Open Access
15 April 2016 Pathways toward high-performance perovskite solar cells: review of recent advances in organo-metal halide perovskites for photovoltaic applications
Zhaoning Song, Suneth C. Watthage, Adam B. Phillips, Michael J. Heben
Author Affiliations +
Abstract
Organo-metal halide perovskite–based solar cells have been the focus of intense research over the past five years, and power conversion efficiencies have rapidly been improved from 3.8 to >21%. This article reviews major advances in perovskite solar cells that have contributed to the recent efficiency enhancements, including the evolution of device architecture, the development of material deposition processes, and the advanced device engineering techniques aiming to improve control over morphology, crystallinity, composition, and the interface properties of the perovskite thin films. The challenges and future directions for perovskite solar cell research and development are also discussed.

1.

Introduction

In 1954, the first practical photovoltaic (PV) device based on crystalline silicon was demonstrated at Bell Laboratories.1 After many decades of progress, crystalline silicon technology dominates the global PV market with a 55% and 36% market share for polycrystalline- and monocrystalline-silicon modules in 2014, respectively.2 The remaining 9% of the market was split between a variety of other established and emerging PV technologies, including polycrystalline thin films, amorphous semiconductors, dye-sensitized solar cells (DSSCs), organics, and quantum dot solar cells.3 To gain market share from crystal silicon solar cells, alternative technologies have to provide a desirable combination of high power conversion efficiency (PCE), low manufacturing costs, and excellent stability. Recent research suggests that organo-metal halide perovskites (OMHPs), with methylammonium lead iodide (CH3NH3PbI3 or MAPbI3) being the prototypical example, have the potential to meet these conditions and become competitive in the marketplace. As a result of intensive research efforts across the globe over the past three years,413 perovskite-based solar cell PCEs are now comparable to or better than most other PV technologies, and the simple device processing promises lower manufacturing costs, suggesting the potential to challenge the prevailing silicon technology in the foreseeable future.14

The term perovskite refers to the crystal structure of calcium titanate (CaTiO3), which was discovered by the German mineralogist Gustav Rose in 1839 and named in honor of the Russian mineralogist Lev Perovski.15 In the field of optoelectronics, OMHPs are a group of materials with the formula AMX3, where A is an organic cation (CH3NH3+ or NH2CH3NH2+), M is a divalent metal cation (Pb2+ or Sn2+), and X is a monovalent halide anion (I, Br, or Cl). Figure 1 shows the crystal structure and a single crystal of MAPbI3. In a unit cell of the OHMP structure, eight A+ cations are located at the vertices of a cubic cage, an M2+ cation is located at the center of the cube, and the latter species is octahedrally coordinated to six X species that sit at the cube’s faces. The OMHP family of materials were studied in the 1990s due to their excellent optoelectronic properties and potential for solution-processed fabrication,17,1820 but the main goal of this early work was to develop new materials for field effect transistors and organic light-emitting diodes.21,22

Fig. 1

(a) Crystal structure of CH3NH3PbI3 perovskite and (b) photo of a CH3NH3PbI3 perovskite single crystal synthesized in our lab using the inverse temperature crystallization method.16

JPE_6_2_022001_f001.png

The first known use of OMHP was as a dye in DSSC, which reported a 3% PCE in 2009.23 However, this OMHP-based solar cell contained a liquid electrolyte and received little attention due to the low efficiency and poor stability. The so-called perovskite fever24 did not fully bloom until a solid-state cell was developed and devices with 10% efficiency were reported in 2012.25,26 Since then, OMHP-based PV device performance has rapidly progressed, and a best efficiency record of >21% was achieved in late 2015.27 The pace of progress has been remarkable and unprecedented in PV history and can likely be attributed to several factors related to inexpensive fabrication costs, ease of processing, and the excellent optoelectronic properties of the materials.7,9,1013

As will be described in Sec. 3, high-quality perovskite thin films can be fabricated using a variety of processes including solution-26,28 and vapor-based2931 deposition methods. Many of these methods are compatible with low-cost, large-scale, industrial production techniques, which strengthens the potential for the commercialization of perovskite solar cells. Due to the ease of processing, many research groups from around the world have been attracted to work in the area. This includes groups that have past histories and relevant expertise in DSSC, organic photovoltaics (OPV), and solution processing. Consequently, the learning curve for developing perovskite solar cells has been relatively short, and progress has been very rapid.

In addition to flexibility in processing, OMHP materials possess several outstanding optoelectronic properties that make them ideal choices for PV applications. The 1.55 eV band gap of MAPbI3 is nearly ideal for single-junction solar cells exposed to the solar irradiance spectrum, and it can be continuously varied in the range from 1.5 to 2.3 eV by exchanging the organic and halide ions.32,33 The optical absorption coefficient of MAPbI3 is higher than other PV materials such as Si, CdTe, CuGaxIn1xSySe1y (CIGS), and amorphous Si:H, so the absorber thickness can be reduced to 300  nm, thereby lowering the material costs.34,35 In contrast to organic PV materials, the low exciton binding energy (30 to 50 meV) allows spontaneous exciton dissociation into free charges after light absorption.3638 Moreover, the high electron and hole mobility in the range of 10 to 60  cm2V1s1 and the long carrier lifetime (100  ns) result in long diffusion lengths (1  μm) so that charge carriers can be freely transported across the 300-nm thick perovskite absorber before recombination.3942 Finally, because the electronic defects are shallow and relatively benign, the nonradiative recombination rates are low, allowing open-circuit voltages >1  V to be achieved.43,44

Although the perovskite solar cells show great potential, there are several challenges that need to be addressed before commercialization will be possible. Perhaps most significantly, OMHPs have not yet demonstrated the long-term stability that is necessary to compete with the 30-year lifetime of commercially available Si and CdTe solar panels. Second, there are questions about the current–voltage (JV) hysteresis during voltage scanning, which could be problematic for large-scale deployment. There are also concerns associated with potential environmental impacts due to the fact that OMHPs contain Pb.

This review focuses on the recent advances that have allowed perovskite PV to improve to efficiencies >21% in the time period of 2013 to 2015. Reviews that discuss early developments and the materials can be found elsewhere.7,9,1013 Here, we provide insights on the key factors that govern the device performance, including device architecture, preparation methods, and advanced device engineering. Notable devices over this time frame are highlighted. Additionally, the stability issues and future directions for perovskite PV devices are discussed.

2.

Device Architectures

The first OMHPs employed in PV were used as direct replacements for the dye sensitizers in the DSSCs.23,45 The typical DSSC structure employs a several-micron thick porous TiO2 layer that is coated and penetrated with an absorber dye material. The electrode assembly is contacted by a liquid electrolyte containing a redox couple.46 In these devices, TiO2 is used to collect and transport the electrons, while the electrolyte acts as a hole conductor. The original perovskite solar cells evolved from this same structure, with the OMHP materials acting simply as a dye replacement.25,26 Interest increased when the so-called mesoscopic device structure [Fig. 2(a)] was formed by replacing the liquid electrolyte with a solid-state hole conductor.25,26 This advance created great interest in the PV community and drew in experts from the thin-film PV and OPV communities. As a result, planar device structures in which the OMHP absorber is sandwiched between electron and hole transporting materials (ETM and HTM) were developed. Depending on which transport material is encountered by the light first, these planar structures can be categorized as either the conventional n-i-p [Fig. 2(b)] or the inverted p-i-n [Fig. 2(c)] structures. Recently, a mesoscopic p-i-n structure [Fig. 2(d)] has also been developed.47,48 Due to processing differences, the device architecture determines the choice of charge transport (ETM and HTM) and collection (cathode and anode) materials, the corresponding material preparation methods, and, consequently, the performance of the devices. To date, no perovskite devices with significant efficiency have been constructed on opaque substrates (e.g., Ti foils)49,50 because the conventional deposition technologies for transparent conducting oxides (TCO) may lead to decomposition of the surface of the OMHP.

Fig. 2

Schematic diagrams of perovskite solar cells in the (a) n-i-p mesoscopic, (b) n-i-p planar, (c) p-i-n planar, and (d) p-i-n mesoscopic structures.

JPE_6_2_022001_f002.png

2.1.

Conventional n-i-p Structure

The mesoscopic n-i-p structure is the original architecture of the perovskite PV devices and is still widely used to fabricate high-performance devices. The structure [Fig. 2(a)] consists of a TCO cathode [fluorine doped tin oxide (FTO)], a 50- to 70-nm thick compact ETM (typically TiO2), a 150- to 300-nm thick mesoporous metal oxide (mp-TiO2 or mp-Al2O3) that is filled with perovskites, followed by an up to 300-nm perovskite capping layer, a 150- to 200-nm thick layer of 2,2’,7,7’-tetrakis(N,N-di-p-methoxyphenylamine)-9,9’-spirobifluorene (spiro-MeOTAD), which is a hole conductor, and 50 to 100 nm of a metal anode (Au or Ag).

In this structure, the mesoscopic layer is thought to enhance charge collection by decreasing the carrier transport distance, preventing direct current leakage between the two selective contacts and increasing photon absorption due to light scattering. Accordingly, the original mesoscopic perovskite devices used a thick (>500  nm) porous layer to efficiently absorb the incident light.25,51 But because the grain growth of the perovskites is confined by the pores in the structure, a significant amount of the material is present in disordered and amorphous phases.52 This leads to relatively low open-circuit voltage (VOC) and short-circuit current density (JSC).53 Surprisingly, thinning the mesoporous layer to 150 to 200 nm results in improved device efficiency due to enhanced crystallinity in the perovskite absorber. Additionally, the pore filling fraction and morphology of the perovskites is critically dependent upon the thickness of mp-TiO2.54,55 When the porous layer thickness is reduced to <300  nm, the pore filling fraction is increased and a perovskite capping layer forms on top of the porous structure. Complete pore filling accompanied by formation of a capping layer assures high charge transport rates and high collection efficiencies at the TiO2 interface. Once the charges are separated, recombination pathways between electrons in the TiO2 and holes in the HTM are blocked due to the relative positions in energy of the respective conduction and valence bands (vide infra).54 Consequently, the meso n-i-p structure is the most popular structure reported in the literature. The previous record efficiency value (20.2%) was measured from a cell formed in the mesoscopic structure that had discrete perovskite nanocrystals embedded in the porous ETM film with an overlaying continuous and dense perovskite capping layer.56

The planar n-i-p structure [Fig. 2(b)] is the natural evolution of the mesoscopic structure. A larger area mesoporous ETM was initially considered critical for high-efficiency perovskite devices because hole extraction at the HTM interfaces is significantly more efficient than electron extraction at the ETM interfaces.57 However, by delicately controlling the formation of the perovskite absorber, and the interfaces among the perovskite, carrier transport layers, and electrodes, high efficiencies can now be achieved without a mesoporous layer.58 To date, the best planar n-i-p device showed a 19.3% efficiency after careful optimization of the electron selective indium tin oxide (ITO)/TiO2 interfaces.58 Although the planar n-i-p perovskite solar cell usually exhibits enhanced VOC and JSC relative to a comparative mesoscopic device processed with the same materials and approach, the planar device usually exhibits more severe JV hysteresis (see Sec. 6). Thus, the state-of-the-art n-i-p devices usually include a thin (150  nm) mesoporous buffer layer filled and capped with the perovskite.56

2.2.

Inverted p-i-n Structure

When the deposition order is changed and the HTM layer is deposited first, the device is fabricated in the p-i-n structure [Fig. 2(c)]. In this case, the p-i-n type perovskite device is built on a 50- to 80-nm p-type conducting polymer such as poly(3,4-ethylenedioxythiophene) poly(styrene-sulfonate) (PEDOT:PSS), which is deposited on ITO-coated substrates. After depositing a 300-nm intrinsic perovskite thin film, the device is completed with a 10- to 60-nm organic hole-blocking layer [6,6]-phenyl C61-butyric acid methyl ester (PCBM) and a metal cathode (Al or Au). Early device design utilized a perovskite and fullerene (C60) donor–acceptor pair, which is typical in OPV.59 In fact, the commonality in structure has allowed OPV researchers to easily move into the field of perovskites. As the field has advanced, the organic acceptor has been omitted in favor of an ETM layer, leaving the planar perovskite absorber sandwiched between two opposite organic charge transporting materials.60 Recently, the efficiency of the planar p-i-n device has improved significantly due to the use of more advanced material preparation methods, such as a multicycle solution coating process, and a best efficiency of 18.9% was achieved.61

Further development of the p-i-n device structure has expanded the selective contact options from organic to inorganic materials. For example, NiO and ZnO/TiO2 layers have recently been used for the hole and electron selective contacts, respectively, which makes the perovskite device distinct from its organic conterpart.62,63 Inorganic charge extraction layers (NiMgLiO and TiNbO2) have been used to fabricate large-area (1  cm2), high-efficiency (15%) perovskite cells, representing a potentially important step in the path toward commercialization.62 The use of oxide HTMs also allows for construction of the mesoscopic p-i-n device structure [Fig. 2(d)], in which NiO/mp-Al2O3 or c-NiO/mp-NiO are used as the HTM.47,48 The best mesoscopic p-i-n device with a nanostructured NiO film demonstrated a 17.3% efficiency.64

3.

Preparation Methods

The device performance of most thin-film solar cells is mainly determined by the film quality of the absorber. High-quality perovskite films with appropriate morphology, uniformity, phase purity, and crystallinity are essential for high-performance PV devices. To meet these quality criteria, well-controlled crystallization and engineering of the composition and interface properties of perovskite films are required. Critical issues include the deposition approach, precursor composition, processing condition, and additive control, all of which can greatly affect the crystallization and quality of the perovskite films. Focusing first on the deposition approach, the preparation processes can be categorized as follows: single-step solution deposition,26 two-step solution deposition,28 two-step vapor-assisted deposition,30 and thermal vapor deposition.29

3.1.

Single-Step Solution Deposition

Single-step solution deposition [Fig. 3(a)] is commonly used for perovskite thin film preparation due to ease of processing and low fabrication cost. Generally, organic halides [methylammonium iodide (MAI)] and lead halides (PbX2, X=I, Br, or Cl) are dissolved in gamma-butyrolactone (GBL), dimethylformamide (DMF), or dimethyl sulfoxide (DMSO) to prepare the precursor solution. The perovskite films can be prepared by spin-coating of the precursor solution followed by a postdeposition heating at 100 to 150°C. Since the perovskite tolerates composition variation,65 high-efficiency devices can be fabricated through a wide range of MAI to PbI2 precursor ratios from MAI-poor (12)66 to MAI-rich (31).58 However, it is critical to choose appropriate processing temperatures and times based on differing precursor compositions to achieve the desired crystallinity, phase, and morphology of the perovskite films.55,65,67 In addition to the choice of precursor composition and processing temperature, the environment (oxygen and humidity levels), substrate material, and deposition parameters must also be controlled. The first solid-state device prepared using the single-step solution process produced a perovskite device that exhibited 9.7% efficiency.68 After developing advanced engineering techniques (discussed in Sec. 4), a best efficiency of 19.7% has been achieved with single-step solution deposition.69

Fig. 3

Deposition methods for perovskite thin films, including (a) single-step solution deposition, (b) two-step solution deposition, (c) two-step hybrid deposition, and (d) thermal vapor deposition.

JPE_6_2_022001_f003.png

In addition to spin-coating, other solution-based deposition methods, including spray,70 doctor-blade,71 inkjet printing,72 and slot-die printing,73 have also been employed to fabricate perovskite PV devices. These techniques demonstrate the potential for large-scale roll-to-roll manufacture of perovskite solar cells. However, the efficiency of devices prepared by these methods is still lower than that of spin-coated devices due to the difficulties associated with controlling the film morphology and compositional uniformity at present.

3.2.

Two-Step Solution Deposition

The two-step solution deposition approach to preparing OMHPs was first introduced by Mitzi et al. in 1998.74 Following this pioneering work, Gratzel et al. developed a sequential deposition method [Fig. 3(b)] to prepare perovskite solar cells, which has resulted in efficiencies >15%.28 In a typical two-step solution procedure, a PbI2 seed layer is spin-coated and then converted to MAPbI3 by dipping the substrate into an MAI/isopropanol solution.28 Spin-coating has also been used to introduce MAI molecules into the PbI2 network.68 Compared with the single-step solution process, the two-step sequential deposition process results in more uniform and dense perovskite films.75 The process can be well controlled and, consequently, has been extensively used to fabricate high-efficiency devices.28,56,76,77

The two-step solution method provides a reproducible way to fabricate high-quality perovskite thin films. Through varying the MAI solution concentration, the perovskite grain size can be controlled.68 However, one of the drawbacks of the two-step solution deposition method is the trade-off between perovskite grain size and surface smoothness. Films with large perovskite grains typically exhibit poor surface coverage, which can limit the performance of devices. The other issue with this method is incomplete perovskite conversion. The conversion from PbI2 to MAPbI3 rapidly occurs as the film is dipped into the solution because the layered structure of heavy metal halide is prone to interaction with small molecules.78 Thus, a dense perovskite capping layer usually forms on the surface of PbI2 and hinders the MAI diffusion to the underlying layer, leading to incomplete perovskite conversion. These issues have been overcome by some new techniques that have been developed recently (Sec. 4), and now the champion cell efficiency using the two-step solution method has been improved to 20.2%.56

3.3.

Vapor-Assisted Solution Deposition

In one modification to the two-step solution deposition method, MAI is introduced through a vapor deposition technique rather than through solution processing [Fig. 3(c)].30 This deposition method allows better control of morphology and grain size via gas–solid crystallization and effectively avoids film delamination that can occur during liquid–solid interaction. The perovskite films prepared by this method exhibit uniform surface coverage, large grain size, and full conversion. However, the use of this method is limited because the gas–solid reaction typically required tens of hours for the full conversion, and devices prepared by this method have exhibited only 10 to 12% efficiency.30,79

3.4.

Thermal Vapor Deposition

Vapor phase deposition is widely used for fabricating high-quality semiconductor thin films with uniform thickness and composition. The thermal vapor deposition of OMHP thin films was first demonstrated by Mitzi et al. in 1999.18 After modifying the technique for dual-source thermal evaporation [Fig. 3(d)], Snaith et al. prepared the first planar heterojunction MAPbI3xClx perovskite solar cell with an efficiency that exceeded 15%.29 Similar vapor-based deposition techniques, such as sequential layer-by-layer vacuum sublimation31 and chemical vapor deposition,80 have also been developed.

The perovskite films prepared by thermal vapor deposition are extremely uniform and pinhole-free. Compared with the incomplete surface coverage that can be found for perovskite films prepared by solution processing, vapor-deposited perovskite layers can conformally coat TiO2 and PEDOT:PSS layers.29,60,81 However, both the precursor sources and the products have low thermal stability, so the vapor deposition requires precise control over temperatures during deposition. Thus, only a few research groups have reported high-efficiency devices prepared by this method.29,31,60,81,82

4.

Advanced Device Engineering

In early 2013, the state-of-the-art of perovskite solar cells prepared by various deposition techniques had demonstrated device efficiencies in the range of 12 to 15%.2830 Since then, the efficiency has improved to 18 to 20%, mainly due to the advancements of several device engineering strategies.10,83,84 These engineering strategies, which focused on controlling the precursor solution, processing condition, perovskite composition, and interface properties, lead to smooth and pinhole-free perovskite thin films consisting of large grains with good crystallinity. The combination of these advanced engineering methods has improved the optoelectronic properties of the perovskite films and, consequently, the device performance as well.

4.1.

Solvent Engineering

Single-step spin-coating is the simplest method for preparing perovskite thin films; however, it is difficult to achieve a homogeneous composition and uniform thickness over large areas. The reason for this is that single-step solution deposition using DMF and GBL solvents often results in the formation of needle- and spherical-shaped colloidal intermediates.28,85 To improve the surface morphology of spin-coated perovskite films, several precursor solution additives have been employed to suppress the formation of deleterious intermediates.

DMSO is one of the best and widely used additives.77,86 The precursor solution with added DMSO forms a uniform and flat MAI-PbI2-DMSO intermediate film when spin-coated. After a thermal treatment, the intermediate film is converted into a uniform perovskite film through a solid-state reaction. Several other additives, such as CH3NH3Cl,87 HI,88 I2,89 NH4Cl,90 H2O/HBr,91 1,8-diiodooctane,92 aminovaleric acid,93 and phosphonic acid ammonium,94 have also been used to improve the crystallinity and morphological uniformity of perovskite films.

The formation of uniform perovskite film by incorporating additives is the result of decoupling the nucleation and grain growth processes. For precursor solutions without additives, these two processes occur simultaneously. Since grain growth favors large-size nuclei (the free energy of volume expansion eclipses that of interface formation), the unbalanced growth rate leads to the formation of large perovskite grains with a significant number of voids between grains. The introduction of additives retards the crystallization kinetics of perovskite formation and results in a uniform intermediate phase film during deposition. A thermal treatment provides the energy for conversion to the perovskite phase and promotes crystal growth to form pinhole-free films.

Additive incorporation was introduced to the two-step methods after its success in single-step deposition. The precursor solution for PbI2 can be mixed with DMSO,95 H2O,96 and low concentrations of MAI76 to improve the surface coverage of the final perovskite film. As with single-step deposition, the introduction of additives results in an intermediate state that retards the rapid reaction between MAI and PbI2 and effectively avoids the formation of a dense perovskite capping layer on the surface of the PbI2 layer that hinders further conversion.

4.2.

Process Engineering

In addition to modifying the precursor solution, improved device performance has been achieved by adapting the deposition and postdeposition processes. While slowing the crystal growth kinetics has resulted in higher-quality films, the same results have been obtained by speeding the nucleation kinetics. Hot casting, in which crystallization of the perovskite film occurs immediately after a hot precursor solution is loaded onto the substrate at an elevated temperature, has been used to obtain pinhole-free perovskite films with millimeter-scale grains.97 Using this approach, the island-shaped grains rapidly integrate into a dense perovskite film with millimeter-size grains following Volmer-Weber growth.98 Devices with efficiency of 18% were fabricated using this technique.97

Another demonstration of process engineering for fabricating extremely uniform and dense perovskite films is adding an antisolvent that does not dissolve perovskite films (e.g., toluene) during the last few seconds of the spin process.77 The introduction of toluene rapidly extracts DMF from the precursor solution, which results in a rapid precipitation of perovskite before significant growth of the perovskite grains. Thus, a dense, small-grain perovskite film can form uniformly across the entire substrate surface. In addition to toluene, other antisolvents, such as diethyl ether,69 chlorobenzene, benzene, and xylene, are also effective in forming highly uniform perovskite films.99 Since the grain growth kinetics are suppressed during the deposition, this process needs an optimized thermal annealing to achieve both smooth morphology and large grain size.

The postdeposition grain growth process can also be engineered to achieve a uniform and high-quality perovskite film. Although thermal annealing helps increase grain size and improve crystallinity, it may cause decomposition of the perovskite phase and reduce surface coverage.55,65 Solvent annealing with DMF leads to recrystallization and regrowth of perovskite grains, resulting in improved crystallinity and electronic properties and enhanced device efficiency (15.6%).100 Annealing with pyridine or MAI vapor has demonstrated enhanced luminescence and carrier lifetimes, indicating the formation of high-quality absorber material and the potential for high-efficiency devices.101,102

The advanced solvent and process engineering techniques both aim to decouple the nucleation and growth processes so that the perovskite film formation can be precisely controlled. By applying one or a combination of these techniques, high-quality perovskite films with smooth morphology and large grains were prepared (Fig. 4) and devices with efficiencies of 15 to 19% were fabricated. Details on these devices are summarized in Table 1.

Fig. 4

SEM images of perovskite films prepared using various deposition techniques and advanced engineering processes, including (a) vapor deposition, (b) vapor-assisted deposition, (c) hot casting, (d) H2O additive, (e) DMSO + toluene, (f) chlorobenzene, (g) sequential deposition, and (h) solvent annealing. Reprinted with permission from Refs. 29, 30, 98, 96, 77, 99, 68, and 100.

JPE_6_2_022001_f004.png

Table 1

Summary of notable perovskite devices.

YearVOC (V)JSC (mA/cm2)FF (%)PCE (%)Device structureMethodAdvanced device engineeringRef.
20120.8917.6629.7FTO/c-TiO2/mp-TiO2/MAPbI3/spiro/AuSSS25
20120.9817.86310.9FTO/c-TiO2/mp-Al2O3/MAPbI3xClx/spiro/AuSSSAl2O3(c)26
20130.9920.07315.0FTO/c-TiO2/mp-TiO2/MAPbI3/spiro/AuTSS28
20131.0721.56715.4FTO/c-TiO2/MAPbI3xClx/spiro/AuTVD29
20130.9219.86612.1FTO/c-TiO2/MAPbI3/spiro/AuVAS30
20140.9423.36514.2FTO/c-TiO2/FAPbI3/spiro/AuSSSFA(b)33
20140.8421.16511.6FTO/c-TiO2/mpTiO2/ZrO2MAPbI3/CSSSZrO2(c)93
20141.1119.67616.5FTO/c-TiO2/mpTiO2/MAPbI3xBrx/PTAA/AuTSSDMSO(a), toluene(p)77
20141.1322.87519.3ITO-PEIE/γ-TiO2/MAPbI3xClx/spiro/AuSSSPolyethyleneimine ethoxylated (PEIE)(c), γ-TiO2(c)58
20140.9222.48217.7FTO/PEDOT:PSS/MAPbI3xClx/PCBM/AlSSSHot-casting(p)97
20151.08227319FTO/c-TiO2/mp-TiO2/FAxMA1xPbBryI1y/PTAA/AuSSSDMSO(a), toluene(p), FA and Br(b)103
20151.1020.97918.1ITO/PEDOT:PSS/MAPbI3/PCBM/AuSSSHI(a)104
20151.10227818.9ITO/PTAA/MAPbI3xClx/PCBM/C60/BCP/AlSSSMulticycle(p)61
20151.0624.77820.2FTO/c-TiO2/mpTiO2/FAxMA1xPbBryI1y/PTAA/AuTSSDMSO(a), FA and Br(b)56
20151.1421.37418.4FTO/SnO2/FAxMA1xPbBryI1y/spiro/AuSSSDMSO(a), FA and Br(b), SnO2(c) chlorobenzene(p)105
20151.0922.48019.1FTO/c-TiO2/mp-TiO2/MAPbI3/spiro/AuSSSPbI2(a)66
20151.0923.87619.7FTO/c-TiO2/mpTiO2//MAPbI3/spiro/AgSSSDMSO(a), diethyl ether(p)69
20151.0823.97519.4FTO/c-TiO2/MAPbI3/spiro/AuSSSChlorobenzene(p)106
20151.0522.27617.7ITO/Cu:NiOx/MAPbI3/C60/AgSSSToluene(p), Cu:NiOx(c)107
20151.0720.67516.2FTO/NiMgLiO/MAPbI3/PCBM/Ti(Nb)Ox/AgSSSDMSO(a), NiMgLiO(c), Ti(Nb)O(c)62
Note: SSS, single-step solution; TSS, two-step solution; TVD, thermal vapor deposition; VAS, vapor-assisted solution. Labels for advanced device engineering techniques: additive (a), processing (p), band gap (b), and contact (c).

4.3.

Band Gap Engineering

The tunability of the OMHP band gap over a wide range of the solar spectrum has led to considerable improvement in device performance. Compositional engineering of MAPbI3 perovskite can be achieved by exchanging the organic [formamidinium (FA)], metal (Sn), or halide (Br or Cl) ions. The band gap of the perovskite can be controllably tuned to cover almost the entire visible spectrum from 1.5 to 2.3 eV by introducing mixed halides (I and Br).32,33 The introduction of Br also enhances the water resistance of the perovskites.32 Partial replacement of MA by FA for the alloyed MAxFA1xPbI3 is an effective way to extend the absorption to longer wavelengths and enhance the thermal stability.108 With the (FAPbI3)1x(MAPbBr3)x absorber, solar cells with over 19% average efficiency have been fabricated with a high degree of reproducibility.56

In addition to varying the halide anions and the organic cations, the divalent metal cation may also be changed. Due to concerns about lead toxicity, lead-free perovskites, such as MASnX3, have attracted increasing attention.109111 Comparing with the Pb-based perovskite, the relatively lower band gap of MASnI3 (1.3  eV) allows absorption over a broader range and an increase of JSC from 15to21  mA/cm2.112 The efficiency (6%) and stability of the lead-free perovskite-based device, though, are not currently comparable to their Pb-based counterparts at this time.

4.4.

Contact Engineering

In addition to controlling and modifying the optoelectronic and structural properties of the absorber, as discussed above, the properties of the electron and hole collecting electrodes and their interfaces are also critical for improving perovskite device performance. The importance of the interface properties has been revealed by electron-beam–induced current investigations, which show that efficient charge separation and collection occur at the interfaces between the perovskite and both charge-selective layers.113 The choice of the ETM and HTM is important to achieve a high degree of charge selectivity while maintaining a low surface recombination to minimize energy loss at the heterojunction interfaces. Recently, a variety of ETMs and HTMs have been explored for achieving high-efficiency perovskite devices. Figure 5 plots the energy levels for several representative components of the most common perovskite solar cells.

Fig. 5

Diagram showing the energy levels, from left to right, for representative cathode, n-type (ETM), absorber, p-type (HTM), and anode materials.

JPE_6_2_022001_f005.png

Metal oxides are the most common ETMs. While TiO2 is predominant in the literature, many other materials can operate as either mesoporous or planar ETMs. Wide band gap metal oxides, such as ZnO,114 Al2O3,26,115 SrTiO3,116 SiO2, and ZrO2,93,117 have been used to fabricate devices in the mesoscopic structure. An electrically insulating mesoporous layer allows high VOC to be achieved if there is a lack of sub-band gap and surface electronic states.26 A variety of ETMs have also been used to form compact layers in the planar n-i-p structure, including ZnO,118 SnO2, 105,119 CdSe,120 CdS,121 and TiO2-graphene.122 Among them, SnO2 has been used to fabricate an 18% efficiency device, presumably due to good band alignment.105

The commonly used HTMs fall into three categories: small molecules, organic polymers, and inorganics. Small molecules, especially spiro-MeOTAD, are very commonly used as the HTM in high-efficiency perovskite PV devices. The conductive organic polymer poly(triarylamine) (PTAA) has recently emerged as a strong competitor to spiro-MeOTAD and was employed in the 20.2% efficiency perovskite device.56 Poly(3-hexylthiophene-2,5-diyl) and other organic molecules and polymers have also been used to fabricate 12 to 15% efficiency perovskite devices. A detailed review of HTMs can be found elsewhere.123 Organic HTMs are typically doped with lithiumbis(trifluoromethanesulfonyl)imide and 4-tertbutylpyridine to improve hole conductivity, doping uniformity, and device performance. Although these organic HTMs provide good carrier transport properties, which lead to high performance, high materials costs and unproven long-term stability are major impediments to industrial application. In contrast, inorganic HTMs, such as CuSCN,124,125 CuI,126 NiO,127 and Cu:NiOx,107 are promising for more cost-effective and stable performance. The highest efficiency reported to date for an inorganic material was 17.7% with Cu:NiOx as the HTM.107

It should be noted that HTM-free and ETM-free designs have also attracted attention. The HTM is not a prerequisite for perovskite solar cells when a high-quality perovskite layer with benign interface properties is presented. A high work function metal (Au or C) may help to extract holes from the perovskite absorber alone. Several groups have demonstrated HTM-free perovskite solar cells with efficiencies ranging from 5 to 12%.79,93,128130 In addition to HTM-free devices, ETM-free devices with efficiencies of 14% have been reported.131,132 In these devices, the interface properties of the TCO cathode were modified, and the perovskite solar cells were grown directly on FTO substrates without an ETM layer. Although the performance of these devices is inferior to state-of-the-art perovskite devices due to a poor charge extraction and undesired surface recombination at the interface, the designs help improve the understanding of the device physics.

5.

Notable Devices

By combining a variety of advanced techniques, several research groups have achieved high-performance perovskite PV devices with efficiencies >17% during the last two years. Over the past three years, the device performance has improved from 10%. Several notable perovskite devices reported in recent years are summarized in Table 1. Figure 6 shows a map of the critical device performance metrics. It is clear that the evolution of the state-of-the-art perovskite device efficiency was achieved by the enhancements in JSC, VOC, and fill factor (FF). Improvements in the 2012 to 2013 timeframe can be attributed to the developments of absorber preparation techniques that led to better morphology and surface coverage and contact engineering techniques that promised efficient charge separation and collection, and, consequently, higher photogenerated current density (Fig. 6). The device efficiency over this time frame was improved from 10 to 15% by optimizing the basic processing of the perovskite absorber.

Fig. 6

Efficiency mapping of recently reported state-of-the-art perovskite solar cells labeled with reference number and colored based on efficiency.

JPE_6_2_022001_f006.png

All high-performance perovskite devices share some common characteristics. First, a high JSC is typically the result of a dense and uniform perovskite film with appropriate thickness, good crystallinity, and large grain size. The high VOC is enabled by reducing intergrain and intragrain defect densities and by good interface properties between the perovskite and the selective charge collectors. FFs are typically very high, with many devices having FFs in the range of 0.75 to 0.80. Additionally, compositional engineering of the perovskite absorber contributes to better device performance. Incorporating FA extends the absorption range to wavelengths longer than 800 nm, hence enhancing the JSC by 4  mA/cm2. The introduction of Br, on the other hand, increases the bang gap of the perovskites and reduces defect density, thus improving the VOC to 1.1  V.

The highest-efficiency devices typically employ a combination of several advanced engineering techniques. For instance, the champion 20.2% device was prepared by the intermolecular exchange process involving the reaction between the PbI2-DMSO intermediate phases and the FAI-MABr contained solution.56 An extremely uniform and dense perovskite film was formed after annealing, and the device exhibited excellent performance (Fig. 7). Other devices with >18% efficiency are fabricated by spin-coating of the mixed PbI2-FAI-PbBr2-MABr precursor in the DMF/DMSO solution followed by antisolvent quenching.69,103,105

Fig. 7

SEM cross-sectional image and the (a) JV and (b) external quantum efficiency characteristics of the best-efficiency perovskite PV devices so far. Reprinted with permission from Ref. 56.

JPE_6_2_022001_f007.png

6.

Issues and Challenges

Perovskite solar cells have demonstrated high efficiency and are being investigated as a viable commercial option. However, the crucial issues and challenges that limited the commercialization of perovskite-based PV remain. Long-term device stability during operation under stressed conditions (high humidity, elevated temperature, and intense illumination) has yet to be demonstrated. The existence of the JV hysteresis limits the standardized characterization of device performance. Environmental impacts during the manufacturing, operational, and disposal phases of perovskite solar cells are unclear, leaving concerns about the toxicity and contamination associated with the water-soluble lead compounds. Although the complexity of the diverse material preparation methods and device architectures make it more difficult to address these issues, recent progress has provided insights into these issues and the corresponding material properties.

6.1.

Device Stability

One of the most important criteria for a practical solar cell is that the cell has to maintain a stable power output under a standard working condition. At present, the efficiency of perovskite devices is determined by the average of the forward and reverse scans or the steady-state power output close to the maximum power point. Although JV hysteresis may exist, the current output of most perovskite devices quickly stabilizes at the maximum power point. Such steady-state output shows the potential for sustainable power generation and is now accepted as one of the criteria to characterize perovskite PV devices.133

Although stability data up to a few hundred hours (one to four months) has been reported,25,93,133 long-term stability that is comparable to the 30-year standard of commercial PV panels has yet to be demonstrated. Early perovskite devices without encapsulation have shown stable operation up to hundreds of hours when stored in the dark and measured infrequently. However, these devices rapidly degraded after sustained exposure to sunlight.10 In addition to light exposure, elevated temperature and humidity may accelerate the degradation due to the moisture-induced decomposition of perovskite crystals.134 These stability issues, though, are being addressed by, for example, proper protective coatings. The stability of perovskite PV devices under high humidity and temperature conditions was improved by employing a moisture-resistant layer (e.g., carbon nanotubes or graphite) to prevent water ingress.93,135137 Encapsulation techniques using glass sealing or laminate plastic films have also been used to improve device stability to over 3000 h at 60°C under simulated sunlight.133 Additionally, when incorporating (FA+ and Br) ions into perovskite (FA1yMAy)Pb(I3xBrx), the thermal and moisture resistivity can be dramatically improved.32,33,103 These results indicate that perovskite PV modules with appropriate composition and encapsulation have the potential to be stable. It should be noted that good stability of perovskite PV devices has recently been demonstrated under hot outdoor conditions in Jeddah, Saudi Arabia.137 After three months of operation at 80°C, the perovskite devices demonstrated impressively stable performance without measureable degradation.

6.2.

JV Hysteresis

One of the major issues that limits advancement of perovskite solar cells is the presence of the anomalous JV hysteresis, which is observed by varying the direction and the rate of voltage sweep [Fig. 8(a)].141 Holding a perovskite device at a forward bias voltage before measurement may result in a higher efficiency than that found when the device is held at the maximum power point or when the device has been reverse biased or held at short circuit.138 Measuring the device at a rate faster than its response time may also result in varying efficiency measurements.142 The presence of JV hysteresis undermines the reporting accuracy of efficiency and may lead to questionable and erroneous device efficiencies.

Fig. 8

(a) JV hysteresis measured using different scan speeds and directions of the scan. Reprinted with permission from Ref. 138. (b) Ferroelectricity of CH3NH3PbI3 perovskite. Reprinted with permission from Ref. 139. (c) Schematic diagrams indicating the influence of ion migration in the perovskite solar cells. Reprinted with permission from Ref. 140. (d) Observation of the appearance of substantial JV hysteresis when cooling the p-i-n perovskite device to 175 K. Reprinted with permission from Ref. 153.

JPE_6_2_022001_f008.png

Parameters affecting JV hysteresis of perovskite solar cells have been investigated;143,144 however, the origins of hysteresis remain controversial. Three possible reasons, including ferroelectricity,145,146 ion migration,138,140 and unbalanced charge collection rates,104,147 have been proposed to explain the origin of JV hysteresis. All of these hypotheses are related to a transient electrical polarization as a response to the change of the external electrical field. Recent investigations have shown that the latter two are more likely the causes of the JV hysteresis.133,148

Ferroelectricity is one possible but unlikely origin for the JV hysteresis. Ferroelectricity may occur in OMHPs due to the shift of ions in the crystal away from their corresponding lattice point or the alignment of organic dipole moments. Evidence of this was observed in polarization loops of MAPbI3 thin films and piezoelectric force microscopy.149,150 However, recent reports indicate that perovskites are not ferroelectric at room temperature [Fig. 8(b)] and that the observed ferroelectric behavior is likely due to piezoelectric or electrochemical behavior.139,151

Ion migration is another possible explanation of JV hysteresis. Under an external electric field, the positive and negative ionic species will migrate to the opposite sides of the device, forming space charge regions closed to the interfaces. Accumulation of the mobile ions changes the density of free electronic charge carries and thus shifts the local quasi-Fermi level in the direction that is favorable (or unfavorable) to charge extraction under positive (or negative) bias [Fig. 8(c)]. Such ion migration has also been demonstrated in polarization-switchable perovskite devices, in which photocurrent direction could be switched by changing the voltage sweep direction.152 Recent modeling work revealed that the ion migration is accompanied by the charge traps serving as recombination centers.148 Therefore, reducing the density of mobile ions or charge traps inside the absorber and at the interfaces may alleviate the hysteresis.

Charge transfer rates at the interfaces of the perovskite absorber also strongly influence JV hysteresis. If unbalanced charge collection exists, i.e., if the charge transfer rates between perovskite and the n-/p-type selective contacts are quite different, charges will accumulate on the interface with a lower charge collection rate and build up a transient capacitance. Evidence of trapped charges was found at two interfaces in the conventional n-i-p structure,147 where the electron and hole mobilities in the ETM and HTM differ, respectively.104 Interestingly, the n-i-p device employing a thin mesoporous ETM and an HTM with desired hole mobility typically exhibits negligible hysteresis, which is likely due to the enhanced surface area for electron injection and improved hole transport, respectively. In contrast, the inverted p-i-n cells exhibit much less JV hysteresis, presumably due to a balanced charge carrier transport and surface passivation on the perovskite/fullerene interface.104 However, it was demonstrated that the so-called hysteresis-free p-i-n devices exhibit substantial JV hysteresis when the temperature is reduced to 175 K [Fig. 8(d)].153 Thus, changing the device architecture may not address the underlying mechanism of hysteresis in the perovskite materials themselves. Moreover, as the devices aged, the JV hysteresis was aggravated due to the degraded electronic quality of perovskite, especially at interfaces.138 This shows the importance of improving the stability of the perovskite and the engineering at the interfaces to prevent materials degradation. Furthermore, compositional engineering may also reduce the JV hysteresis. Unlike MAPbI3, FAPbI3 possesses an asymmetric charge transfer rate, which balances the charge extraction at either side of the perovskite and alleviates the JV hysteresis.56 Further development may show that the JV hysteresis could be reduced or eliminated.

6.3.

Toxicity and Pollution

Because most perovskite solar cells are lead-based, there are environmental concerns with the possibility of large-scale development. Recent environmental research should reduce these concerns. Ex-ante life cycle analysis and environmental impact assessment of perovskite solar cells have revealed that the lead bears very little proportion on the overall environmental impact during manufacturing process.154,155 Compared with other lead emission sources, such as mining, fossil fuels, and the manufacture of common products (batteries, plumbing, soldering, electronics, and so on), the potential lead pollution from a 1-GW perovskite PV plant is insignificant, even assuming the worst-case leakage scenario during operation.156 In fact, perovskite PV may actually be able to reduce the amount of Pb contamination in the environment by providing an opportunity to reuse it from other applications. Recently, perovskite PV devices were fabricated using lead sources recycled from used car batteries.157 Although no industrial data exist at present, these LCA results are based on the best projections of an industrial process and are likely an overestimate of the potential hazard.

7.

Future Directions

A new approach to the design of device architecture for perovskite solar cells is based on the flexible and lightweight substrates (e.g., flexible plastic foils). Such PV devices are of commercial interest for low-cost, large-scale roll-to-roll processing and applications as portable power sources and building/vehicle integrated materials. In the last two years, a great deal of effort has been made on perovskite PV devices on flexible, conductive substrates, such as poly(ethylene terephthalate),118,158161 polyethylene naphthalate,162 and Ti foils.163 A detailed review of flexible perovskite solar cells can be found elsewhere.164 The device performances on these substrates were stable after bending.160,162 Flexible perovskite PV mini-modules have demonstrated the potential to transfer laboratory-based perovskite techniques to industrial roll-to-roll processing165 and have been used to power aviation models.166 Additionally, high-transparency and colorful perovskite PV for building integration have also been demonstrated.167170

Because perovskites have a tunable band gap (Eg=1.5 to 2.3 eV) and high VOC, there is great interest in incorporating them in tandem devices with crystal silicon or CIGS cells.171 It is predicted that the ultimate efficiency of the monolithic tandem perovskite devices can exceed 35% in the future.11 Several designs of perovskite-based tandem devices have been reported, including two-terminal monolithic devices and four-terminal devices (with a light splitting component).171176 However, the overall efficiencies, 19.5% for the best four-terminal perovskite/CIGS device175 and 21% for the best two-terminal perovskite/Si device,177 are still lower than that of the state-of-the-art single-junction perovskite devices due to the electric loss at the tunneling junction and the transparent electrode. Practical fabrication of monolithic perovskite tandem devices is challenging because the device efficiency is strongly determined by the choice of the materials and the processing methods. In particular, the tunneling layer needs to be adequately adjusted so that the optical transparency and electrical conductivity can be appropriately matched to both the top and bottom cells. Additionally, the high energy associated with the sputtering of the transparent electrode (ITO or ZnO:Al) may deteriorate the perovskite layer. Therefore, the deposition process needs to be well controlled to prevent any degradation of perovskite or organic HTM layers.

8.

Summary

The last five years have witnessed a rapid development of OMHP solar cells. A variety of device architectures and material preparation methods have been developed for fabricating high-performance PV devices. Recent advances in engineering the bulk and interface properties of perovskite thin films and contacts have been tremendously effective in enhancing device performance. These advanced engineering techniques are beneficial to increase perovskite grain size and crystallinity, to improve surface coverage and film morphology, and to passivate surface and bulk defects. Further improvement of perovskite PV devices depends on a precise control of the processing of the organic and inorganic precursors and a corresponding understanding of the fundamental material properties of the perovskites. With progress in device stability, perovskite solar cells may well be a very promising technology for the future PV market.

Acknowledgments

The work was supported by the Air Force Research Laboratory, Space Vehicles Directorate (Contract No. FA9453-11-C-0253) and the National Science Foundation (Contract No. CHE-1230246).

References

1. 

D. M. Chapin, C. S. Fuller and G. L. Pearson, “A new silicon p‐n junction photocell for converting solar radiation into electrical power,” J. Appl. Phys., 25 (5), 676 –677 (1954). http://dx.doi.org/10.1063/1.1721711 JAPIAU 0021-8979 Google Scholar

2. 

Fraunhofer Institute for Solar Energy Systems, “Photovoltaic report,” (2015) https://www.ise.fraunhofer.de/de/downloads/pdf-files/aktuelles/photovoltaics-report-in-englischer-sprache.pdf December 2015). Google Scholar

3. 

M. A. Green et al., “Solar cell efficiency tables (version 47),” Prog. Photovolt., 24 (1), 3 –11 (2016). http://dx.doi.org/10.1002/pip.2728 PPHOED 1062-7995 Google Scholar

4. 

H. J. Snaith, “Perovskites: the emergence of a new era for low-cost, high-efficiency solar cells,” J. Phys. Chem. Lett., 4 (21), 3623 –3630 (2013). http://dx.doi.org/10.1021/jz4020162 JPCLCD 1948-7185 Google Scholar

5. 

N.-G. Park, “Organometal perovskite light absorbers toward a 20% efficiency low-cost solid-state mesoscopic solar cell,” J. Phys. Chem. Lett., 4 (15), 2423 –2429 (2013). http://dx.doi.org/10.1021/jz400892a JPCLCD 1948-7185 Google Scholar

6. 

R. F. Service, “Perovskite solar cells keep on surging,” Science, 344 (6183), 458 (2014). http://dx.doi.org/10.1126/science.344.6183.458 SCIEAS 0036-8075 Google Scholar

7. 

M. A. Green, A. Ho-Baillie and H. J. Snaith, “The emergence of perovskite solar cells,” Nat. Photon., 8 (7), 506 –514 (2014). http://dx.doi.org/10.1038/nphoton.2014.134 NPAHBY 1749-4885 Google Scholar

8. 

M. Gratzel, “The light and shade of perovskite solar cells,” Nat. Mater., 13 (9), 838 –842 (2014). http://dx.doi.org/10.1038/nmat4065 Google Scholar

9. 

P. Gao, M. Gratzel and M. K. Nazeeruddin, “Organohalide lead perovskites for photovoltaic applications,” Energy Environ. Sci., 7 (8), 2448 –2463 (2014). http://dx.doi.org/10.1039/C4EE00942H EESNBY 1754-5692 Google Scholar

10. 

S. D. Stranks and H. J. Snaith, “Metal-halide perovskites for photovoltaic and light-emitting devices,” Nat. Nanotechnol., 10 (5), 391 –402 (2015). http://dx.doi.org/10.1038/nnano.2015.90 NNAABX 1748-3387 Google Scholar

11. 

Q. Chen et al., “Under the spotlight: the organic-inorganic hybrid halide perovskite for optoelectronic applications,” Nano Today, 10 (3), 355 –396 (2015). http://dx.doi.org/10.1016/j.nantod.2015.04.009 NTAOCG 1748-0132 Google Scholar

12. 

T. C. Sum and N. Mathews, “Advancements in perovskite solar cells: photophysics behind the photovoltaics,” Energy Environ. Sci., 7 (8), 2518 –2534 (2014). http://dx.doi.org/10.1039/C4EE00673A EESNBY 1754-5692 Google Scholar

13. 

A. K. Chilvery et al., “Perovskites: transforming photovoltaics, a mini-review,” J. Photon. Energy, 5 (1), 057402 (2015). http://dx.doi.org/10.1117/1.JPE.5.057402 Google Scholar

14. 

V. S. Sivaram, S. D. Stranks and H. J. Snaith, “Perovskite solar cells join the major league,” Science, 350 (6263), 917 –917 (2015). http://dx.doi.org/10.1126/science.aad5891 SCAMAC 0036-8733 Google Scholar

15. 

L. J. Schmidt, “Tracking down the truth of perovski,” in 38th Rochester Mineralogical Symp. Program Notes, 31 –32 (2011). Google Scholar

16. 

M. I. Saidaminov et al., “High-quality bulk hybrid perovskite single crystals within minutes by inverse temperature crystallization,” Nat. Commun., 6 7586 (2015). http://dx.doi.org/10.1038/ncomms8586 NCAOBW 2041-1723 Google Scholar

17. 

D. B. Mitzi, “Synthesis, crystal structure, and optical and thermal properties of (C4H9NH3)2MI4 (M=Ge, Sn, Pb),” Chem. Mater., 8 (3), 791 –800 (1996). http://dx.doi.org/10.1021/cm9505097 CMATEX 0897-4756 Google Scholar

18. 

D. B. Mitzi, M. T. Prikas and K. Chondroudis, “Thin film deposition of organic-inorganic hybrid materials using a single source thermal ablation technique,” Chem. Mater., 11 (3), 542 –544 (1999). http://dx.doi.org/10.1021/cm9811139 CMATEX 0897-4756 Google Scholar

19. 

D. B. Mitzi, C. D. Dimitrakopoulos and L. L. Kosbar, “Structurally tailored organic-inorganic perovskites: optical properties and solution-processed channel materials for thin-film transistors,” Chem. Mater., 13 (10), 3728 –3740 (2001). http://dx.doi.org/10.1021/cm010105g CMATEX 0897-4756 Google Scholar

20. 

J. Calabrese et al., “Preparation and characterization of layered lead halide compounds,” J. Am. Chem. Soc., 113 (6), 2328 –2330 (1991). http://dx.doi.org/10.1021/ja00006a076 JACSAT 0002-7863 Google Scholar

21. 

K. Chondroudis and D. B. Mitzi, “Electroluminescence from an organic-inorganic perovskite incorporating a quaterthiophene dye within lead halide perovskite layers,” Chem. Mater., 11 (11), 3028 –3030 (1999). http://dx.doi.org/10.1021/cm990561t CMATEX 0897-4756 Google Scholar

22. 

D. B. Mitzi, K. Chondroudis and C. R. Kagan, “Organic-inorganic electronics,” IBM J. Res. Dev., 45 (1), 29 –45 (2001). http://dx.doi.org/10.1147/rd.451.0029 IBMJAE 0018-8646 Google Scholar

23. 

A. Kojima et al., “Organometal halide perovskites as visible-light sensitizers for photovoltaic cells,” J. Am. Chem. Soc., 131 (17), 6050 –6051 (2009). http://dx.doi.org/10.1021/ja809598r JACSAT 0002-7863 Google Scholar

24. 

“Perovskite fever,” Nat. Mater., 13 (9), 837 (2014). http://dx.doi.org/10.1038/nmat4079 NMAACR 1476-1122 Google Scholar

25. 

H.-S. Kim et al., “Lead iodide perovskite sensitized all-solid-state submicron thin film mesoscopic solar cell with efficiency exceeding 9%,” Sci. Rep., 2 591 (2012). http://dx.doi.org/10.1038/srep00591 SRCEC3 2045-2322 Google Scholar

26. 

M. M. Lee et al., “Efficient hybrid solar cells based on meso-superstructured organometal halide perovskites,” Science, 338 (6107), 643 –647 (2012). http://dx.doi.org/10.1126/science.1228604 SCIEAS 0036-8075 Google Scholar

27. 

NREL, “Solar cell efficiency chart,” (2016) http://www.nrel.gov/ncpv/images/efficiency_chart.jpg March ). 2016). Google Scholar

28. 

J. Burschka et al., “Sequential deposition as a route to high-performance perovskite-sensitized solar cells,” Nature, 499 (7458), 316 –319 (2013). http://dx.doi.org/10.1038/nature12340 Google Scholar

29. 

M. Liu, M. B. Johnston and H. J. Snaith, “Efficient planar heterojunction perovskite solar cells by vapour deposition,” Nature, 501 (7467), 395 –398 (2013). http://dx.doi.org/10.1038/nature12509 Google Scholar

30. 

Q. Chen et al., “Planar heterojunction perovskite solar cells via vapor-assisted solution process,” J. Am. Chem. Soc., 136 (2), 622 –625 (2014). http://dx.doi.org/10.1021/ja411509g JACSAT 0002-7863 Google Scholar

31. 

C.-W. Chen et al., “Efficient and uniform planar-type perovskite solar cells by simple sequential vacuum deposition,” Adv. Mater., 26 (38), 6647 –6652 (2014). http://dx.doi.org/10.1002/adma.201402461 ADVMEW 0935-9648 Google Scholar

32. 

J. H. Noh et al., “Chemical management for colorful, efficient, and stable inorganic-organic hybrid nanostructured solar cells,” Nano Lett., 13 (4), 1764 –1769 (2013). http://dx.doi.org/10.1021/nl400349b NALEFD 1530-6984 Google Scholar

33. 

G. E. Eperon et al., “Formamidinium lead trihalide: a broadly tunable perovskite for efficient planar heterojunction solar cells,” Energy Environ. Sci., 7 (3), 982 –988 (2014). http://dx.doi.org/10.1039/c3ee43822h EESNBY 1754-5692 Google Scholar

34. 

W. J. Yin, T. T. Shi and Y. F. Yan, “Unique properties of halide perovskites as possible origins of the superior solar cell performance,” Adv. Mater., 26 (27), 4653 –4658 (2014). http://dx.doi.org/10.1002/adma.201306281 ADVMEW 0935-9648 Google Scholar

35. 

S. De Wolf et al., “Organometallic halide perovskites: sharp optical absorption edge and its relation to photovoltaic performance,” J. Phys. Chem. Lett., 5 (6), 1035 –1039 (2014). http://dx.doi.org/10.1021/jz500279b JPCLCD 1948-7185 Google Scholar

36. 

Jr. C. S. Ponseca et al., “Organometal halide perovskite solar cell materials rationalized: ultrafast charge generation, high and microsecond-long balanced mobilities, and slow recombination,” J. Am. Chem. Soc., 136 (14), 5189 –5192 (2014). http://dx.doi.org/10.1021/ja412583t JACSAT 0002-7863 Google Scholar

37. 

V. D’Innocenzo et al., “Excitons versus free charges in organo-lead tri-halide perovskites,” Nat. Commun., 5 3586 (2014). http://dx.doi.org/10.1038/ncomms4586 NCAOBW 2041-1723 Google Scholar

38. 

S. D. Stranks et al., “Recombination kinetics in organic-inorganic perovskites: excitons, free charge, and subgap states,” Phys. Rev. Appl., 2 (3), 034007 (2014). http://dx.doi.org/10.1103/PhysRevApplied.2.034007 PRAHB2 2331-7019 Google Scholar

39. 

S. D. Stranks et al., “Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber,” Science, 342 (6156), 341 –344 (2013). http://dx.doi.org/10.1126/science.1243982 SCIEAS 0036-8075 Google Scholar

40. 

G. Xing et al., “Long-range balanced electron- and hole-transport lengths in organic-inorganic CH3NH3PbI3,” Science, 342 (6156), 344 –347 (2013). http://dx.doi.org/10.1126/science.1243167 SCIEAS 0036-8075 Google Scholar

41. 

C. Wehrenfennig et al., “High charge carrier mobilities and lifetimes in organolead trihalide perovskites,” Adv. Mater., 26 (10), 1584 –1589 (2014). http://dx.doi.org/10.1002/adma.201305172 ADVMEW 0935-9648 Google Scholar

42. 

T. Leijtens et al., “Electronic properties of meso-superstructured and planar organometal halide perovskite films: charge trapping, photodoping, and carrier mobility,” ACS Nano, 8 (7), 7147 –7155 (2014). http://dx.doi.org/10.1021/nn502115k ANCAC3 1936-0851 Google Scholar

43. 

W.-J. Yin, T. Shi and Y. Yan, “Unusual defect physics in CH3NH3PbI3 perovskite solar cell absorber,” Appl. Phys. Lett., 104 (6), 063903 (2014). http://dx.doi.org/10.1063/1.4864778 APPLAB 0003-6951 Google Scholar

44. 

W. Tress et al., “Predicting the open-circuit voltage of CH3NH3PbI3 perovskite solar cells using electroluminescence and photovoltaic quantum efficiency spectra: the role of radiative and non-radiative recombination,” Adv. Energy Mater., 5 (3), 1400812 (2015). http://dx.doi.org/10.1002/aenm.201400812 ADEMBC 1614-6840 Google Scholar

45. 

J. H. Im et al., “6.5% efficient perovskite quantum-dot-sensitized solar cell,” Nanoscale, 3 (10), 4088 –4093 (2011). http://dx.doi.org/10.1039/c1nr10867k NANOHL 2040-3364 Google Scholar

46. 

B. O’Regan and M. Gratzel, “A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films,” Nature, 353 (6346), 737 –740 (1991). http://dx.doi.org/10.1038/353737a0 Google Scholar

47. 

W. Chen et al., “Hybrid interfacial layer leads to solid performance improvement of inverted perovskite solar cells,” Energy Environ. Sci., 8 (2), 629 –640 (2015). http://dx.doi.org/10.1039/C4EE02833C EESNBY 1754-5692 Google Scholar

48. 

K. C. Wang et al., “P-type mesoscopic nickel oxide/organometallic perovskite heterojunction solar cells,” Sci. Rep., 4 4756 (2014). http://dx.doi.org/10.1038/srep04756 SRCEC3 2045-2322 Google Scholar

49. 

M. Lee et al., “Efficient, durable and flexible perovskite photovoltaic devices with Ag-embedded ITO as the top electrode on a metal substrate,” J. Mater. Chem. A, 3 (28), 14592 –14597 (2015). http://dx.doi.org/10.1039/C5TA03240G Google Scholar

50. 

J. Troughton et al., “Highly efficient, flexible, indium-free perovskite solar cells employing metallic substrates,” J. Mater. Chem. A, 3 (17), 9141 –9145 (2015). http://dx.doi.org/10.1039/C5TA01755F Google Scholar

51. 

J. H. Heo et al., “Efficient inorganic-organic hybrid heterojunction solar cells containing perovskite compound and polymeric hole conductors,” Nat. Photon., 7 (6), 486 –491 (2013). http://dx.doi.org/10.1038/nphoton.2013.80 NPAHBY 1749-4885 Google Scholar

52. 

J. J. Choi et al., “Structure of methylammonium lead iodide within mesoporous titanium dioxide: active material in high-performance perovskite solar cells,” Nano Lett., 14 (1), 127 –133 (2014). http://dx.doi.org/10.1021/nl403514x NALEFD 1530-6984 Google Scholar

53. 

T. Leijtens et al., “Overcoming ultraviolet light instability of sensitized TiO2 with meso-superstructured organometal tri-halide perovskite solar cells,” Nat. Commun., 4 2885 (2013). http://dx.doi.org/10.1038/ncomms3885 NCAOBW 2041-1723 Google Scholar

54. 

T. Leijtens et al., “The importance of perovskite pore filling in organometal mixed halide sensitized TiO2-based solar cells,” J. Phys. Chem. Lett., 5 (7), 1096 –1102 (2014). http://dx.doi.org/10.1021/jz500209g JPCLCD 1948-7185 Google Scholar

55. 

G. E. Eperon et al., “Morphological control for high performance, solution-processed planar heterojunction perovskite solar cells,” Adv. Funct. Mater., 24 (1), 151 –157 (2014). http://dx.doi.org/10.1002/adfm.201302090 AFMDC6 1616-301X Google Scholar

56. 

W. S. Yang et al., “High-performance photovoltaic perovskite layers fabricated through intramolecular exchange,” Science, 348 (6240), 1234 –1237 (2015). http://dx.doi.org/10.1126/science.aaa9272 SCIEAS 0036-8075 Google Scholar

57. 

E. Edri et al., “Why lead methylammonium tri-iodide perovskite-based solar cells require a mesoporous electron transporting scaffold (but not necessarily a hole conductor),” Nano Lett., 14 (2), 1000 –1004 (2014). http://dx.doi.org/10.1021/nl404454h NALEFD 1530-6984 Google Scholar

58. 

H. Zhou et al., “Interface engineering of highly efficient perovskite solar cells,” Science, 345 (6196), 542 –546 (2014). http://dx.doi.org/10.1126/science.1254050 SCIEAS 0036-8075 Google Scholar

59. 

J.-Y. Jeng et al., “CH3NH3PbI3 perovskite/fullerene planar-heterojunction hybrid solar cells,” Adv. Mater., 25 (27), 3727 –3732 (2013). http://dx.doi.org/10.1002/adma.v25.27 ADVMEW 0935-9648 Google Scholar

60. 

O. Malinkiewicz et al., “Perovskite solar cells employing organic charge-transport layers,” Nat. Photon., 8 (2), 128 –132 (2014). http://dx.doi.org/10.1038/nphoton.2013.341 NPAHBY 1749-4885 Google Scholar

61. 

Q. Dong et al., “Abnormal crystal growth in CH3NH3PbI3xCLx using a multi-cycle solution coating process,” Energy Environ. Sci., 8 (8), 2464 –2470 (2015). http://dx.doi.org/10.1039/C5EE01179E EESNBY 1754-5692 Google Scholar

62. 

W. Chen et al., “Efficient and stable large-area perovskite solar cells with inorganic charge extraction layers,” Science, 350 (6263), 944 –948 (2015). http://dx.doi.org/10.1126/science.aad1015 SCIEAS 0036-8075 Google Scholar

63. 

J. You et al., “Improved air stability of perovskite solar cells via solution-processed metal oxide transport layers,” Nat. Nanotechnol., 11 (1), 75 –81 (2016). http://dx.doi.org/10.1038/nnano.2015.230 NNAABX 1748-3387 Google Scholar

64. 

J. H. Park et al., “Efficient CH3NH3PbI3 perovskite solar cells employing nanostructured p-type nio electrode formed by a pulsed laser deposition,” Adv. Mater., 27 (27), 4013 –4019 (2015). http://dx.doi.org/10.1002/adma.201500523 ADVMEW 0935-9648 Google Scholar

65. 

Z. Song et al., “Impact of processing temperature and composition on the formation of methylammonium lead iodide perovskites,” Chem. Mater., 27 (13), 4612 –4619 (2015). http://dx.doi.org/10.1021/acs.chemmater.5b01017 CMATEX 0897-4756 Google Scholar

66. 

C. Roldan-Carmona et al., “High efficiency methylammonium lead triiodide perovskite solar cells: the relevance of non-stoichiometric precursors,” Energy Environ. Sci., 8 (12), 3550 –3556 (2015). http://dx.doi.org/10.1039/C5EE02555A EESNBY 1754-5692 Google Scholar

67. 

Q. Wang et al., “Large fill-factor bilayer iodine perovskite solar cells fabricated by a low-temperature solution-process,” Energy Environ. Sci., 7 (7), 2359 –2365 (2014). http://dx.doi.org/10.1039/C4EE00233D EESNBY 1754-5692 Google Scholar

68. 

J.-H. Im et al., “Growth of CH3NH3PbI3 cuboids with controlled size for high-efficiency perovskite solar cells,” Nat. Nanotechnol., 9 (11), 927 –932 (2014). http://dx.doi.org/10.1038/nnano.2014.181 NNAABX 1748-3387 Google Scholar

69. 

N. Ahn et al., “Highly reproducible perovskite solar cells with average efficiency of 18.3% and best efficiency of 19.7% fabricated via Lewis base adduct of lead(II) iodide,” J. Am. Chem. Soc., 137 (27), 8696 –8699 (2015). http://dx.doi.org/10.1021/jacs.5b04930 JACSAT 0002-7863 Google Scholar

70. 

A. T. Barrows et al., “Efficient planar heterojunction mixed-halide perovskite solar cells deposited via spray-deposition,” Energy Environ. Sci., 7 (9), 2944 –2950 (2014). http://dx.doi.org/10.1039/C4EE01546K EESNBY 1754-5692 Google Scholar

71. 

Y. H. Deng et al., “Scalable fabrication of efficient organolead trihalide perovskite solar cells with doctor-bladed active layers,” Energy Environ. Sci., 8 (5), 1544 –1550 (2015). http://dx.doi.org/10.1039/C4EE03907F EESNBY 1754-5692 Google Scholar

72. 

S.-G. Li et al., “Inkjet printing of CH3NH3PbI3 on a mesoscopic TiO2 film for highly efficient perovskite solar cells,” J. Mater. Chem. A, 3 (17), 9092 –9097 (2015). http://dx.doi.org/10.1039/C4TA05675B Google Scholar

73. 

K. Hwang et al., “Toward large scale roll-to-roll production of fully printed perovskite solar cells,” Adv. Mater., 27 (7), 1241 –1247 (2015). http://dx.doi.org/10.1002/adma.201404598 ADVMEW 0935-9648 Google Scholar

74. 

K. N. Liang, D. B. Mitzi and M. T. Prikas, “Synthesis and characterization of organic-inorganic perovskite thin films prepared using a versatile two-step dipping technique,” Chem. Mater., 10 (1), 403 –411 (1998). http://dx.doi.org/10.1021/cm970568f CMATEX 0897-4756 Google Scholar

75. 

Z. Song et al., “Spatially resolved characterization of solution processed perovskite solar cells using the lbic technique,” in IEEE 42nd Photovoltaic Specialist Conf., 1 –5 (2015). http://dx.doi.org/10.1109/PVSC.2015.7355676 Google Scholar

76. 

T. Zhang et al., “Controllable sequential deposition of planar CH3NH3PbI3 perovskite films via adjustable volume expansion,” Nano Lett., 15 (6), 3959 –3963 (2015). http://dx.doi.org/10.1021/acs.nanolett.5b00843 NALEFD 1530-6984 Google Scholar

77. 

N. J. Jeon et al., “Solvent engineering for high-performance inorganic-organic hybrid perovskite solar cells,” Nat. Mater., 13 (9), 897 –903 (2014). http://dx.doi.org/10.1038/nmat4014 NMAACR 1476-1122 Google Scholar

78. 

V. Nicolosi et al., “Liquid exfoliation of layered materials,” Science, 340 (6139), 1226419 (2013). http://dx.doi.org/10.1126/science.1226419 SCIEAS 0036-8075 Google Scholar

79. 

F. Hao et al., “Controllable perovskite crystallization at a gas-solid interface for hole conductor-free solar cells with steady power conversion efficiency over 10%,” J. Am. Chem. Soc., 136 (46), 16411 –16419 (2014). http://dx.doi.org/10.1021/ja509245x JACSAT 0002-7863 Google Scholar

80. 

M. M. Tavakoli et al., “Fabrication of efficient planar perovskite solar cells using a one-step chemical vapor deposition method,” Sci. Rep., 5 14083 (2015). http://dx.doi.org/10.1038/srep14083 SRCEC3 2045-2322 Google Scholar

81. 

Q. Lin et al., “Electro-optics of perovskite solar cells,” Nat. Photon., 9 (2), 106 –112 (2015). http://dx.doi.org/10.1038/nphoton.2014.284 NPAHBY 1749-4885 Google Scholar

82. 

D. Zhao et al., “Annealing-free efficient vacuum-deposited planar perovskite solar cells with evaporated fullerenes as electron-selective layers,” Nano Energy, 19 88 –97 (2016). http://dx.doi.org/10.1016/j.nanoen.2015.11.008 Google Scholar

83. 

T. Salim et al., “Perovskite-based solar cells: impact of morphology and device architecture on device performance,” J. Mater. Chem. A, 3 (17), 8943 –8969 (2015). http://dx.doi.org/10.1039/C4TA05226A Google Scholar

84. 

Y. Zhao and K. Zhu, “Solution chemistry engineering toward high-efficiency perovskite solar cells,” J. Phys. Chem. Lett., 5 (23), 4175 –4186 (2014). http://dx.doi.org/10.1021/jz501983v JPCLCD 1948-7185 Google Scholar

85. 

K. Yan et al., “Hybrid halide perovskite solar cell precursors: colloidal chemistry and coordination engineering behind device processing for high efficiency,” J. Am. Chem. Soc., 137 (13), 4460 –4468 (2015). http://dx.doi.org/10.1021/jacs.5b00321 JACSAT 0002-7863 Google Scholar

86. 

W. Li et al., “Controllable grain morphology of perovskite absorber film by molecular self-assembly toward efficient solar cell exceeding 17%,” J. Am. Chem. Soc., 137 (32), 10399 –10405 (2015). http://dx.doi.org/10.1021/jacs.5b06444 JACSAT 0002-7863 Google Scholar

87. 

Y. Zhao and K. Zhu, “Efficient planar perovskite solar cells based on 1.8 eV band gap CH3NH3PbI32Br nanosheets via thermal decomposition,” J. Am. Chem. Soc., 136 (35), 12241 –12244 (2014). http://dx.doi.org/10.1021/ja5071398 JACSAT 0002-7863 Google Scholar

88. 

J. H. Heo et al., “Planar CH3NH3PbI3 perovskite solar cells with constant 17.2% average power conversion efficiency irrespective of the scan rate,” Adv. Mater., 27 (22), 3424 –3430 (2015). http://dx.doi.org/10.1002/adma.201500048 ADVMEW 0935-9648 Google Scholar

89. 

H. Tsai et al., “Optimizing composition and morphology for large-grain perovskite solar cells via chemical control,” Chem. Mater., 27 (16), 5570 –5576 (2015). http://dx.doi.org/10.1021/acs.chemmater.5b02378 CMATEX 0897-4756 Google Scholar

90. 

Y. Chen, Y. Zhao and Z. Liang, “Non-thermal annealing fabrication of efficient planar perovskite solar cells with inclusion of NH4Cl,” Chem. Mater., 27 (5), 1448 –1451 (2015). http://dx.doi.org/10.1021/acs.chemmater.5b00041 CMATEX 0897-4756 Google Scholar

91. 

J. H. Heo, D. H. Song and S. H. Im, “Planar CH3NH3PbBr3 hybrid solar cells with 10.4% power conversion efficiency, fabricated by controlled crystallization in the spin-coating process,” Adv. Mater., 26 (48), 8179 –8183 (2014). http://dx.doi.org/10.1002/adma.201403140 ADVMEW 0935-9648 Google Scholar

92. 

P. W. Liang et al., “Additive enhanced crystallization of solution-processed perovskite for highly efficient planar-heterojunction solar cells,” Adv. Mater., 26 (22), 3748 –3754 (2014). http://dx.doi.org/10.1002/adma.v26.22 ADVMEW 0935-9648 Google Scholar

93. 

A. Mei et al., “A hole-conductor-free, fully printable mesoscopic perovskite solar cell with high stability,” Science, 345 (6194), 295 –298 (2014). http://dx.doi.org/10.1126/science.1254763 SCIEAS 0036-8075 Google Scholar

94. 

X. Li et al., “Improved performance and stability of perovskite solar cells by crystal crosslinking with alkylphosphonic acid ω-ammonium chlorides,” Nat. Chem., 7 (9), 703 –711 (2015). http://dx.doi.org/10.1038/nchem.2324 NCAHBB 1755-4330 Google Scholar

95. 

Y. Wu et al., “Retarding the crystallization of PbI2 for highly reproducible planar-structured perovskite solar cells via sequential deposition,” Energy Environ. Sci., 7 (9), 2934 –2938 (2014). http://dx.doi.org/10.1039/C4EE01624F EESNBY 1754-5692 Google Scholar

96. 

C. G. Wu et al., “High efficiency stable inverted perovskite solar cells without current hysteresis,” Energy Environ. Sci., 8 (9), 2725 –2733 (2015). http://dx.doi.org/10.1039/C5EE00645G EESNBY 1754-5692 Google Scholar

97. 

W. Nie et al., “High-efficiency solution-processed perovskite solar cells with millimeter-scale grains,” Science, 347 (6221), 522 –525 (2015). http://dx.doi.org/10.1126/science.aaa0472 SCIEAS 0036-8075 Google Scholar

98. 

Y. C. Zheng et al., “Thermal-induced Volmer-Weber growth behavior for planar heterojunction perovskites solar cells,” Chem. Mater., 27 (14), 5116 –5121 (2015). http://dx.doi.org/10.1021/acs.chemmater.5b01924 CMATEX 0897-4756 Google Scholar

99. 

M. Xiao et al., “A fast deposition-crystallization procedure for highly efficient lead iodide perovskite thin-film solar cells,” Angew. Chem., 53 (37), 9898 –9903 (2014). http://dx.doi.org/10.1002/anie.201405334 Google Scholar

100. 

Z. Xiao et al., “Solvent annealing of perovskite-induced crystal growth for photovoltaic-device efficiency enhancement,” Adv. Mater., 26 (37), 6503 –6509 (2014). http://dx.doi.org/10.1002/adma.201401685 ADVMEW 0935-9648 Google Scholar

101. 

D. W. de Quilettes et al., “Impact of microstructure on local carrier lifetime in perovskite solar cells,” Science, 348 (6235), 683 –686 (2015). http://dx.doi.org/10.1126/science.aaa5333 SCIEAS 0036-8075 Google Scholar

102. 

B. S. Tosun and H. W. Hillhouse, “Enhanced carrier lifetimes of pure iodide hybrid perovskite via vapor-equilibrated re-growth (VERG),” J. Phys. Chem. Lett., 6 (13), 2503 –2508 (2015). http://dx.doi.org/10.1021/acs.jpclett.5b00842 JPCLCD 1948-7185 Google Scholar

103. 

N. J. Jeon et al., “Compositional engineering of perovskite materials for high-performance solar cells,” Nature, 517 (7535), 476 –480 (2015). http://dx.doi.org/10.1038/nature14133 Google Scholar

104. 

J. H. Heo et al., “Hysteresis-less inverted CH3NH3PbI3 planar perovskite hybrid solar cells with 18.1% power conversion efficiency,” Energy Environ. Sci., 8 (5), 1602 –1608 (2015). http://dx.doi.org/10.1039/C5EE00120J EESNBY 1754-5692 Google Scholar

105. 

J. P. Correa Baena et al., “Highly efficient planar perovskite solar cells through band alignment engineering,” Energy Environ. Sci., 8 (10), 2928 –2934 (2015). http://dx.doi.org/10.1039/C5EE02608C EESNBY 1754-5692 Google Scholar

106. 

H. D. Kim et al., “Photovoltaic performance of perovskite solar cells with different grain sizes,” Adv. Mater., 28 (5), 917 –922 (2016). http://dx.doi.org/10.1002/adma.201504144 ADVMEW 0935-9648 Google Scholar

107. 

J. W. Jung, C.-C. Chueh and A. K. Y. Jen, “A low-temperature, solution-processable, Eu-doped nickel oxide hole-transporting layer via the combustion method for high-performance thin-film perovskite solar cells,” Adv. Mater., 27 (47), 7874 –7880 (2015). http://dx.doi.org/10.1002/adma.201503298 ADVMEW 0935-9648 Google Scholar

108. 

N. Pellet et al., “Mixed-organic-cation perovskite photovoltaics for enhanced solar-light harvesting,” Angew. Chem., 53 (12), 3151 –3157 (2014). http://dx.doi.org/10.1002/anie.201309361 Google Scholar

109. 

N. K. Noel et al., “Lead-free organic-inorganic tin halide perovskites for photovoltaic applications,” Energy Environ. Sci., 7 (9), 3061 –3068 (2014). http://dx.doi.org/10.1039/C4EE01076K EESNBY 1754-5692 Google Scholar

110. 

F. Hao et al., “Lead-free solid-state organic-inorganic halide perovskite solar cells,” Nat. Photon., 8 (6), 489 –494 (2014). http://dx.doi.org/10.1038/nphoton.2014.82 NPAHBY 1749-4885 Google Scholar

111. 

Y. Ogomi et al., “CH3NH3SnxPb(1x)I3 perovskite solar cells covering up to 1060 nm,” J. Phys. Chem. Lett., 5 (6), 1004 –1011 (2014). http://dx.doi.org/10.1021/jz5002117 JPCLCD 1948-7185 Google Scholar

112. 

F. Hao et al., “Solvent-mediated crystallization of CH3NH3SnI3 films for heterojunction depleted perovskite solar cells,” J. Am. Chem. Soc., 137 (35), 11445 –11452 (2015). http://dx.doi.org/10.1021/jacs.5b06658 JACSAT 0002-7863 Google Scholar

113. 

H. S. Kim et al., “Mechanism of carrier accumulation in perovskite thin-absorber solar cells,” Nat. Commun., 4 7 (2013). http://dx.doi.org/10.1038/ncomms3242 NCAOBW 2041-1723 Google Scholar

114. 

D. Y. Son et al., “11% efficient perovskite solar cell based on Zno nanorods: an effective charge collection system,” J. Phys. Chem. C, 118 (30), 16567 –16573 (2014). http://dx.doi.org/10.1021/jp412407j JPCCCK 1932-7447 Google Scholar

115. 

J. M. Ball et al., “Low-temperature processed meso-superstructured to thin-film perovskite solar cells,” Energy Environ. Sci., 6 (6), 1739 –1743 (2013). http://dx.doi.org/10.1039/c3ee40810h EESNBY 1754-5692 Google Scholar

116. 

A. Bera et al., “Perovskite oxide SrTiO3 as an efficient electron transporter for hybrid perovskite solar cells,” J. Phys. Chem. C, 118 (49), 28494 –28501 (2014). http://dx.doi.org/10.1021/jp509753p JPCCCK 1932-7447 Google Scholar

117. 

D. Q. Bi et al., “Using a two-step deposition technique to prepare perovskite (CH3NH3PbI3) for thin film solar cells based on ZrO2 and TiO2 mesostructures,” RSC Adv., 3 (41), 18762 –18766 (2013). http://dx.doi.org/10.1039/c3ra43228a Google Scholar

118. 

D. Liu and T. L. Kelly, “Perovskite solar cells with a planar heterojunction structure prepared using room-temperature solution processing techniques,” Nat. Photon., 8 (2), 133 –138 (2014). http://dx.doi.org/10.1038/nphoton.2013.342 NPAHBY 1749-4885 Google Scholar

119. 

W. Ke et al., “Low-temperature solution-processed tin oxide as an alternative electron transporting layer for efficient perovskite solar cells,” J. Am. Chem. Soc., 137 (21), 6730 –6733 (2015). http://dx.doi.org/10.1021/jacs.5b01994 JACSAT 0002-7863 Google Scholar

120. 

L. Wang et al., “Low temperature solution processed planar heterojunction perovskite solar cells with a CDSE nanocrystal as an electron transport/extraction layer,” J. Mater. Chem. C, 2 (43), 9087 –9090 (2014). http://dx.doi.org/10.1039/C4TC01875C Google Scholar

121. 

J. Liu et al., “Low-temperature, solution processed metal sulfide as an electron transport layer for efficient planar perovskite solar cells,” J. Mater. Chem. A, 3 (22), 11750 –11755 (2015). http://dx.doi.org/10.1039/C5TA01200G Google Scholar

122. 

J. T. W. Wang et al., “Low-temperature processed electron collection layers of graphene/TiO2 nanocomposites in thin film perovskite solar cells,” Nano Lett., 14 (2), 724 –730 (2014). http://dx.doi.org/10.1021/nl403997a NALEFD 1530-6984 Google Scholar

123. 

Z. Yu and L. Sun, “Recent progress on hole-transporting materials for emerging organometal halide perovskite solar cells,” Adv. Energy Mater., 5 (12), 1500213 (2015). http://dx.doi.org/10.1002/aenm.201500213 ADEMBC 1614-6840 Google Scholar

124. 

P. Qin et al., “Inorganic hole conductor-based lead halide perovskite solar cells with 12.4% conversion efficiency,” Nat. Commun., 5 3834 (2014). http://dx.doi.org/10.1038/ncomms4834 NCAOBW 2041-1723 Google Scholar

125. 

S. Ye et al., “CuSCN-based inverted planar perovskite solar cell with an average PCE of 15.6%,” Nano Lett., 15 (6), 3723 –3728 (2015). http://dx.doi.org/10.1021/acs.nanolett.5b00116 NALEFD 1530-6984 Google Scholar

126. 

J. A. Christians, R. C. M. Fung and P. V. Kamat, “An inorganic hole conductor for organo-lead halide perovskite solar cells. Improved hole conductivity with copper iodide,” J. Am. Chem. Soc., 136 (2), 758 –764 (2014). http://dx.doi.org/10.1021/ja411014k JACSAT 0002-7863 Google Scholar

127. 

X. Xu et al., “Hole selective NiO contact for efficient perovskite solar cells with carbon electrode,” Nano Lett., 15 (4), 2402 –2408 (2015). http://dx.doi.org/10.1021/nl504701y NALEFD 1530-6984 Google Scholar

128. 

L. Etgar et al., “Mesoscopic CH3NH3PbI3/TiO2 heterojunction solar cells,” J. Am. Chem. Soc., 134 (42), 17396 –17399 (2012). http://dx.doi.org/10.1021/ja307789s JACSAT 0002-7863 Google Scholar

129. 

W. A. Laban and L. Etgar, “Depleted hole conductor-free lead halide iodide heterojunction solar cells,” Energy Environ. Sci., 6 (11), 3249 –3253 (2013). http://dx.doi.org/10.1039/c3ee42282h EESNBY 1754-5692 Google Scholar

130. 

J. Shi et al., “Hole-conductor-free perovskite organic lead iodide heterojunction thin-film solar cells: high efficiency and junction property,” Appl. Phys. Lett., 104 (6), 063901 (2014). http://dx.doi.org/10.1063/1.4864638 APPLAB 0003-6951 Google Scholar

131. 

W. Ke et al., “Efficient hole-blocking layer-free planar halide perovskite thin-film solar cells,” Nat. Commun., 6 6700 (2015). http://dx.doi.org/10.1038/ncomms7700 NCAOBW 2041-1723 Google Scholar

132. 

D. Liu, J. Yang and T. L. Kelly, “Compact layer free perovskite solar cells with 13.5% efficiency,” J. Am. Chem. Soc., 136 (49), 17116 –17122 (2014). http://dx.doi.org/10.1021/ja508758k JACSAT 0002-7863 Google Scholar

133. 

T. Leijtens et al., “Stability of metal halide perovskite solar cells,” Adv. Energy Mater., 5 1500963 (2015). http://dx.doi.org/10.1002/aenm.201500963 ADEMBC 1614-6840 Google Scholar

134. 

J. M. Frost et al., “Atomistic origins of high-performance in hybrid halide perovskite solar cells,” Nano Lett., 14 (5), 2584 –2590 (2014). http://dx.doi.org/10.1021/nl500390f NALEFD 1530-6984 Google Scholar

135. 

S. N. Habisreutinger et al., “Carbon nanotube/polymer composites as a highly stable hole collection layer in perovskite solar cells,” Nano Lett., 14 (10), 5561 –5568 (2014). http://dx.doi.org/10.1021/nl501982b NALEFD 1530-6984 Google Scholar

136. 

Z. Song et al., “Investigation of degradation mechanisms of perovskite-based photovoltaic devices using laser beam induced current mapping,” Proc. SPIE, 9561 956107 (2015). http://dx.doi.org/10.1117/12.2195789 PSISDG 0277-786X Google Scholar

137. 

X. Li et al., “Outdoor performance and stability under elevated temperatures and long-term light soaking of triple-layer mesoporous perovskite photovoltaics,” Energy Technol., 3 (6), 551 –555 (2015). http://dx.doi.org/10.1002/ente.201500045 ENTED9 Google Scholar

138. 

W. Tress et al., “Understanding the rate-dependent J-V hysteresis, slow time component, and aging in CH3NH3PbI3 perovskite solar cells: the role of a compensated electric field,” Energy Environ. Sci., 8 (3), 995 –1004 (2015). http://dx.doi.org/10.1039/C4EE03664F EESNBY 1754-5692 Google Scholar

139. 

Z. Fan et al., “Ferroelectricity of CH3NH3PbI3 perovskite,” J. Phys. Chem. Lett., 6 (7), 1155 –1161 (2015). http://dx.doi.org/10.1021/acs.jpclett.5b00389 JPCLCD 1948-7185 Google Scholar

140. 

C. Eames et al., “Ionic transport in hybrid lead iodide perovskite solar cells,” Nat. Commun., 6 7497 (2015). http://dx.doi.org/10.1038/ncomms8497 NCAOBW 2041-1723 Google Scholar

141. 

H. J. Snaith et al., “Anomalous hysteresis in perovskite solar cells,” J. Phys. Chem. Lett., 5 (9), 1511 –1515 (2014). http://dx.doi.org/10.1021/jz500113x JPCLCD 1948-7185 Google Scholar

142. 

E. L. Unger et al., “Hysteresis and transient behavior in current-voltage measurements of hybrid-perovskite absorber solar cells,” Energy Environ. Sci., 7 (11), 3690 –3698 (2014). http://dx.doi.org/10.1039/C4EE02465F EESNBY 1754-5692 Google Scholar

143. 

H.-S. Kim and N.-G. Park, “Parameters affecting I-V hysteresis of CH3NH3PbI3 perovskite solar cells: effects of perovskite crystal size and mesoporous TiO2 layer,” J. Phys. Chem. Lett., 5 (17), 2927 –2934 (2014). http://dx.doi.org/10.1021/jz501392m JPCLCD 1948-7185 Google Scholar

144. 

Y. Shao et al., “Origin and elimination of photocurrent hysteresis by fullerene passivation in CH3NH3PbI3 planar heterojunction solar cells,” Nat. Commun., 5 5784 (2014). http://dx.doi.org/10.1038/ncomms6784 NCAOBW 2041-1723 Google Scholar

145. 

J. Wei et al., “Hysteresis analysis based on the ferroelectric effect in hybrid perovskite solar cells,” J. Phys. Chem. Lett., 5 (21), 3937 –3945 (2014). http://dx.doi.org/10.1021/jz502111u JPCLCD 1948-7185 Google Scholar

146. 

J. M. Frost, K. T. Butler and A. Walsh, “Molecular ferroelectric contributions to anomalous hysteresis in hybrid perovskite solar cells,” APL Mater., 2 (8), 081506 (2014). http://dx.doi.org/10.1063/1.4890246 Google Scholar

147. 

V. W. Bergmann et al., “Real-space observation of unbalanced charge distribution inside a perovskite-sensitized solar cell,” Nat. Commun., 5 5001 (2014). http://dx.doi.org/10.1038/ncomms6001 NCAOBW 2041-1723 Google Scholar

148. 

S. van Reenen, M. Kemerink and H. J. Snaith, “Modeling anomalous hysteresis in perovskite solar cells,” J. Phys. Chem. Lett., 6 (19), 3808 –3814 (2015). http://dx.doi.org/10.1021/acs.jpclett.5b01645 JPCLCD 1948-7185 Google Scholar

149. 

J. Beilsten-Edmands et al., “Non-ferroelectric nature of the conductance hysteresis in CH3NH3PbI3 perovskite-based photovoltaic devices,” Appl. Phys. Lett., 106 (17), 173502 (2015). http://dx.doi.org/10.1063/1.4919109 APPLAB 0003-6951 Google Scholar

150. 

B. Chen et al., “Ferroelectric solar cells based on inorganic-organic hybrid perovskites,” J. Mater. Chem. A, 3 (15), 7699 –7705 (2015). http://dx.doi.org/10.1039/C5TA01325A Google Scholar

151. 

M. Coll et al., “Polarization switching and light-enhanced piezoelectricity in lead halide perovskites,” J. Phys. Chem. Lett., 6 (8), 1408 –1413 (2015). http://dx.doi.org/10.1021/acs.jpclett.5b00502 JPCLCD 1948-7185 Google Scholar

152. 

Z. Xiao et al., “Giant switchable photovoltaic effect in organometal trihalide perovskite devices,” Nat. Mater., 14 (2), 193 –198 (2015). http://dx.doi.org/10.1038/nmat4150 NMAACR 1476-1122 Google Scholar

153. 

D. Bryant et al., “Observable hysteresis at low temperature in ‘hysteresis free’ organic-inorganic lead halide perovskite solar cells,” J. Phys. Chem. Lett., 6 (16), 3190 –3194 (2015). http://dx.doi.org/10.1021/acs.jpclett.5b01381 JPCLCD 1948-7185 Google Scholar

154. 

J. Gong, S. B. Darling and F. You, “Perovskite photovoltaics: life-cycle assessment of energy and environmental impacts,” Energy Environ. Sci., 8 (7), 1953 –1968 (2015). http://dx.doi.org/10.1039/C5EE00615E EESNBY 1754-5692 Google Scholar

155. 

N. Espinosa et al., “Solution and vapour deposited lead perovskite solar cells: ecotoxicity from a life cycle assessment perspective,” Sol. Energy Mater. Sol. Cells, 137 303 –310 (2015). http://dx.doi.org/10.1016/j.solmat.2015.02.013 SEMCEQ 0927-0248 Google Scholar

156. 

B. Hailegnaw et al., “Rain on methylammonium lead iodide based perovskites: possible environmental effects of perovskite solar cells,” J. Phys. Chem. Lett., 6 (9), 1543 –1547 (2015). http://dx.doi.org/10.1021/acs.jpclett.5b00504 JPCLCD 1948-7185 Google Scholar

157. 

P. Y. Chen et al., “Environmentally responsible fabrication of efficient perovskite solar cells from recycled car batteries,” Energy Environ. Sci., 7 (11), 3659 –3665 (2014). http://dx.doi.org/10.1039/C4EE00965G EESNBY 1754-5692 Google Scholar

158. 

J. You et al., “Low-temperature solution-processed perovskite solar cells with high efficiency and flexibility,” ACS Nano, 8 (2), 1674 –1680 (2014). http://dx.doi.org/10.1021/nn406020d ANCAC3 1936-0851 Google Scholar

159. 

P. Docampo et al., “Efficient organometal trihalide perovskite planar-heterojunction solar cells on flexible polymer substrates,” Nat. Commun., 4 2761 (2013). http://dx.doi.org/10.1038/ncomms3761 NCAOBW 2041-1723 Google Scholar

160. 

C. Roldan-Carmona et al., “Flexible high efficiency perovskite solar cells,” Energy Environ. Sci., 7 (3), 994 –997 (2014). http://dx.doi.org/10.1039/C3EE43619E EESNBY 1754-5692 Google Scholar

161. 

D. Bryant et al., “A transparent conductive adhesive laminate electrode for high-efficiency organic-inorganic lead halide perovskite solar cells,” Adv. Mater., 26 (44), 7499 –7504 (2014). http://dx.doi.org/10.1002/adma.201403939 ADVMEW 0935-9648 Google Scholar

162. 

B. J. Kim et al., “Highly efficient and bending durable perovskite solar cells: toward a wearable power source,” Energy Environ. Sci., 8 (3), 916 –921 (2015). http://dx.doi.org/10.1039/C4EE02441A EESNBY 1754-5692 Google Scholar

163. 

X. Y. Wang et al., “TiO2 nanotube arrays based flexible perovskite solar cells with transparent carbon nanotube electrode,” Nano Energy, 11 728 –735 (2015). http://dx.doi.org/10.1016/j.nanoen.2014.11.042 Google Scholar

164. 

Y. Wang et al., “High-efficiency flexible solar cells based on organometal halide perovskites,” Adv. Mater., (2015). http://dx.doi.org/10.1002/adma.201504260 ADVMEW 0935-9648 Google Scholar

165. 

F. Di Giacomo et al., “Flexible perovskite photovoltaic modules and solar cells based on atomic layer deposited compact layers and uv-irradiated TiO2 scaffolds on plastic substrates,” Adv. Energy Mater., 5 (8), 1401808 (2015). http://dx.doi.org/10.1002/aenm.201401808 ADEMBC 1614-6840 Google Scholar

166. 

M. Kaltenbrunner et al., “Flexible high power-per-weight perovskite solar cells with chromium oxide-metal contacts for improved stability in air,” Nat. Mater., 14 (10), 1032 –1039 (2015). http://dx.doi.org/10.1038/nmat4388 NMAACR 1476-1122 Google Scholar

167. 

G. E. Eperon et al., “Neutral color semitransparent microstructured perovskite solar cells,” ACS Nano, 8 (1), 591 –598 (2014). http://dx.doi.org/10.1021/nn4052309 ANCAC3 1936-0851 Google Scholar

168. 

E. Della Gaspera et al., “Ultra-thin high efficiency semitransparent perovskite solar cells,” Nano Energy, 13 249 –257 (2015). http://dx.doi.org/10.1016/j.nanoen.2015.02.028 Google Scholar

169. 

A. Cannavale et al., “Perovskite photovoltachromic cells for building integration,” Energy Environ. Sci., 8 (5), 1578 –1584 (2015). http://dx.doi.org/10.1039/C5EE00896D EESNBY 1754-5692 Google Scholar

170. 

W. Zhang et al., “Highly efficient perovskite solar cells with tunable structural color,” Nano Lett., 15 (3), 1698 –1702 (2015). http://dx.doi.org/10.1021/nl504349z NALEFD 1530-6984 Google Scholar

171. 

C. D. Bailie et al., “Semi-transparent perovskite solar cells for tandems with silicon and CIGS,” Energy Environ. Sci., 8 (3), 956 –963 (2015). http://dx.doi.org/10.1039/C4EE03322A EESNBY 1754-5692 Google Scholar

172. 

P. Loper et al., “Organic-inorganic halide perovskite/crystalline silicon four-terminal tandem solar cells,” Phys. Chem. Chem. Phys., 17 (3), 1619 –1629 (2015). http://dx.doi.org/10.1039/C4CP03788J PPCPFQ 1463-9076 Google Scholar

173. 

C. C. Chen et al., “Perovskite/polymer monolithic hybrid tandem solar cells utilizing a low-temperature, full solution process,” Mater. Horiz., 2 (2), 203 –211 (2015). http://dx.doi.org/10.1039/C4MH00237G Google Scholar

174. 

J. P. Mailoa et al., “A 2-terminal perovskite/silicon multijunction solar cell enabled by a silicon tunnel junction,” Appl. Phys. Lett., 106 (12), 121105 (2015). http://dx.doi.org/10.1063/1.4914179 APPLAB 0003-6951 Google Scholar

175. 

L. Kranz et al., “High-efficiency polycrystalline thin film tandem solar cells,” J. Phys. Chem. Lett., 6 (14), 2676 –2681 (2015). http://dx.doi.org/10.1021/acs.jpclett.5b01108 JPCLCD 1948-7185 Google Scholar

176. 

S. Albrecht et al., “Monolithic perovskite/silicon-heterojunction tandem solar cells processed at low temperature,” Energy Environ. Sci., 9 81 –88 (2016). http://dx.doi.org/10.1039/C5EE02965A EESNBY 1754-5692 Google Scholar

177. 

J. Werner et al., “Efficient monolithic perovskite/silicon tandem solar cell with cell area >1  cm2,” J. Phys. Chem. Lett., 7 (1), 161 –166 (2016). http://dx.doi.org/10.1021/acs.jpclett.5b02686 JPCLCD 1948-7185 Google Scholar

Biography

Zhaoning Song received his BS degree in physics from Xiamen University, China, in 2009. He is currently a PhD student in Dr. Michael Heben’s group at the University of Toledo. His research interests include solution processing of thin-film photovoltaics and nanomaterials for optoelectronic applications. The goal of his research is to develop more cost-effective alternatives to conventional PV technologies and to explore underlying physical processes of preparing the materials.

Suneth C. Watthage is a PhD candidate in the Department of Physics of the University of Toledo in the concentration of material science. He received his BS degree in engineering physics from the University of Colombo, Sri Lanka, in 2008 and his MS degree in engineering technology from Middle Tennessee State University in 2012. His research interests mainly focus on fabrication and characterization of nanomaterials for thin film photovoltaics and opto-electronic devices.

Adam B. Phillips received his BS degree in physics from Case Western Reserve University in 1999 and his PhD in physics from the University of Virginia in 2007. He joined the University of Toledo in 2008 and is currently a research associate professor at UT’s Wright Center for Photovoltaic Innovation and Commercialization. His research includes investigating low cost materials and methods for carbon-free energy sources.

Michael J. Heben is a professor and the Wright Center Endowed chair for Photovoltaics in the Department of Physics and Astronomy at the University of Toledo (UT). He earned his MS degree in materials science and engineering from Stanford and his PhD in chemistry from Caltech under the guidance of N.S. Lewis. He became a postdoc with A.J. Nozik at SERI/NREL in 1990, and was a principal scientist and group leader when he left NREL for UT in 2008.

© 2016 Society of Photo-Optical Instrumentation Engineers (SPIE)
Zhaoning Song, Suneth C. Watthage, Adam B. Phillips, and Michael J. Heben "Pathways toward high-performance perovskite solar cells: review of recent advances in organo-metal halide perovskites for photovoltaic applications," Journal of Photonics for Energy 6(2), 022001 (15 April 2016). https://doi.org/10.1117/1.JPE.6.022001
Published: 15 April 2016
Lens.org Logo
CITATIONS
Cited by 231 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Perovskite

Solar cells

Crystals

Interfaces

Photovoltaics

Transparent conductors

Gold


CHORUS Article. This article was made freely available starting 15 April 2017

Back to Top