost identical to those of polycrystalline films, and their observations indicated that the perovskite polycrystalline films were much closer to possessing ‘single-crystal-like’ optoelectronic properties than previously thought, and also highlighted the discrepancies in the estimation of trap densities from the electronic and optical methods ( Figure 4d ). For the further development of perovskite crystals, more detailed experimental investigations combined with theoretical calculations that focus on the optical features are required, which would assist in the preparation of the high-quality perovskite single crystals and the development of the high-performance device applications.

3.2 Charge transport properties

As for hybrid perovskite crystals, in addition to the remarkable optical properties, their promising electrical properties have caught the great attention. In general, there are five common methods to measure the transport mobilities in perovskite crystals, including the space charge limited current (SCLC), time-of-flight (TOF), Hall Effect, THz pulse and field-effect transistor (FET) measurement methods. Among these methods, the SCLC method is widely employed to determine the carrier mobility and trap density of perovskite crystals. The current–voltage (I-V) curve can be divided into three parts: the first region, where an Ohmic contact exists, hence the conductivity can be estimated; the second region is the trap-filling region, which is increased sharply at trap-filled limit voltage (V TFL); and the third region, known as the child region. Trap density (n trap) can be obtained by following the relation: n trap = (2V TFL εε 0)/(eL 2), where ε 0 is the vacuum permittivity, ε is the relative dielectric constant, L is the crystal thickness, and e is the electron charge. Moreover, the mobility (μ) is determined by fitting the I-V curve with Mott-Gurney’s law: μ = (8JL 3)/(9εε 0 V 2), where J is the current density. Liu’s group designed the hole-only device ( Figure 5a ), and a large hole mobility of 67.27 cm2/Vs was estimated [62]. An SCLC method was also applied on MAPbBr3 crystals, with an n trap of 5.8 × 109 cm−3 and a μ of 38 cm2/Vs [15]. I-V response of a MAPbCl3 crystal was measured by Bakr’s group with n trap = 3.1 × 1010 cm−3 and μ = 42 cm2/Vs [60]. Another method to measure the μ is the TOF method. Bakr’s group obtained the μ via using the TOF method ( Figure 5b ) [15], from which μ can be defined by the equation: μ = d 2/( t), where d is the sample thickness, V is the applied voltage, and τ t is the transit time that be provided by the transient current under different driving voltages [67, 68]. The same method was also applied by Huang’s group and the electron μ was verified to be 24.0 ± 6.8 cm2/Vs ( Figure 5c ) [63]. Apart from the above two methods, Bakr et al. also carried out the complementary Hall Effect measurements on perovskite crystals, confirming the μ ranging from 20 to 60 cm2/Vs [15]. Meanwhile, Huang’s group applies the Hall Effect measurement [68], and they showed the crystals were slightly p-doped with a low free holes concentration. Thereafter, Podzorov’s group increased the conductivity of MAPbBr3 single crystals by sputtering Ti on the flat-faceted single crystal to form Hall bars ( Figure 5d ) [64], from which the Hall mobility was calculated to be 10 cm2/Vs.

Figure 5.

5a, dark I-V curve of hole-only MAPbI3 crystal device. J. Energy Chem. [62], Copyright 2018. 5b, ToF traces of MAPbBr3 crystal. Science [15], copyright 2015. 5c, transient current curves of perovskite crystal devices. Science [63], copyright 2015. 5d, schematic of hall effect measurement. Adv. Mater. [64], Copyright 2016. 5e, schematic of time-resolved multi-THz spectroscopy experiment. 5f, incident (black), transmitted (blue) and reflected (red) multi-THz pulses after interaction with the crystal. Energy Environ. Sci. [65], Copyright 2015. 5 g, schematic of bottom-gate, top-contact perovskite crystal FET. 5 h, transfer characteristics of a MAPbCl3 device. Nature Commun. [66], Copyright 2018. 5i, PL time decay trace of a MAPbBr3 crystal. Science [15], copyright 2015.

Although the above measurement approaches have been widely used in the perovskite crystals, the obtained results from different groups are sometimes different. Sargent et al. demonstrated that one main challenge that may explain these order-of-magnitude discrepancies is that the Hall Effect, TOF, and SCLC methods all probe the mobilities near the respective Fermi levels during the experiments, and the (non-equilibrium, high-injection-level) Fermi level is widely different in each experiment [64]. In this regard, they developed a contactless method to measure the mobility of a perovskite crystal directly [64]. Plus, THz pulse measurement was also used to estimate μ. David et al. used a two-color laser plasma in dry air to generate multi-THz pulses and excited the large MAPbI3 single crystals and detected the electric field by an air-biased coherent detection scheme with 1–30 THz ultra-bandwidth after normal incidence reflection off the crystal facet ( Figure 5e , f ) [65]. Such spectra measurements indicate the ultrafast dynamics and efficiencies of free charge creation and remarkably high μ as high as 500–800 cm2/Vs. Furthermore, FETs are the fundamental components to realize digital integrated circuits, which are also often used as a platform to evaluate charge transport mechanism in the active materials. In this regard, bottom-gate, top-contact FETs were fabricated via using micrometer-thin MAPbX3 (X = Cl, Br, and I) crystals as active layer ( Figure 4g )[66], from which the field-effect μ values are up to 4.7 and 1.5 cm2/Vs in p- and n-channel devices, respectively ( Figure 5h ).

Carrier lifetime (τ) is an important parameter that should be considered when designing an optoelectronic device. Upon excitation by photons, the active materials will be in an excited state. After that, the photo-induced holes and electrons will recombine back to the ground state. Usually, if this recombination process, that is, the carrier lifetime of carriers, is sufficiently long, a high performance device will be expected. The τ of semiconductors strongly depends on the nature, dimension, and purity of the materials. Generally, τ can be obtained from the PL decay, transient absorption, as well as the transient photo-voltage decay and impedance methods [69]. Among these methods, the PL decay approach has been widely applied. The superposition of fast and slow components of carrier dynamics from the PL spectra measurement result yield τ ≈ 41 and 357 ns for MAPbBr3 ( Figure 5i ) [15, 70, 71]. Transient absorption (TA) also suggests the recombination property of excitons which is used to determine the carrier lifetime through a bi-exponential fitting [60]. The carrier diffusion length L D can be further estimated based on the equation: L D = [((k BT)/eμτ)]1/2, where k B is Boltzmann’s constant and T is the sample temperature. From the above-examined values of μ and τ, L D was calculated [63, 64].

Advertisement

4. Applications of perovskite single crystals

4.1 Photovoltaic cells

The widely studied hybrid perovskite solar cells with high performance are usually made from polycrystalline films; however, the current studies have also focused on the developments and optimization of single crystal perovskite solar cells, owing to their significant advantages. Huang et al. fabricated photovoltaic devices based on MAPbI3 bulk crystals by depositing gold (Au) as anodes and gallium (Ga) as cathodes ( Figure 6a ) [63]. A red-shift of 50 nm of the EQE cutoff to 850 nm showed that MAPbI3 crystals increased the upper limit of short-circuit current density (JSC) compared with the polycrystalline solar cells from 27.5 mA/cm2 to 33.0 mA/cm2. Notably, as compared with the perovskite polycrystalline solar cells, the bulk crystal devices showed much lower efficiency, which was attributed to the fact that photo-generated carriers could not be fully collected in a thick active layer. Much thinner MAPbBr3 monocrystalline films grown on indium tin oxide (ITO)-coated glass were applied into the solar cells, and the devices showed the best cell performance with a fill factor (FF) of 0.58, a JSC of 7.42 mA/cm2, an open-circuit voltage (VOC) of 1.24 V, and a PCE of 5.37% ( Figure 6b ) [54]. To enhance the device performance, Huang’s group further fabricated crystal solar cells through interface engineering ( Figure 6c ), of which the best device showed a JSC of 20.5 mA/cm2, a VOC of 1.06 V, a FF of 74.1%, and a PCE of 16.1% [72]. The single crystal solar cell also displayed the better device stability of remaining nearly unchanged after storage in air for 30 days.

Figure 6.

6a, schematic of MAPbI3 crystal solar cell. Science [63], copyright 2015. 6b, dark and illuminated J-V curves of MAPbBr3 crystal solar cells with a device illustration in the inset. Adv. Mater. [54], Copyright 2016. 6c, device structure of single-crystal solar cells. Nature Commun. [72], Copyright 2017. 6d, schematic of MAPbI3 crystal solar cells with lateral structure. Adv. Mater. [73], Copyright 2016. 6e, cross-sectional SEM image of a MAPbI3 crystal device. 6f, statistical summary of photovoltaic parameters from 12 devices. ACS Energy Lett. [74], Copyright 2019.

In addition to the vertical-structured solar cells, Huang’s group also fabricated the lateral structure perovskite crystal device ( Figure 6d ) [73], which showed a VOC of 0.82 V and the highest PCE of 5.36% at 170 K. More recently, a 20-μm MAPbI3 single crystal inverted p-i-n solar cell with a PCE as high as 21.09% and a FF up to 84.3% was fabricated [74], of which the cross-sectional SEM image and photovoltaic performance are shown in Figure 6e and f . To further realize the optimized performance of perovskite crystal solar cells, more efforts will be performed to enhance the sample quality and to design promising device structures.

4.2 Photodetectors

Photodetectors which can convert incident light into electrical signals are critical for various industrial and scientific applications. To evaluate the photodetector performance, several parameters are important, including responsivity (R), detectivity (D*), Gain (G), and linear dynamic range (LDR), which are listed and are explained in Table 1 briefly.

Quantity Unit Definition
Photocurrent (I light) A Current through a photodetector resulting from illumination.
Dark-current (I dark) A Current through a device in the absence of illumination.
Photoresponsivity (R) A/W R is calculated according to: R = (I lightI dark)/P light, where P light is power of the incident light.
Detectivity (D *) Jones D * can be calculated as R/(2eJ d)1/2, where e is elementary charge and J d is dark current density.
Gain (G) G can be calculated as [(J lightJ dark)/e]/(P light/), where is the incident photon energy.
Linear dynamic range (LDR) dB LDR is calculated by LDR = 20log(P sat/P low), where P sat (P low) is the light intensity when the incident light intensity stronger (weaker) than which the photocurrent begins to deviate from linearity.
External quantum efficiency (EQE) % Carrier number divided by the number of incident photons.
Internal quantum efficiency (IQE) % It is the ratio of carrier number to the number of incident photons that are absorbed by the device.

Table 1.

Parameters for evaluating the perovskite single crystal photodetectors.

4.2.1 In visible region

Huang’s group fabricated perovskite crystal photodetectors that exhibited a high sensitivity capacity, which led to a narrow-band photo-response with a full width at half maximum (FWHM) of less than 20 nm under V = −1 V ( Figure 7a ) [49]. EQE spectra of the single crystals showed a narrow peak near the absorption edge, which promised a detection application at a specific wavelength, with a peak D* over 2 × 1010 Jones at 570 nm under V = −4 V ( Figure 7b ). Also, Huang et al. further fabricated vertical structured perovskite crystal photodetectors by using the non-wetting hole transport layer-coating substrates [75]. The noise currents are as low as 1.4 and 1.8 fA/Hz1/2 at an 8-Hz frequency for the devices based on MAPbBr3and MAPbI3, respectively. Additionally, the photocurrent responses of both the MAPbBr3 and MAPbI3 devices were linear, and their LDRs are up to 256 and 222 dB, respectively. Sun’s group introduced a planar-type photodetector on the MAPbI3 crystal (001) facet with a highest R value of 953 A/W and EQE of 2.22 × 105% at a light power density of 2.12 nW/cm2 [76]. Wei’s group used a two-step method to fabricate a self-powered photodetector based on a MAPbBr3 crystal core-shell heterojunction [77]. The device showed a broad photo-response ranging from 350 to 800 nm and a peak R up to 11.5 mA/W. Hu’s group fabricated photodetectors based on MAPbI3 single crystal nanowires and nanoplates by transferring them to SiO2/Si slides [78]. The highest On/Off ratio approached 103 under a light illumination of 73.7 mW/cm2.

Figure 7.

7a, schematic of device structure. 7b, D* spectrum and total noise at −4 V. Nature Photonics [49], copyright 2015. 7c, illustration of planar-integrated MAPbBr3 photodetector. Nature Commun. [42], Copyright 2015. Photograph of ≈100 photodetectors on a perovskite crystal wafer (7d) and the R values (7e). Adv. Mater. [55], Copyright 2016.

Although perovskite crystal photodetectors have shown better performance, macroscopic crystals cannot be grown on a planar substrate, restricting their potential for device integration. To overcome this shortcoming, Bakr et al. grew large-area planar-integrated crystal films onto the ITO-patterned substrates ( Figure 7c ) [42], and the fabricated photodetector possessed a high G (above 104) and a high gain-bandwidth product (above 108 Hz) relative to other perovskite devices. Furthermore, Liu’s group fabricated a photodetector based on a thin perovskite crystal wafer by the space-limited crystallization method, which has about 100 pairs of interdigitated Au wire electrodes ( Figure 7d ) [55], and the R increased linearly as the radiance intensity decreased ( Figure 7e ). Moreover, Su’s group sputtered the thin Au electrodes on a large-area MAPbBr3 thin crystal to fabricate a narrowband photodetector [56]. Furthermore, Ma’s group reported the superior-performance photodetectors based on MAPbBr3 thin crystals [79], which displayed the R as high as 1.6 × 107 A/W and the highest G up to 5 × 107.

4.2.2 In ultraviolet (UV) region

UV detection is a key technology in the fields of flame detection [80], remote security monitoring [81], environmental monitoring [82], and so forth. Researchers have endeavored to develop UV photodetectors based on perovskite crystals considering their excellent UV absorption properties. Visible-blind UV photodetectors based on MAPbCl3 crystals a suitable bandgap of about 3.11 eV were fabricated ( Figure 8a ) [60], and the device showed the dark current as low as 4.15 × 10−7 A at 15 V and a drastically high stability ( Figure 8b ). Planar-integrated MAPbCl3 crystal UV photodetectors on ITO-deposited glass substrate were reported by Sargent et al. ( Figure 8c ) [83], which showed decreased R and G values as increased power density of a 385-nm laser ( Figure 8d ) [85].

Figure 8.

8a, device architecture of MAPbCl3 crystal photodetector. 8b, I-V curves of the photodetector under UV light (λ = 365 nm) and in the dark. J. Phys. Chem. Lett. [60], Copyright 2015. 8c, schematic of planar-integrated MAPbCl3 UV-detectors. 8d, R and G values vs. incident light power. Adv. Mater. [83], Copyright 2016. 8e, sub-gap electron trap state absorptions. 8f, R values of MAPbI3 photo-resistors under the illumination above the gap (visible, 600 nm) and below the gap (NIR, 900 nm). Laser Photonics Rev. [84], copyright 2016.

4.2.3 In near-infrared (NIR) region

NIR photodetectors have widespread uses in telecommunications [86], as well as thermal and biological imaging [87, 88, 89, 90]. Meredith’s group demonstrated the perovskite crystal that overcame the large bandgap and presented photodetectors with performance metrics appropriate for NIR detection by using the trap-related linear sub-gap absorption ( Figure 8e ) [84]. A strong NIR photo-response was achieved in photodiodes based on MAPbI3 crystals illuminated by a continuous 808-nm laser (∼10 mW/cm2). The photodiodes could also respond to a laser with a wavelength as long as 1064 nm ( Figure 8f ).

4.2.4 In X-ray region

In addition to the common light detections from UV to IR, perovskite crystals have been employed for the detection of X-rays, which have important applications in medical diagnostics, clinical treatment, and the non-destructive testing of products [53]. Huang et al. fabricated a sensitive MAPbBr3 crystal X-ray detector with the structure of Au/MAPbBr3/crystal/C60/BCP/Ag or Au ( Figure 9a ) [53]. Through reducing the bulk defects and passivating surface traps, the devices showed a detection efficiency of 16.4% at a near zero bias under irradiation with continuum X-ray energy up to 50 keV. The lowest detectable X-ray dose rate was 0.5 μGyair/s with a sensitivity of 80 μC/Gyaircm2, which is four times higher than the sensitivity achieved in α-Se-based X-ray detectors ( Figure 9b ). An X-ray detector based on p-i-n diode array made of a thick MAPbBr3 single crystal was introduced by Chen’s group [94], which displayed the highest sensitivity of 23.6 μC/mGyaircm2, indicating high potential for practical applications.

Figure 9.

9a, structure of MAPbBr3 crystal X-ray detector. 9b, X-ray-generated photocurrent at various dose rates. Nature Photonics [53], copyright 2016. 9c, attenuation coefficient and penetration depth of MAPbI3 and CdTe. 9d, photocurrent and a fit with Hecht model generated by Cu X-ray radiation (8 keV) in a MAPbI3 crystal. Nature Photonics [91], copyright 2016. 9e, pictures of guard ring electrode side, anode side and side view of a MAPbBr2.94Cl0.06 crystal detector. 9f, 137Cs energy spectrum obtained by crystal, CZT and NaI (Tl) detectors. Nature Mater. [92], Copyright 2017. 9 g, schematic of a Schottky-type MAPbI3 detector with asymmetrical electrode and the energy level diagram. Energy-resolved spectrum by Schottky-type MAPbI3 detector (9 h) under 241Am 59.5 keV γ-ray under −50 V and (9i) under 57Co 122 keV γ-ray under −70 V. ACS Photonics [93], copyright 2018.

4.2.5 In gamma-ray (γ-ray) region

Similar to X-ray detectors, the γ-ray detectors are also widely used in many fields, owing to the non-invasive detections. However, γ-ray detectors need to work in a weak radiation field pulse mode and perform event-by-event detections to sort out the intensity vs. the energy of the radiation quanta. Large and balanced μ and τ are needed for high-energy detection. Huang et al. reported high-quality MAPbI3 crystals that were applied to γ-ray detection with a 4% efficiency when operating in the γ-voltaic mode [63]. Kovalenko et al. demonstrated MAPbI3 crystals for γ-ray detection ( Figure 9c ), and a 59.6 keV 241Am energy spectrum was acquired [91]. A fit of bias dependence of photocurrent with Hecht model indicated a high μτ product of ∼10−2 cm2/V ( Figure 9d ) [95, 96].

Huang’s group further reported a Cl dopant compensation of MAPbBr3 single crystal process to fabricate a low-cost γ-ray detector [92]. MAPbBr2.94Cl0.06 crystals with a larger μτ product were equipped with a guard ring electrode to mitigate their leakage current ( Figure 9e ). The 137Cs energy spectrum obtained by such crystals with a full-energy peak resolution of 6.50% is compared with the spectrum obtained by CZT and NaI(Tl) detectors ( Figure 9f ). A high-performance MAPbI3 crystal γ-ray spectrometer was designed by Kanatzidis et al. [93], and the asymmetrical electrodes (Schottky-type) were applied to prohibit the hole injection from the anode or to reduce the leakage current ( Figure 9g ). The best energy resolution of the device for 241Am 59.5 keV γ-rays was ∼12%; while the best energy resolution achieved for 57Co 122 keV was 6.8% ( Figure 9h and i ).

4.3 Light-emitting diodes (LEDs) and lasers

With the exceptional PL efficiency and high color purity, perovskite crystals can also perform as high-performance LEDs [97]. Most of the existing perovskite LEDs employ a polycrystalline film with sizes of nanometers to micrometers, and coherent light emission is a challenge [98]. In Yu’s work, the LEDs with the structure of ITO/MAPbBr3 micro-platelet/Au cathode had the turn-on voltage of about 1.8 V and could last for at least 54 h with a luminance of ∼5000 cd/m2 ( Figure 10a ) [99].

Figure 10.

10a, light emission intensity vs. time of a perovskite LED at −193°C. inset: A microscopic image at t = 12 h. ACS Nano [99], copyright 2018. 10b, schematic for optical setup of a CH3NH3PbI3 nanoplatelet. 10c, evolution from spontaneous emission to lasing in a typical CH3NH3PbI3 nanoplatelet. Inset left: Optical image of a nanoplate and plot of integrated P out. Inset right: PL decay curve below (pink) and above (dark green) the threshold. Nano Lett. [100], Copyright 2014. 10d, nanowire emission spectra. Inset: Integrated emission intensity and FWHM vs. P. Nature Mater. [101], Copyright 2015. 10e, integrated PL intensity as a function of excitation density. Adv. Mater. [102], Copyright 2015. 10f, emission spectra of perovskite microplates excited by different pump densities. Inset: Integrated PL intensity vs. pump density. Adv. Mater. [103], Copyright 2016.

The excellent properties, including a small trap density, long lifetime and electron–hole diffusion length, and large carrier mobility, also make perovskite crystals suitable for laser devices with low lasing thresholds and high qualities. Xiong’s group grew typical MAPbI3 triangular nano-platelets and optically pumped them by a femtosecond-pulsed laser ( Figure 10b ) [100], and the peaks centered at λ = 776.7, 779.2, 781.9, 784.3, and 786.8 nm appeared over the spontaneous emission band with a FWHM of ∼1.2 nm ( Figure 10c ), when the pump fluence was increased to 40.6 μJ/cm2. Zhu et al. demonstrated room-temperature lasing via using MAPbI3 crystal nanowire, which had a broad tenability covering the NIR to visible region [101]. From Figure 10d , a sharp peak appeared at 787 nm in the representative emission spectra and grew rapidly with increasing the pump laser fluence (P) with the lasing threshold P Th of ∼595 nJ/cm2. Additionally, MAPbBr3 crystal square micro-disks were synthesized into a 557-nm single-mode laser based on a built-in whispering gallery mode micro-resonator by Fu’s group [102], from which a P Th = 3.6 μJ/cm2 was observed, and a sublinear regime was observed below the threshold ( Figure 10e ). Uniform-sized MAPbBr3 microplates were also created by Jiang et al. by using “liquid knife” and were made into lasers [103]. A 400-nm pulsed laser beam was used as a pump source to excite microplates, and a spontaneous emission peak centered at 550 nm with a FWHM of 20 nm was observed ( Figure 10f ).

Advertisement

5. Challenges and open issues

Hybrid perovskite single crystals have shown great potential in high-performance optoelectronic devices; however, several challenges and issues still remain in terms of their practical applications. They mainly include (1) the effects of surface defects, (2) the large-area fabrication, as well as (3) the stability of the perovskite single crystal devices and (4) the health and environmental concerns.

5.1 Surface defects

The absence of grain boundaries makes perovskite crystals acquire better optical and charge transport properties than their polycrystalline counterparts. However, the surface of crystals usually possesses lots of chemical impurities, dangling bonds, surface dislocations, and under-coordinated atoms, and becomes disordered owing to hydration, thus decreasing the carrier mobility and carrier diffusion length and promote the recombination of carriers near the crystal surface [76, 104, 105, 106]. Thus, the further decrease of defects, especially the surface defects, is required, aiming to gain high-quality perovskite crystals. To realize high-performance optoelectronic devices based on perovskite crystals with low-level surface defects, more research should be carried out on the surface passivation.

5.2 Large-area fabrication

Hybrid perovskite thin crystals are freer of grain boundaries and exhibit better transport properties than those of the polycrystalline candidates, so their large-area fabrication will ensure a promising future. However, the embedding of volatile and vulnerable organic components on fragile inorganic framework makes them difficult to be fabricated with a large area by deposition techniques or solution-based methods [42, 54]. Furthermore, thin crystals were grown directly on conductive substrates like FTO- or ITO-glass [42, 56], and tailored substrates, such as SiO2/Si [97], which provide in-situ growth for thin crystals and be directly made into devices. Nevertheless, these large-area thin crystals have rough surfaces and a great number of surface defects, and thus their optoelectronic properties remain inferior to the bulk counterparts. Further optimization of growth methods for large-area thin crystals is needed for industry productions in future.

5.3 Long-term stability

Low stability of the current hybrid perovskite crystal devices hinders their broad practical application. Several factors that affect the device stability, like ion migration [107, 108], can cause hysteresis and photo-induced phase separation, and the interaction between single crystals and their surroundings lies in the degradation of perovskite by humidity and light [109, 110, 111]. Therefore, to further enhance the stability of single crystal devices, optimized device structures should be designed to control the ion migrations. Meanwhile, various compositions and interface engineering approaches are also intensively investigated to confront this critical issue. In addition, encapsulation has been demonstrated to be a valid method to protect hybrid perovskite devices.

5.4 Health and environmental concerns

The growth of hybrid perovskite crystals adopt heavy metal ions, like lead (Pb) or tin (Sn), and organic functional groups, which can impact both the environment and human health. This critical issue needs to be overcome, aiming for further commercialization. As for the common MAPbI3 perovskite crystal, the Pb-ion is toxic to both the human health and natural environment; while the organic solvents used during the growth process of crystals are also toxic and easily penetrate into the human body [112]. To solve these problems, capsulation and recycling are needed in the use of crystal materials and organic solvents. Furthermore, other alternative metals to Pb, with a lower toxicity, are also being studied, such as bismuth and antimony [113, 114], and thus, the optoelectronic properties of these Pb-free perovskite crystals need to be explored further for device applications.

Advertisement

6. Conclusions

More recently, hybrid perovskite crystals, having different dimensional forms: bulk and thin crystals, and micro−/nano-plates, have been widely explored as functional layers for optoelectronic devices owing to their excellent physical properties combined with the advantage of ease of processing. Although these types of devices are still in the early stages of development, a strong potential for a variety of technological and commercial applications clearly remains. Here, we presented a comprehensive overview of the recent advances in hybrid perovskite crystals with respect to the background knowledge on the optoelectronic properties and charge transport dynamics of crystals, and their applications in the area of optoelectronic devices, and a fundamental understanding of the device performance. We summarized the main growth methods for the bulk crystals and also some modified and optimized approaches to synthesize thin crystals. The detailed discussions are focused on charge transport characteristics, operation mechanisms, and challenges, aiming to provide a critical understanding of further advance in materials design and device engineering in a variety of optoelectronic technologies.

In conclusion, the research progress achieved to date in the area of perovskite crystal optoelectronic devices, with the emphasis placed on challenges faced by the research community, has been summarized systematically, and finally perspective on the opportunities offered by this emerging family to photoactive materials in practical and commercial technologies is also proposed. Further exploration of high-quality perovskite crystals, combined with an in-depth understandings of working mechanism of devices, indicates a promising future for wide applications with markedly-enhanced performance.

Advertisement

Acknowledgments

The author acknowledges support from Discovery Early Career Researcher Award (DECRA) (DE180100167) from the Australian Research Council (ARC).

Advertisement

Conflict of interest

There are no conflicts to declare.

References

  1. 1. Mitzi, D. B. Templating and structural engineering in organic–inorganic perovskites. J. Chem. Soc., Dalton Trans., 2001, 1-12
  2. 2. Green, M. A.; Ho-Baillie, A.; Snaith, H. J. The emergence of perovskite solar cells. Nature photonics 2014, 8, 506
  3. 3. Zhang, D.; Zhu, Y.; et al. Atomic-resolution transmission electron microscopy of electron beam–sensitive crystalline materials. Science 2018, 359, 675-679
  4. 4. Ke, W.; Mao, L.; et al. Compositional and solvent engineering in Dion–Jacobson 2D perovskites boosts solar cell efficiency and stability. Adv. Energy Mater. 2019, 9, 1803384
  5. 5. Li, S.; Tong, S.; et al. Fast-response and high-responsivity FAxMA(1−x)PbI3 photodetectors fabricated via doctor-blading deposition in ambient condition. Org. Electron. 2018, 52, 190-194
  6. 6. Wang, H.; Wu, H.; et al. Controllable CsxFA1–xPbI3 single-crystal morphology via rationally regulating the diffusion and collision of micelles toward high-performance photon detectors. ACS Appl. Mater. Interfaces 2019, 11, 13812-13821
  7. 7. Hou, X.; Xu, M.; et al. High performance printable perovskite solar cells based on Cs0.1FA0.9PbI3 in mesoporous scaffolds. J. Power Sources 2019, 415, 105-111
  8. 8. Wang, M.; Jiang, X.; et al. High-performance and stable mesoporous perovskite solar cells via well-crystallized FA0.85MA0.15Pb(I0.8Br0.2)3. ACS Appl. Mater. Interfaces 2018, 11, 2989-2996
  9. 9. Babayigit, A.; Thanh, D. D.; et al. Assessing the toxicity of Pb- and Sn-based perovskite solar cells in model organism Danio rerio. Sci. Rep. 2016, 6, 1-11
  10. 10. Ma, H.; Ma, Y.; et al. Experimental phonon dispersion and lifetimes of tetragonal CH3NH3PbI3 perovskite crystals. J. Phys. Chem. Let. 2018, 10, 1-6
  11. 11. Ju, D.; Zheng, X.; et al. Reversible band gap narrowing of Sn-based hybrid perovskite single crystal with excellent phase stability. Angew. Chem. Int. Ed. 2018, 57, 14868-14872
  12. 12. Ju, D.; Dang, Y.; et al. Tunable band gap and long carrier recombination lifetime of stable mixed CH3NH3PbxSn1–xBr3 single crystals. Chem. Mater. 2018, 30, 1556-1565
  13. 13. Shao, S.; Liu, J.; et al. Highly reproducible Sn-based hybrid perovskite solar cells with 9% efficiency. Adv. Energy Mater. 2018, 8, 1702019
  14. 14. Li, C.; Lu, X.; et al. Formability of ABX3 (X= F, Cl, Br, I) halide perovskites. Acta Crystallogr. B Struct. Sci. 2008, 64, 702-707
  15. 15. Shi, D.; Adinolfi, V.; et al. Low trap-state density and long carrier diffusion in organolead trihalide perovskite single crystals. Science 2015, 347, 519-522
  16. 16. Kojima, A.; Teshima, K.; et al. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc. 2009, 131, 6050-6051
  17. 17. Xiao, Z.; Bi, C.; et al. Efficient, high yield perovskite photovoltaic devices grown by interdiffusion of solution-processed precursor stacking layers. Energy Environ. Sci. 2014, 7, 2619-2623
  18. 18. Liu, M.; Johnston, M. B.; Snaith, H. J. Efficient planar heterojunction perovskite solar cells by vapour deposition. Nature 2013, 501, 395-398
  19. 19. Xiao, M.; Huang, F.; et al. A fast deposition-crystallization procedure for highly efficient lead iodide perovskite thin-film solar cells. Angew. Chem. Int. Ed. 2014, 53, 9898-9903
  20. 20. Zhu, W.; Bao, C.; et al. An efficient planar-heterojunction solar cell based on wide-bandgap CH3NH3PbI2.1Br0.9 perovskite film for tandem cell application. Chem. Commun. 2016, 52, 304-307
  21. 21. Heo, J. H.; Im, S. H.; et al. Efficient inorganic–organic hybrid heterojunction solar cells containing perovskite compound and polymeric hole conductors. Nature photonics 2013, 7, 486
  22. 22. Jeon, N. J.; Noh, J. H.; et al. Solvent engineering for high-performance inorganic–organic hybrid perovskite solar cells. Nature Mater. 2014, 13, 897-903
  23. 23. Lee, M. M.; Teuscher, J.; et al. Efficient hybrid solar cells based on meso-super structured organometal halide perovskites. Science 2012, 338, 643-647
  24. 24. Filip, M. R.; Eperon, G. E.; et al. Steric engineering of metal-halide perovskites with tunable optical band gaps. Nature Commun. 2014, 5, 1-9
  25. 25. D’Innocenzo, V.; Srimath Kandada, A. R.; et al. Tuning the light emission properties by band gap engineering in hybrid lead halide perovskite. J. Am. Chem. Soc. 2014, 136, 17730-17733
  26. 26. Sum, T. C.; Mathews, N. Advancements in perovskite solar cells: photophysics behind the photovoltaics. Energy Environ. Sci. 2014, 7, 2518-2534
  27. 27. Choi, H.; Jeong, J.; et al. Cesium-doped methylammonium lead iodide perovskite light absorber for hybrid solar cells. Nano Energy 2014, 7, 80-85
  28. 28. Kim, Y. H.; Cho, H.; et al. Multicolored organic/inorganic hybrid perovskite light-emitting diodes. Adv. Mater. 2015, 27, 1248-1254
  29. 29. NREL. Efficiency chart. https://www.nrel.gov/pv/assets/pdfs/pv-efficiency-chart.20181214.pdf
  30. 30. Hu, X.; Zhang, X.; et al. High-performance flexible broadband photodetector based on organolead halide perovskite. Adv. Funct. Mater. 2014, 24, 7373-7380
  31. 31. Chin, X. Y.; Cortecchia, D.; et al. Lead iodide perovskite light-emitting field-effect transistor. Nature Commun. 2015, 6, 7383
  32. 32. Li, F.; Ma, C.; et al. Ambipolar solution-processed hybrid perovskite phototransistors. Nature Commun. 2015, 6, 8238
  33. 33. Meng, L.; Yao, E. P.; et al. Pure formamidinium-based perovskite light-emitting diodes with high efficiency and low driving voltage. Adv. Mater. 2017, 29, 1603826
  34. 34. Liu, P.; He, X.; et al. Organic–inorganic hybrid perovskite nanowire laser arrays. ACS Nano 2017, 11, 5766-5773
  35. 35. Duan, H.-S.; Zhou, H.; et al. The identification and characterization of defect states in hybrid organic–inorganic perovskite photovoltaics. Phys. Chem. Chem. Phys. 2015, 17, 112-116
  36. 36. Zheng, X.; Chen, B.; et al. Defect passivation in hybrid perovskite solar cells using quaternary ammonium halide anions and cations. Nature Energy 2017, 2, 1-9
  37. 37. Liu, Z.; Hu, J.; et al. Chemical reduction of intrinsic defects in thicker heterojunction planar perovskite solar cells. Adv. Mater. 2017, 29, 1606774
  38. 38. Fang, H.-H.; Adjokatse, S.; et al. Ultrahigh sensitivity of methylammonium lead tribromide perovskite single crystals to environmental gases. Sci. Adv. 2016, 2, e1600534
  39. 39. Huang, J.; Shao, Y.; Dong, Q . Organometal trihalide perovskite single crystals: a next wave of materials for 25% efficiency photovoltaics and applications beyond? J. Phys. Chem. Lett. 2015, 6, 3218-3227
  40. 40. Zhang, F.; Yang, B.; et al. Extra-long electron–hole diffusion lengths in CH3NH3PbI3−xClx perovskite single crystals. J. Mater. Chem. C 2017, 5, 8431-8435
  41. 41. Chen, B.-X.; Li, W.-G.; et al. Large-grained perovskite films via FAxMA1−xPb(IxBr1−x)3 single crystal precursor for efficient solar cells. Nano Energy 2017, 34, 264-270
  42. 42. Saidaminov, M. I.; Adinolfi, V.; et al. Planar-integrated single-crystalline perovskite photodetectors. Nature Commun. 2015, 6, 1-7
  43. 43. Li, S.; Zhang, C.; et al. Metal halide perovskite single crystals: from growth process to application. Crystals 2018, 8, 220
  44. 44. Liu, Y.; Yang, Z.; Liu, S. Recent progress in single-crystalline perovskite research Including crystal preparation, property evaluation, and applications. Adv. Sci. 2018, 5, 1700471
  45. 45. Chen, Y.; He, M.; et al. Structure and growth control of organic–inorganic halide perovskites for optoelectronics: from polycrystalline films to single crystals. Adv. Sci. 2016, 3, 1500392
  46. 46. Ding, J.; Yan, Q . Progress in organic-inorganic hybrid halide perovskite single crystal: growth techniques and applications. Sci. China Mater. 2017, 60, 1063-1078
  47. 47. Dang, Y.; Liu, Y.; et al. Bulk crystal growth of hybrid perovskite material CH3NH3PbI3. CrystEngComm 2015, 17, 665-670
  48. 48. Su, J.; Chen, D.; Lin, C. Growth of large CH3NH3PbX3 (X= I, Br) single crystals in solution. J. Cryst. Growth 2015, 422, 75-79
  49. 49. Fang, Y.; Dong, Q .; et al. Highly narrowband perovskite single-crystal photodetectors enabled by surface-charge recombination. Nature Photonics 2015, 9, 679
  50. 50. Saidaminov, M. I.; Abdelhady, A. L.; et al. High-quality bulk hybrid perovskite single crystals within minutes by inverse temperature crystallization. Nature Commun. 2015, 6, 1-6
  51. 51. Zhang, Y.; Liu, Y.; et al. Perovskite CH3NH3Pb(BrxI1−x)3 single crystals with controlled composition for fine-tuned bandgap towards optimized optoelectronic applications. J. Mater. Chem. C 2016, 4, 9172-9178
  52. 52. Zhang, B.; Yan, J.; et al. Effect of the modulating of organic content on optical properties of single-crystal perovskite. Opt. Mater. 2016, 62, 273-278
  53. 53. Wei, H.; Fang, Y.; et al. Sensitive X-ray detectors made of methylammonium lead tribromide perovskite single crystals. Nature Photonics 2016, 10, 333
  54. 54. Peng, W.; Wang, L.; et al. Solution-grown monocrystalline hybrid perovskite films for hole-transporter-free solar cells. Adv. Mater. 2016, 28, 3383-3390
  55. 55. Liu, Y.; Zhang, Y.; et al. Thinness-and shape-controlled growth for ultrathin single-crystalline perovskite wafers for mass production of superior photoelectronic devices. Adv. Mater. 2016, 28, 9204-9209
  56. 56. Rao, H. S.; Li, W. G.; et al. In-situ growth of 120-cm2 CH3NH3PbBr3 perovskite crystal film on FTO glass for narrowband-photodetectors. Adv. Mater. 2017, 29, 1602639
  57. 57. Chen, Y.-X.; Ge, Q .-Q .; et al. General space-confined on-substrate fabrication of thickness-adjustable hybrid perovskite single-crystalline thin films. J. Am. Chem. Soc. 2016, 138, 16196-16199
  58. 58. Lee, L.; Baek, J.; et al. Wafer-scale single-crystal perovskite patterned thin films based on geometrically-confined lateral crystal growth. Nature Commun. 2017, 8, 1-8
  59. 59. Gu, Z.; Huang, Z.; et al. A general printing approach for scalable growth of perovskite single-crystal films. Sci. Adv. 2018, 4, eaat2390
  60. 60. Maculan, G.; Sheikh, A. D.; et al. CH3NH3PbCl3 single crystals: inverse temperature crystallization and visible-blind UV-photodetector. J. Phys. Chem. Lett. 2015, 6, 3781-3786
  61. 61. Wenger, B.; Nayak, P. K.; et al. Consolidation of the optoelectronic properties of CH3NH3PbBr3 perovskite single crystals. Nature Commun. 2017, 8, 1-10
  62. 62. Zhang, Y.; Liu, Y.; et al. High-quality perovskite MAPbI3 single crystals for broad-spectrum and rapid response integrate photodetector. J. Energy Chem. 2018, 27, 722-727
  63. 63. Dong, Q .; Fang, Y.; et al. Electron-hole diffusion lengths >175 μm in solution-grown CH3NH3PbI3 single crystals. Science 2015, 347, 967-970
  64. 64. Yi, H. T.; Wu, X.; et al. Intrinsic charge transport across phase transitions in hybrid organo-inorganic perovskites. Adv. Mater. 2016, 28, 6509-6514
  65. 65. Valverde-Chávez, D. A.; Ponseca, C. S.; et al. Intrinsic femtosecond charge generation dynamics in single crystal CH3NH3PbI3. Energy Environ. Sci. 2015, 8, 3700-3707
  66. 66. Yu, W.; Li, F.; et al. Single crystal hybrid perovskite field-effect transistors. Nature Commun. 2018, 9, 1-10
  67. 67. Edri, E.; Kirmayer, S.; et al. High open-circuit voltage solar cells based on organic–inorganic lead bromide perovskite. J. Phys. Chem. Lett. 2013, 4, 897-902
  68. 68. Giorgi, G.; Yamashita, K. Organic–inorganic halide perovskites: an ambipolar class of materials with enhanced photovoltaic performances. J. Mater. Chem. A 2015, 3, 8981-8991
  69. 69. Ball, J. M.; Petrozza, A., Defects in perovskite-halides and their effects in solar cells. Nature Energy 2016, 1, 1-13
  70. 70. Zhang, M.; Yu, H.; et al. Composition-dependent photoluminescence intensity and prolonged recombination lifetime of perovskite CH3NH3PbBr3−xClx films. Chem. Commun. 2014, 50, 11727-11730
  71. 71. Fang, Y.; Wei, H.; et al. Quantification of re-absorption and re-emission processes to determine photon recycling efficiency in perovskite single crystals. Nature Commun. 2017, 8, 14417
  72. 72. Chen, Z.; Dong, Q .; et al. Thin single crystal perovskite solar cells to harvest below-bandgap light absorption. Nature Commun. 2017, 8, 1-7
  73. 73. Dong, Q .; Song, J.; et al. Lateral-structure single-crystal hybrid perovskite solar cells via piezoelectric poling. Adv. Mater. 2016, 28, 2816-2821
  74. 74. Chen, Z.; Turedi, B.; et al. Single-crystal MAPbI3 perovskite solar cells exceeding 21% power conversion efficiency. ACS Energy Lett. 2019, 4, 1258-1259
  75. 75. Bao, C.; Chen, Z.; et al. Low-noise and large-linear-dynamic-range photodetectors based on hybrid-perovskite thin-single-crystals. Adv. Mater. 2017, 29, 1703209
  76. 76. Lian, Z.; Yan, Q .; et al. High-performance planar-type photodetector on (100) facet of MAPbI3 single crystal. Sci. Rep. 2015, 5, 16563
  77. 77. Cao, M.; Tian, J.; et al. Perovskite heterojunction based on CH3NH3PbBr3 single crystal for high-sensitive self-powered photodetector. Appl. Phys. Lett. 2016, 109, 233303
  78. 78. Qin, X.; Yao, Y.; et al. Perovskite photodetectors based on CH3NH3PbI3 single crystals. Chem. Asian J. 2016, 11, 2675-2679
  79. 79. Yang, Z.; Deng, Y.; et al. High-performance single-crystalline perovskite thin-film photodetector. Adv. Mater. 2018, 30, 1704333
  80. 80. Cheong, P.; Chang, K.-F.; et al. A ZigBee-based wireless sensor network node for ultraviolet detection of flame. IEEE Trans. Ind. Electron. 2011, 58, 5271-5277
  81. 81. Li, M.; Li, W.; et al. On-site determination and monitoring of real-time fluence delivery for an operating UV reactor based on a true fluence rate detector. Environ. Sci. Technol. 2017, 51, 8094-8100
  82. 82. Ojeda, C. B.; Rojas, F. S. Process analytical chemistry: applications of ultraviolet/visible spectrometry in environmental analysis: an overview. Appl. Spectrosc. Rev. 2009, 44, 245-265
  83. 83. Adinolfi, V.; Ouellette, O.; et al. Fast and sensitive solution-processed visible-blind perovskite UV photodetectors. Adv. Mater. 2016, 28, 7264-7268
  84. 84. Lin, Q .; Armin, A.; et al. Near infrared photodetectors based on sub-gap absorption in organohalide perovskite single crystals. Laser Photonics Rev. 2016, 10, 1047-1053
  85. 85. Dong, R.; Fang, Y.; et al. High-gain and low-driving-voltage photodetectors based on organolead triiodide perovskites. Adv. Mater. 2015, 27, 1912-1918
  86. 86. Chin, A.; Vaddiraju, S.; et al. Near-infrared semiconductor subwavelength-wire lasers. Appl. Phys. Lett. 2006, 88, 163115
  87. 87. Cheng, S.-Q .; Cai, B.; Zhu, Y.-M. Black silicon as absorber for near-infrared photo-thermal conversion, 2015 Opto-Electronics and Communications Conference (OECC), IEEE: 2015; pp 1-3
  88. 88. Du, K.-K.; Li, Q .; et al. Control over emissivity of zero-static-power thermal emitters based on phase-changing material GST. Light Sci. Appl. 2017, 6, e16194-e16194
  89. 89. Wang, Z.; Li, Y.; et al. Pure near-infrared to near-infrared up-conversion of multifunctional Tm3+ and Yb3+ co-doped NaGd (WO4)2 nanoparticles. J. Mater. Chem. C 2014, 2, 4495-4501
  90. 90. Naczynski, D.; Tan, M.; et al Rare-earth-doped biological composites as in vivo shortwave infrared reporters. Nature Commun. 2013, 4, 1-10
  91. 91. Yakunin, S.; Dirin, D. N.; et al. Detection of gamma photons using solution-grown single crystals of hybrid lead halide perovskites. Nature Photonics 2016, 10, 585
  92. 92. Wei, H.; DeSantis, D.; et al. Dopant compensation in alloyed CH3NH3PbBr3−xClx perovskite single crystals for gamma-ray spectroscopy. Nature Mater. 2017, 16, 826-833
  93. 93. He, Y.; Ke, W.; et al. Resolving the energy of γ-ray photons with MAPbI3 single crystals. ACS Photonics 2018, 5, 4132-4138
  94. 94. Wang, X.; Zhao, D.; et al. PIN diodes array made of perovskite single crystal for X-ray imaging. Phys. Status Solidi RRL 2018, 12, 1800380
  95. 95. Street, R.; Ready, S.; et al. Comparison of PbI2 and HgI2 for direct detection active matrix X-ray image sensors. J. Appl. Phys. 2002, 91, 3345-3355
  96. 96. Kabir, M. Z.; Kasap, S. Charge collection and absorption-limited sensitivity of X-ray photoconductors: Applications to a-Se and HgI2. Appl. Phys. Lett. 2002, 80, 1664-1666
  97. 97. Nguyen, V.-C.; Katsuki, H.; et al. Single-crystal perovskite CH3NH3PbBr3 prepared by cast capping method for light-emitting diodes. Jpn. J. Appl. Phys. 2018, 57, 04FL10
  98. 98. Chih, Y. K.; Wang, J. C.; et al. NiOx electrode interlayer and CH3NH2/CH3NH3PbBr3 Interface treatment to markedly advance hybrid perovskite-based light-emitting diodes. Adv. Mater. 2016, 28, 8687-8694
  99. 99. Chen, M.; Shan, X.; et al. Manipulating ion migration for highly stable light-emitting diodes with single-crystalline organometal halide perovskite microplatelets. ACS Nano 2017, 11, 6312-6318
  100. 100. Zhang, Q .; Ha, S. T.; et al. Room-temperature near-infrared high-Q perovskite whispering-gallery planar nanolasers. Nano Lett. 2014, 14, 5995-6001
  101. 101. Zhu, H.; Fu, Y.; et al. Lead halide perovskite nanowire lasers with low lasing thresholds and high quality factors. Nature Mater. 2015, 14, 636-642
  102. 102. Liao, Q .; Hu, K.; et al. Perovskite microdisk microlasers self-assembled from solution. Adv. Mater. 2015, 27, 3405-3410
  103. 103. Feng, J.; Yan, X.; et al. “Liquid knife” to fabricate patterning single-crystalline perovskite microplates toward high-performance laser arrays. Adv. Mater. 2016, 28, 3732-3741
  104. 104. Wu, B.; Nguyen, H. T.; et al. Discerning the surface and bulk recombination kinetics of organic–inorganic halide perovskite single crystals. Adv. Energy Mater. 2016, 6, 1600551
  105. 105. Murali, B.; Yengel, E.; et al. The surface of hybrid perovskite crystals: a boon or bane. ACS Energy Lett. 2017, 2, 846-856
  106. 106. Yang, Y.; Yang, M.; et al. Top and bottom surfaces limit carrier lifetime in lead iodide perovskite films. Nature Energy 2017, 2, 1-7
  107. 107. Yuan, Y.; Huang, J. Ion migration in organometal trihalide perovskite and its impact on photovoltaic efficiency and stability. Acc. Chem. Res. 2016, 49, 286-293
  108. 108. Deng, Y.; Xiao, Z.; Huang, J. Light-induced self-poling effect on organometal trihalide perovskite solar cells for increased device efficiency and stability. Adv. Energy Mater. 2015, 5, 1500721
  109. 109. Li, W.; Fan, J.; et al. Aquointermediate assisted highly orientated perovskite thin films toward thermally stable and efficient solar cells. Adv. Energy Mater. 2017, 7, 1601433
  110. 110. Grancini, G.; D'Innocenzo, V.; et al. CH3NH3PbI3 perovskite single crystals: surface photophysics and their interaction with the environment. Chem. Sci. 2015, 6, 7305-7310
  111. 111. Ecker, B. R.; Wang, C.; et al. Intrinsic behavior of CH3NH3PbBr3 single crystals under light illumination. Adv. Mater. Interfaces 2018, 5, 1801206
  112. 112. Williams, A. C.; Barry, B. W. Penetration enhancers. Adv. Drug Deliv. Rev. 2012, 64, 128-137
  113. 113. Lyu, M.; Yun, J.-H.; et al. Organic–inorganic bismuth (III)-based material: A lead-free, air-stable and solution-processable light-absorber beyond organolead perovskites. Nano Res. 2016, 9, 692-702
  114. 114. Leng, M.; Chen, Z.; et al. Lead-free, blue emitting bismuth halide perovskite quantum dots. Angew. Chem. Int. Ed. 2016, 55, 15012-15016

Written By

Feng Li

Submitted: 15 September 2020 Reviewed: 17 November 2020 Published: 09 December 2020