• Journal of Semiconductors
  • Vol. 41, Issue 5, 051202 (2020)
Zhizai Li and Zhiwen Jin
Author Affiliations
  • School of Physical Science and Technology & Key Laboratory for Magnetism and Magnetic Materials (MoE) & Key Laboratory of Special Function Materials and Structure Design (MoE) & National & Local Joint Engineering Laboratory for Optical Conversion Materials and Technology, Lanzhou University, Lanzhou 730000, China
  • show less
    DOI: 10.1088/1674-4926/41/5/051202 Cite this Article
    Zhizai Li, Zhiwen Jin. HI hydrolysis-derived intermediate as booster for CsPbI3 perovskite: from crystal structure, film fabrication to device performance[J]. Journal of Semiconductors, 2020, 41(5): 051202 Copy Citation Text show less

    Abstract

    Nowadays, inorganic CsPbI3 perovskite solar cells (PSCs) have become one of the most attractive research hotspots in photovoltaic field for its superior chemical stability and excellent photo-electronic properties. Since the first independent report in 2015, the power conversion efficiency (PCE) of CsPbI3 based PSCs has sharply increased from 3.9% to 19.03%. Importantly, during the developing process of CsPbI3 PSCs, HI hydrolysis-derived intermediate plays an important role: from stabilizing the crystal structure, optimizing the fabricated film to boosting the device performance. In this review, the different crystal and electronic structures of CsPbI3 are introduced. We then trace the history and disputes of HI hydrolysis-derived intermediate to make this review more logical. Meanwhile, we highlight the functions of HI hydrolysis-derived intermediate, and systematically summarize the advanced works on CsPbI3 PSCs. Finally, the bottlenecks and prospects are revealed to further increase the CsPbI3 PSCs performance.

    1. Introduction

    Since the first organic–inorganic hybrid perovskite solar cell (PSCs) was proposed by Miyasaka’s group[1], its power conversion efficiency (PCE) has sky-rocketed from 3.8% to 25.2% over the past decade[2-4]. This quick growth rate has made it as a hotspot in recent years because of its fascinating properties, such as high absorption coefficient[5, 6], low exciton binding energy[7-9], tunable band gap[10, 11], long carrier diffusion length[12, 13], and superb carrier mobility[14, 15]. Though great progresses have been achieved in term of its PCE, there still exist some problems. Volatility and hygroscopic A-site cations decompose the perovskite structure under extreme environment and destroy the performance of the device[16-21].

    CsPbX3 is a promising candidate to conquer these problems because Cs+ is the most feasible inorganic cation to replay volatility and hygroscopic A-site with suitable tolerance factor[22-26]. Besides, the structures of organic cation (e.g., CH3NH3+) have orientation freedom, while Cs+ is symmetric without a multiple structure, which make hybrid perovskites show an unstable structure under extreme environments[27-31]. Among all of the CsPbX3 materials, CsPbI3 with a bandgap of ~1.7 eV is a suitable and promising candidate for high performance and stable output photovoltaic material[32-34].

    CsPbI3 has four different phases (cubic (α), tetragonal (β), orthorhombic (γ) and non-perovskite yellow (δ) phase) and each phase transforms under different temperatures[35]. At room temperature (RT), CsPbI3 will finally transfer into non-perovskite phase (δ-phase) with an unsuitable bandgap of 2.75 eV, which limits its practical application[36-38]. Many researchers have conducted a series of methods to conquer this problem[39-41]. In the CsPbI3 PSCs development process, HI hydrolysis-derived intermediate plays an important role[42-44]: stabilizing the crystal structure, optimizing the fabricated film and improving the device performance.

    In this review, we aim to summarize the latest works about CsPbI3 PSCs based on HI hydrolysis-derived intermediate. First, we briefly review the different crystal and electronic structures of CsPbI3. We then trace the history and disputes of HI hydrolysis-derived intermediate to make this review more logical. Afterward, we highlight the functions of HI hydrolysis-derived intermediate, and systematically summarize some advanced works about HI hydrolysis-derived intermediate on CsPbI3 PSCs. Finally, present issues and outlines are discussed to further increase the CsPbI3 PSCs performance.

    2. Crystal/electronic structure

    Photo-electric properties (e.g., optical transitions, charger transfer) are greatly related to crystal and electronic properties (e.g., phase transition, energy band)[45-47]. In this section, we mainly discuss the CsPbI3 perovskite from two aspects: crystal structure and electronic structure.

    2.1. Crystal structure

    The CsPbI3 perovskite structure can be described as: Pb-site and I-site ion form a corner sharing [PbI6]4– octahedron, while the Cs cation resides in the cuboctahedral cavities[48, 49]. There are mainly four types of structures: cubic structure (α, Pm m), tetragonal structure (β, P4/mbm), orthorhombic structure (γ, Pbnm), and non-perovskite phase (δ, Pnma)[46], as shown in Fig. 1(a). CsPbI3 possesses an unsuitable Goldschmidt tolerance factor (t) in the range of 0.81–0.84, which is smaller than the ideal value (t, 0.9–1) and leads an unstable perovskite structure[50]. The transitions of each phase in different temperatures are shown in Fig. 1(b)[35].

    (Color online) (a) The structure and transition of CsPbI3 phases versus temperature. Reproduced with permission[46]. Copyright 2018, American Chemical Society Publications. (b) The transition of CsPbI3 thermal phase and their transition mechanism. Reproduced with permission[35]. Copyright 2019, Science Publishing Group. (c) Schematic of bonding/antibonding orbitals in CsPbX3. Reproduce with permission[54]. Copyright 2016, American Chemical Society Publications. (d) Electronic band structure of CsPbI3 calculated by DFT and (e) tight-binding model. Reproduced with permission[46]. Copyright 2018, American Chemical Society Publications.

    Figure 1.(Color online) (a) The structure and transition of CsPbI3 phases versus temperature. Reproduced with permission[46]. Copyright 2018, American Chemical Society Publications. (b) The transition of CsPbI3 thermal phase and their transition mechanism. Reproduced with permission[35]. Copyright 2019, Science Publishing Group. (c) Schematic of bonding/antibonding orbitals in CsPbX3. Reproduce with permission[54]. Copyright 2016, American Chemical Society Publications. (d) Electronic band structure of CsPbI3 calculated by DFT and (e) tight-binding model. Reproduced with permission[46]. Copyright 2018, American Chemical Society Publications.

    The detailed transition temperature between each phase was researched by Even et al., based on density functional theory (DFT) analyzation. The increasing thermal parameters of I tended to strength the dynamic motion of the corner-connected [PbI6/2] octahedral, which further induced a change of the unit cell volume and made δ-phase transformed to α-phase at 595 K. Then, with dynamic states gradually relaxing, α-phase transitions to β-phase at 539 K; β-phase transitions to γ-phase at 425 K; finally γ-phase turns to yellow non-perovskite phase (δ-phase) at RT[46].

    The different stability of each phases can be ascribed to the different dissociated energies. The dissociation energy from CsPbI3 to CsI and PbI2 for α-, γ-, and δ-phase are 0.04, –0.09 and –0.16 eV, respectively. δ-phase CsPbI3 shows a small Pb–I–Pb bond angel (95.09° and 91.40°) than α-phase CsPbI3 (180°) and γ-phase CsPbI3 (154.74°), which reduces the orbital overlap between Pb and I atoms and also makes δ-phase CsPbI3 with a deeper defect transition energy level than α-, γ-phase. This indicates that δ-phase is the most stable phase because of its lowest dissociation energy[51]. Most importantly, different cooling rates change the formation energies for CsPbI3 phases. α-phase converts to γ-phase when rapidly cooled in dry air, while slowly cooling leads to δ-phase phase because of different formation energies[52]. Besides, changing the temperature to cause the structure transition, a polar solvent also induces lattice distortion in CsPbI3. A polar solvent would induce the lattice distortion of CsPbI3 nanocubes by triggering the dipole moment, which leads to the self-assembly from α-CsPbI3 to a γ-phase through oriented attachment process[53]. These studies indicate that metastable phases (contained β, γ-phase CsPbI3) are more promising than α-CsPbI3.

    2.2. Electronic structure

    The valence band maximum (VBM) of CsPbI3 perovskite is constituted of antibonding hybridized Pb 6s and X np orbitals, among which X np takes the lead. However, Pb 6p is in the dominant place of conduction band minimum (CBM), as shown in Fig. 1(c)[54]. Compared with organic A-site cations, Cs+ has little effect on CsPbI3 electronic properties. A-site cations could indirectly influence perovskite electronic properties through Coulombic interactions and steric hindrance to deform the perovskite lattice, which makes perovskite electronic structure close to the band edges and further changes the band gap energetics[55].

    The calculated electronic of α-, β- and γ-CsPbI3 are depicted in Figs. 1(d) and 1(e)[46]. According to the tight-binding (TB) and DFT structures, taking band folding into account, we can vividly draw that all the different band gaps of CsPbI3 phases are direct, and the band gap of α-CsPbI3 shifts from the R point in the Brillouin zone to Z and Γ for β-CsPbI3 and γ-CsPbI3, respectively. This change indicates that the electronic band gap gradually increases with the transition from α-CsPbI3 to more distorted β- and γ-CsPbI3 because the [PbI6]4– rotations stabilize the top of VBM and destabilize the bottom of CBM[56].

    3. The functions of HI hydrolysis-derived intermediate

    We summarize the performance of CsPbI3 PSCs after introducing HI hydrolysis-derived intermediate in Table 1 (sPCE is the stable PCE). Its main functions can be summarized as following:

    Table Infomation Is Not Enable

    1) Reducing crystallization energy barrier in low temperature fabrication;

    2) Increasing iodide coordination numbers to decrease structural disorder, modifying structure and forming higher-order iodoplumbate complexes;

    3) Slowing down the rapid crystalline process and obtaining high-quality CsPbI3 film;

    4) Inducing strain to generate distorted metastable phase (β- and γ- CsPbI3);

    5) Modifying the band gap of perovskites films.

    4. History and disputes of HI hydrolysis-derived intermediate

    The solution one-step method has advantages of simple, convenience and facile process, and can also be compatible roll-to-roll fabrication technology[57]. However, the one-step deposition process shows poor morphology and low performance because of the rapid reaction in the solution. The common solution is using intermediates to slow down the quick reaction and make it controllable. However, understanding the composition of the HI hydrolysis-derived intermediate still needs a long time.

    4.1. HI

    In the early stage, researchers focused on using HI additive in CsPbI3 PSCs fabrication to cause a microstrain and induce a low temperature phase transition process. Meanwhile, extra halides in HI precursor solution tended to fill the vacancies of perovskites, resulting in change of metal–halogen–metal bond connectivity, and consequently cell volumes and optical bandgap[58]. Besides, PbI2 first coordinated with DMF in the precursor through Pb–O bonds, but further added HI would eliminate PbI2–DMF coordination and form higher-order iodoplumbate complexes (e.g., PbI42−, PbI53−, and PbI64−), which benefited the formation of high-quality CsPbI3 film[59, 60].

    In 2015, HI was first used as an additive in CsPbI3 PSCs. Snaith et al. introduced a small amount of HI in the precursor solution before spin-coating. They found that HI additive could change the solubility of precursor materials and induce a strain to lower the temperature phase transition. Then, strain triggered small crystals appearing and significantly stabilized its structure in RT, as shown in Fig. 2(a)[42]. Uddin et al. controlled the concentration of HI on purpose to modify CsPbI3 bandgap. They demonstrated that introducing 36 μL/mL HI would decrease the bandgap from 1.75 eV to optimized value of 1.7 eV and show excellence electronic properties with low charge-transport (12.8 kΩ). From the scanning electron microscopy (SEM) images, optimal concentration HI additive led to the appearance of small grain sizes with a few nanometers, which is beneficial to increase the stability of black phase CsPbI3[61]. This conclusion was also proven by Kim et al., who found that HI formed small grains and stabilized the black phase of CsPbI3 at low temperature[62]. Furthermore, they investigated the function of NH4+ (e.g., NH4Cl, NH4Br and NH4I) and H+ (HCl, HBr and HI) based additive and found that HI additive is the most efficient, and which could reduce the roughness and increase the stability of perovskites films. In particular, when it was exposed in ambient for 3 h, the optimal device with CsPbI3 perovskite only dropped its PCE from 3.55% to 2.78%[63].

    (Color online) (a) The diagrammatic of HI fabricated CsPbI3. Reproduced with permission[42]. Copyright 2015, The Royal Society of Chemistry. (b) Schematic of using HPbI3 to fabricate FAPbI3 PSCs. Reproduced with permission[65]. Copyright 2015, Wiley-VCH Publications. (c) Detail information of PbI2 and HPbI3 fabricated perovskite film. Reproduced with permission[69]. Copyright 2018, Wiley-VCH Publications. (d) The molecular structure of FA and DMA, and the tolerance factor of corresponding perovskite (CsPbI3, Cs0.7DMA0.3PbI3 and DMAPbI3). Reproduced with permission[71]. Copyright 2018, Nature Publishing Group.

    Figure 2.(Color online) (a) The diagrammatic of HI fabricated CsPbI3. Reproduced with permission[42]. Copyright 2015, The Royal Society of Chemistry. (b) Schematic of using HPbI3 to fabricate FAPbI3 PSCs. Reproduced with permission[65]. Copyright 2015, Wiley-VCH Publications. (c) Detail information of PbI2 and HPbI3 fabricated perovskite film. Reproduced with permission[69]. Copyright 2018, Wiley-VCH Publications. (d) The molecular structure of FA and DMA, and the tolerance factor of corresponding perovskite (CsPbI3, Cs0.7DMA0.3PbI3 and DMAPbI3). Reproduced with permission[71]. Copyright 2018, Nature Publishing Group.

    4.2. PbI2·xHI or HPbI3

    In addition to HI additive, HI hydrolysis-derived intermediate is more effective because it eliminates water in the HI solution and is an intermediate compound to increase perovskite crystallinity[64]. It was first proposed in 2015 by Zhao et al. They developed a new precursor compound (named HPbI3) through a reaction of HI and PbI2 in DMF solution, and used it to replace PbI2 in fabrication FAPbI3-based PSCs, as shown in Fig. 2(b)[65]. Such HPbI3 has a pseudo-3D crystal structure, where 1D face-shared [PbI6]4– octahedra with intercalated protons (H+) for charge balance[66].

    Zhu et al. introduced HPbI3 into the CsPbI3 PSCs and assisted with a triple cation NH3+C2H4NH2+C2H4NH3+ (named as DETA3+) to further stabilize the α-CsPbI3 perovskite phases[67]. Subsequently, Chen et al. used HPbI3 to substitute PbI2 and found that tensile lattice strain appeared in HPbI3-process CsPbI3 perovskite[68]. The tensile lattice strain generated because HPbI3 crystals serve as a template to guide the nucleation and growth of α-CsPbI3. A similar work was conducted by Zhao et al., who introduced the I-excess precursor HPbI3+x and replace PbI2 to reduce crystallization energy barrier, and fabricated α-CsPbI3–xBrx PSCs in low temperature (130 °C). The champion PCE of 13.61% was measured by a reverse scanning, as shown in Fig. 2(c)[69]. More recently, in order to overcome the low PCE problem of CsPbI3-based PSC, You et al. used CsI, HPbI3+x and PbBr2 as precursor to fabricate high-quality CsPbI3–xBrx perovskite films. They developed an inorganic shunt-blocking layer lithium fluoride (LiF) in the ETL/perovskite interface to align the bandgap and suppress the surface defect. Furthermore, a small amount of PbCl2 were introduced to further suppress the recombination. Finally, they obtained CsPbI3–xBrx with the highest PCE of 18.64%, and boosted VOC to 1.25 V with little loss. After continuous 1 sun equivalent illumination, the best device dropped only 6% of its initial PCE after 1000 h[70].

    4.3. PbI2·xDMAI or DMAPbI3

    Kanatzidis et al. recently claimed that HPbI3 did not exist and was replaced by a compound of DMAPbI3, which generated through DMF hydrolysis in HI solution. Importantly, they pointed out that some early reports of inorganic perovskite are actually the hybrid perovskite. They found that DMAPbI3 possessed a larger tolerance factor and mixing with Cs+ could adjust tolerance factor (t) of the compounds (Cs1−xDMAxPbI3) toward an ideal factor (t, 0.9–1). Finally, they achieved a champion PCE of 12.62% in Cs1−xDMAxPbI3-based PSCs, as shown in Fig. 2(d)[71]. Liu et al. dissolved PbI2 and HI in DMF to synthesize DMAPbI3 and further confirmed that no HPbI3 existed. They used it as precursor to fabricate high-quality CsxDMA1−xPbI3 perovskite films with 14.3% PCE, and the initial PCE kept more than 85% when exposed in air 20 days without encapsulation, as shown in Fig. 3(a)[72].

    (Color online) (a) Schematic illustration the fabrication process of CsxDMA1–xPbI3. Reproduced with permission[72]. Copyright 2019, Elsevier Inc Publications. (b) Schematic diagram of using DMAI additive to form CsPbI3 films. Reproduced with permission[74]. Copyright 2019, Wiley-VCH Publications. (c) The changeable component of DMAI-fabricated perovskite versus annealing temperature. Reproduced with permission[75]. Copyright 2020, American Chemical Society Publications. (d) Schematic diagram of DMAPbI3 synthesis process and the information of corresponding perovskite. Reproduced with permission[76]. Copyright 2019, Wiley-VCH Publications.

    Figure 3.(Color online) (a) Schematic illustration the fabrication process of CsxDMA1–xPbI3. Reproduced with permission[72]. Copyright 2019, Elsevier Inc Publications. (b) Schematic diagram of using DMAI additive to form CsPbI3 films. Reproduced with permission[74]. Copyright 2019, Wiley-VCH Publications. (c) The changeable component of DMAI-fabricated perovskite versus annealing temperature. Reproduced with permission[75]. Copyright 2020, American Chemical Society Publications. (d) Schematic diagram of DMAPbI3 synthesis process and the information of corresponding perovskite. Reproduced with permission[76]. Copyright 2019, Wiley-VCH Publications.

    Zhao et al. used PbI2·xDMAI to fabricate CsPbI3 PSCs recently, and they concluded that the fabricated perovskites are actually all inorganic composition because the organic ion DMA+ are easily lost during the high-temperature (210 °C) annealing process[73]. Later, they proved that DMAI is a volatile additive, and used it to assist with phenyltrimethylammonium chloride (PTACl) passivation. Finally, they obtained the highest PCE of CsPbI3 PSC, 19.03%, as shown in Fig. 3(b)[74]. This conclusion was also confirmed by Pang et al., who found that DMAPbI3 and Cs4PbI6 first formed in annealing 100 °C for 10 min. When the annealing temperature was increased to 180 °C for 15 min, DMAPbI3 converted into DMA0.15Cs0.85PbI3 with a small amount Cs4PbI6 residue. Once the annealing temperature exceeded 200 °C, the γ-phase CsPbI3 was formed with a small number of DMA0.15Cs0.85PbI3 residue. The number of DMA0.15Cs0.85PbI3 greatly decreased and a little δ-phase appeared, as shown in Fig. 3(c)[75].

    Our groups also confirmed this conclusion. We synthesized a series of intermediate compounds (DMAI and DMAPbI3) by different ratio of HI/DMF, and used them to fabricate CsPbI3 PSCs. After detailed analysis, we found that the major component of CsPbI3 was still inorganic in this reaction route. Most of DMA+ organic molecules lost during the annealing process, and only a small amount of DMA+ remained to stabilize perovskite structure. Excessive DMA+ interacted with Pb2+ to further passivate CsPbI3 surface, as shown in Fig. 3(d)[76].

    In conclusion, the organic molecule DMAI mainly influence the crystallization kinetics and perovskite phase. During the annealing process, DMAI will sublimate quickly, change the rate of crystallization and form metastable (β- and γ-) phase based CsPbI3. The controllable crystallization kinetics and stable (β- and γ-) phase are beneficial to morphology and stability of perovskite, respectively. Besides, DMA+ (2.72 Å) possesses a larger ionic radius than Cs+ (1.88 Å)[77]. Therefore, if the DMA+ is residual (non-sublimate) in the CsPbI3 film: (1) DMA+ doped into the lattice of CsPbI3 can increase its tolerance factor for avoiding crystal structure distortion; (2) DMA+ reacted with Pb2+ to passivate CsPbI3 surface, further reducing leakage current generation for increasing the device performance.

    5. Applying HI hydrolysis-derived intermediate in CsPbI3 PSCs

    5.1. α-phase CsPbI3 based PSCs

    As we discussed earlier, HI hydrolysis-derived intermediate showed a lot of advantages in high-quality film fabrication and device performance. Importantly, perovskite films with better crystallinity, morphology, and higher range of absorption are the foundation of efficiency.

    The first working α-CsPbI3 PSCs with a PCE of 2.9% was fabricated in low temperature (100 °C) by Snaith and his cooperators via a small amount HI additive adding[42]. Many relevant works have been done to boost its performance. Uddin et al.[61] and Kim et al.[62] used HI to modify the α-CsPbI3 film morphology and boost its PCE to 6.44% and 4.88%, respectively. Later, Zhao et al. discovered that PbI2.xHI could reduce the crystallization energy barrier for α-CsPbI3 phase. They used PbI2.xHI assist with two-dimension EDAPbI4 perovskite to stabilize α-CsPbI3 phase and avoid lattice distortion. Finally, α-CsPbI3 phase based on the EDAPbI4 passivation showed a record PCE of 11.8% together with superior stability, as shown in Fig. 4(a). The α-CsPbI3 phase perovskite kept its structure after annealing at 100 °C for more than 150 h and stable at RT for months[78].

    (Color online) (a) The detail information of PbI2.HI and PbI2 fabricated perovskite, inserted pictures are their digital photos. Reproduced with permission[78]. Copyright 2017, Wiley-VCH Publications. (b) The diagram of PbI2 and HPbI3 fabricated CsPbI3 film, respectively. Reproduced with permission[68]. Copyright 2018, American Chemical Society Publications. (c) Schematic of PEA+ organic ligand treatment on CsPbI3 thin film. Reproduced with permission[79]. Copyright 2018, Elsevier Inc Publications. (d) Diagram illustrates the mechanism of with/without OTG passivation. Reproduced with permission[80]. Copyright 2019, Wiley-VCH Publications.

    Figure 4.(Color online) (a) The detail information of PbI2.HI and PbI2 fabricated perovskite, inserted pictures are their digital photos. Reproduced with permission[78]. Copyright 2017, Wiley-VCH Publications. (b) The diagram of PbI2 and HPbI3 fabricated CsPbI3 film, respectively. Reproduced with permission[68]. Copyright 2018, American Chemical Society Publications. (c) Schematic of PEA+ organic ligand treatment on CsPbI3 thin film. Reproduced with permission[79]. Copyright 2018, Elsevier Inc Publications. (d) Diagram illustrates the mechanism of with/without OTG passivation. Reproduced with permission[80]. Copyright 2019, Wiley-VCH Publications.

    Compared with HI, the absence of H2O molecules in HI hydrolysis-derived intermediate can optimize the perovskite crystallinity and morphology. Chen and his cooperators replaced PbI2 with HPbI3 in fabricating stable α-CsPbI3 film. They found that the bandgap was shifted from 1.72 to 1.68 eV owing to formation of tensile lattice strain. Finally, a HTL free α-CsPbI3 was obtained with a higher PCE of 9.5%. Besides, the optimal device showed enhanced stability, which maintained 90% of its initial PCE under illumination for more than 3000 h in dry environment, as shown in Fig. 4(b)[68]. Then, organic terminal groups were widely used to assist HPbI3 and further improve α-phase CsPbI3 stability. Zhu et al. introduced DETA3+ additive into the HPbI3 containing precursor to stabilize α-CsPbI3 perovskite phases. DETA3+ has NH3+ or RHN2+ group, which could combine with I or [PbI6]4– and avoid octahedral tilting. Besides, oil-wet (hydrophobic) hydrocarbon chains of DETA3+ increased the CsPbI3 humidity-resistance. Ultimately, the optimal device kept its structure for more than 6 h in a humid environment (~30 °C, 60%–70% RH), while the reference device changed its color from black to yellow in the same condition in 55 min, indicating a phase transition[67]. Zhao et al. fabricated HPbI3+x to form α-CsPbI3 perovskite film. With the help of PEAI post-treated on the α-CsPbI3, superior PCE of 13.5% was obtained with improved stability. Because of PEA+ terminated on the CsPbI3 surface and did as a capping layer, PEA+-CsPbI3 remained stable structure after 80 °C annealing for 7 days, while the pure one degraded into non-perovskite in the same condition, as shown in Fig. 4(c)[79] . Han et al. added organic terminal groups (OTG) into CsI and HPbI3 precursors to design an inverted planar CsPbI3 PSCs. They report that OTG induced a steric hindrance and suppressed octahedral [PbI6]4– tilting. Moreover, OTG passivated the surface electronic traps states to further increase its performances. Finally, the inverted planar OTG-CsPbI3 PSCs showed the highest PCE of 13.2% and retained about 85% of its initial PCE for 30 days at RT, while the reference device degraded completely in 3 weeks, as shown in Fig. 4(d)[80].

    5.2. Metastable (β- and γ-) phase CsPbI3 based PSCs

    Recently, the CsPbI3 films fabricated by HI hydrolysis-derived intermediate were proved metastable phases (combined β-phase CsPbI3 with γ-phase CsPbI3).

    The β-phase CsPbI3 can also be formed at low temperature and show more stable perovskite structure than α-phase one. However, it is difficult to deposit and stabilize its perovskite structure[81]. Zhao et al. adopted PbI2·DMAI and CsI as precursor to fabricate stable β-phase CsPbI3 with stable structure. Furthermore, they used choline iodine (CHI) to passiviate the surface trap states and aligned the energy level in the TiO2/β-phase interface. β-phase CsPbI3 PSCs has a particularly high PCE of 18.4% with distinguished ambient stability because of PbI2·DMAI and CHI. The stability of β-CsPbI3 PSCs were greatly improved: it retained 92% of its initial PCE after 500 h illumination at the maximum power point, as shown in Fig. 5(a)[73].

    (Color online) (a) Schematic illustration of CHI crack-filling interface engineering. Reproduced with permission[73]. Copyright 2019, Science Publishing Group. (b) Schematic diagram CsPbI3 crystal formation by using HI and H2O. Reproduced with permission[83]. Copyright 2018, American Chemical Society Publications. (c) Mechanism of STCG-CsPbI3 film formation by assistant of ADMA molecule. Reproduced with permission[84]. Copyright 2020, Wiley-VCH Publications. (d) The schematic illustration of HI and PEAI do on the CsPbI3. Reproduced with permission[43]. Copyright 2018, Nature Publishing Group.

    Figure 5.(Color online) (a) Schematic illustration of CHI crack-filling interface engineering. Reproduced with permission[73]. Copyright 2019, Science Publishing Group. (b) Schematic diagram CsPbI3 crystal formation by using HI and H2O. Reproduced with permission[83]. Copyright 2018, American Chemical Society Publications. (c) Mechanism of STCG-CsPbI3 film formation by assistant of ADMA molecule. Reproduced with permission[84]. Copyright 2020, Wiley-VCH Publications. (d) The schematic illustration of HI and PEAI do on the CsPbI3. Reproduced with permission[43]. Copyright 2018, Nature Publishing Group.

    The γ-phase CsPbI3 is the most stable black phase because of its lowest dissociation energy[82]. Hence, γ-phase CsPbI3 was systematically researched to boost its stability and performance. Hu et al. used HI to fabricate high-quality CsPbI3 film. After that, they introduced a small amount of H2O into precursor to induce a proton transfer process in CsPbI3 film, which could manipulate the grains size and improve stability of perovskite. When stored in ambient conditions, the optimal device showed no drop of its performance while the reference’s PCE was greatly degraded, as shown in Fig. 5(b)[83]. Then, Nazeeruddin et al. developed a soft template-controlled growth method to fabricate pinhole-free γ-phase CsPbI3 film, where (adamantan-1-yl)methanammonium (ADMA) acted as a template and ionized by HI. They pointed that ADMA absorbed on CsPbI3 surface and induced a steric effect to further increase the nucleation rate at the initial stage of CsPbI3 formation. Taking into account of controllable nucleation rate and excellent morphology, superior PCE of 16.04% was obtained with improved stability (drop only 10% after continuous light soaking and heating for 1000 h), as shown in Fig. 5(c)[84].

    In our recent research, we reported the synergistic effect of HI and PEAI additives, where HI transferred to an intermediate (HPbI3+x) to fabricate distorted black phase-based CsPbI3 thin films and PEAI induced a steric effects to avoid phase transition. It is noteworthy that the best device maintained 92% of its initial PCE for 60 days storage in ambient (RH ~ 20%–30%, 25 °C), while the reference one degraded to 0.65% in the same condition for 8 days, as shown in Fig. 5(d)[43]. Then, HPbI3 was used to fabricate CsPbI3, assisted with anti-solvent hot substrate spin-coating method, and 15.91% PCE was obtained in a humidity environment (RH ~ 50%)[85]. Furthermore, we added a small amount of Br- (5%) to the HPbI3-CsPbI3 lattice to increase phase stability by suppressing bulk trap-assisted non-radiative recombination and relaxing lattice strain. The performance was boost up to a record PCE of 17.17% together with excellent stability[86]. Another work we reported was that 3 mol% Cl was added to the γ-CsPbI3 film with a 16.07% PCE obtained. It not only showed the excepted increase crystalline dynamics for an excellent CsPbI3 morphology, but also improved crystalline orientation. The Cl-doping CsPbI3 showed high stability, for optimal device dropped only 0.45% of its initial PCE under continuous light soaking for 200 h. More important, non-encapsulated CsPbI3 PSCs with Cl doping degraded only 6% when stored in RH ~ 30% for 60 days (the fresh one dropped to 85% of its initial PCE in the same conditions)[87]. We also introduced 2% Pb(SCN)2 into the DMAPbI3 and CsI precursor to control the morphology of the CsPbI3 film. In this case, a PCE of 17.04% with VOC of 1.09 V was obtained[88].

    One of the notorious problems to limit CsPbI3 performance is the lower JSC compared with hybrid one. Thus, we have also developed several strategies to increase its JSC, such as harvesting short wavelength ultraviolet light (UV-light) or near-infrared (NIR) light, and designing device structure to capture light. First, we developed a downconversion nanoparticles (DCNPs) nitrogen-doped graphene quantum dots (N-GQDs) as an energy-down-shift to harvest the short wavelength (< 350 nm) UV-light. After combining it with HPbI3-formed γ-CsPbI3, the optimal device showed an improved short circuit current density (JSC) from 18.67 to 19.15 mA/cm2, with an increase of 2.57%. Meanwhile, its performance was greatly increased 3.15%, from 15.53% to 16.02%[89]. Furthermore, we developed a core-shell-structured upconversion nanoparticles (UCNPs) to capture the NIR light, making it possible to obtain 15.86% PCE (noted that the improvement of PCE is negligible), JSC = 19.17 mA/cm2[90]. We also investigated the influence of different haze glass substrates, and used the optimized one to fabricate CsPbI3 PSCs with 16.24% PCE. We found that the improvement came from scattering effect of FTO, refractive index and roughness of each layer[91].

    5.3. Low dimension CsPbI3 based PSCs

    Reducing dimension can further increase the stability of CsPbI3 PSCs because reducing materials dimension can lead to more symmetric crystal structure and show a smaller surface energy[92-96].

    However, the poor solubility of CsX in the precursor solution would severely limited the thickness of CsPbI3 film and influence the light absorption. Chen et al. used HPbX3 and CsAc as new precursor to overcome the poor solubility of Cs+ precursor and fabricate α-CsPbX3 with optimal thickness. They introduced phenylethylammonium iodide (PEAI) to HPbX3 and CsAc system and further controlled the dimension of CsPbX3 from three dimension (3D) to two dimension (2D). Finally, a champion PCE of 12.4% in 2D CsPbI3 perovskite was obtained, and maintained 93% of its initial PCE in ambient for 40 days, as shown in Fig. 6(a)[97]. Similarly, our group used DMAPbI3 as a new precursor to fabricated γ-CsPbI3 perovskite, and then we introduced a judicious amount of PEAI into the DMAPbI3 contained precursor to convert γ-CsPbI3 PSCs into a 2D Ruddlesden–Popper (RP) structure (PEA)2(Cs)n−1PbnI3n+1 perovskite. And the optimal 2D RP PSCs with highest PCE of 13.65%, showed a similar charge extraction and carrier lifetime compared with the 3D samples. The thermostability was tested in N2 filled glovebox. The superior 2D RP PSCs kept 88% of its initial PCE when stored at 80 °C for 15 days, while the reference 3D γ-CsPbI3 degraded to 69%. Besides, 2D RP PSCs maintained its black color whereas 3D γ-CsPbI3 appeared yellow phase in RH ~ 30% for 12 days, as shown in Fig. 6(b)[98]. In addition to PEAI, Chen et al. first introduced a novel dual ammonium cation piperazine-1,4-diium (PZD+) to generate 2D RP CsPbI3. It’s noted that one of the ammonium groups coordinated with [PbI6]4– octahedral, and another one interacted with I to balance charge. Finally, the optimized device achieved a PCE of 9.39% and maintained its performance without decomposition under heating at 100 °C for 24 h[99].

    (Color online) (a) The structure and decomposition energies of different n values PEA2Csn-1PbnX3n+1. Reproduced with permission[97]. Copyright 2018, Elsevier Inc Publications. (b) The controllable n values and structures of PEA2Csn-1PbnX3n+1. Reproduced with permission[98]. Copyright 2019, Wiley-VCH Publications. (c) Schematic illustration the fabrication process of shell ligand, HPbI3, H2PbI4 and in-suit assembled of them. Reproduced with permission[101]. Copyright 2019, Wiley-VCH Publications.

    Figure 6.(Color online) (a) The structure and decomposition energies of different n values PEA2Csn-1PbnX3n+1. Reproduced with permission[97]. Copyright 2018, Elsevier Inc Publications. (b) The controllable n values and structures of PEA2Csn-1PbnX3n+1. Reproduced with permission[98]. Copyright 2019, Wiley-VCH Publications. (c) Schematic illustration the fabrication process of shell ligand, HPbI3, H2PbI4 and in-suit assembled of them. Reproduced with permission[101]. Copyright 2019, Wiley-VCH Publications.

    Pradhan et al. fabricated stable CsPbI3 nanocrystals (NCs) with superior stability by using a higher temperature (260 °C) than usual (160 °C), and adding olelyamine (OLA) and HI respectively in the reaction process (noted that only OLA or HI are less efficient). Taking the NMR analyzation into account, they found that higher temperature helped the OLA+ ligands to occupy the Cs+ position on the surface and further stabilized its structure[100]. To solve adverse shell ligands and unstable core lattices in nanocrystals, Choi et al. developed a rational core-shell design method in CsPbI3 NCs. They used a novel ligand named 4-fluorophenethylammonium iodide (FPEAI) to enhance the binding force between the ligand and CsPbI3 NCs and efficient charge coupling between NCs, to increase charge extraction. Besides, H2PbI4, which was synthesized with excessive HI, was used to assist doping Mn2+ ion into perovskite lattices and to further raise the performance of CsPbI3 NCs to 13.4% with superb stability (maintained 92% of its PCE in air for 500 h), as shown in Fig. 6(c)[101].

    6. Prospects and outlook

    Although lots of advanced works about HI hydrolysis-derived intermediate have been done to boost PCE of CsPbI3 PSCs, its PCE still far behind the hybrid ones. Therefore, we need to analyze the urgent problems that remain and develop corresponding strategies to improve the performance of CsPbI3 perovskites.

    Increasing the light absorption.As we show in Table 1, though the performance of CsPbI3 PSCs has reach 19.03%[74] , the JSC (about 20 mA/cm2) is still lower than the hybrid one. This maybe cause by current loss and insufficient utilization of light, resulting from the relatively large bandgap (compared with hybrid perovskite films). Developing high efficiency DCNPs/UCNPs, to capture UV-light or NIR could be an useful solution[90].

    Doping X-site halide to improve CsPbI3 PSCs stability.The unstable nature of CsPbI3 is an unsuitable tolerance factor (t), as we mentioned earlier. Partially replacing I (2.2 Å) with smaller radius halides, e.g., Cl (1.87 Å) and Br (1.96 Å) or pseudohalide SCN (2.15 Å), could enlarge t value to reach ideal one (0.9–1)[32].

    Ligand assists to stabilize CsPbI3 crystal structure. Introducing ligand (e.g., OTG[80], poly-vinylpyrrolidone (PVP)[102], ADMA[84], PEAI[79], and DETA[67]) into a perovskite precursor and inducing a steric hindrance to further stable CsPbI3 crystal structure is a useful way to boost its stability. Therefore, it is promising to develop and design new ligands.

    Increasing the carrier transport for low dimension perovskite. Low dimension CsPbI3 tends to increase surface-area-to-volume ratio and raise the Gibbs free energy to make the structure more stable[103]. It is noted that though low dimension CsPbI3 can keep CsPbI3 structure more stable, the PCE of low dimension CsPbI3 is still far below the 3D CsPbI3. Hence, we should pay more attention to increase the carrier transport for low dimension CsPbI3.

    Avoiding the disadvantage of intermediate.Liu et al. claimed that intermediates had negative influence during low dimension perovskite deposition process because they slowed down intercalation of ions and increased nucleation barrier, and further caused the byproduct formation[104]. Although there are no negative effects on HI hydrolysis-derived intermediate so far, we should be more vigilant against HI hydrolysis-derived intermediate.

    In conclusion, CsPbI3 perovskite, particularly the metastable phases (β- and γ-Phase CsPbI3), is a promising material to replace the unstable hybrid perovskite. Besides, CsPbI3 PSCs with suitable bandgap make it more suitable to apply in tandem solar cells and commercialization. Based on these advantages, we conclude that CsPbI3 PSCs maybe the mainstream research direction in the near future, and we should adopt a positive attitude to it.

    Acknowledgements

    This work was funded by the National Natural Science Foundation of China (51902148, 61704099 and 51801088), the Fundamental Research Funds for the Central Universities (lzujbky-2020-61, lzujbky-2019-88 and lzujbky-2020-kb06), and the Special Funding for Open and Shared Large-Scale Instruments and Equipments of Lanzhou University (LZU-GXJJ-2019C023 and LZU-GXJJ-2019C019).

    References

    [1] A Kojima, K Teshima, Y Shirai et al. Organometal halide perovskites as visible-light sensitizers for photovoltaic cells. J Am Chem Soc, 131, 6050(2009).

    [2]

    [3] J Jiang, Q Wang, Z Jin et al. Polymer doping for high-efficiency perovskite solar cells with improved moisture stability. Adv Energy Mater, 8, 1701757(2018).

    [4] J Jiang, Z Jin, F Gao et al. CsPbCl3-driven low-trap-density perovskite grain growth for > 20% solar cell efficiency. Adv Sci, 5, 1800474(2018).

    [5] C Wehrenfennig, G E Eperon, M B Johnston et al. High charge carrier mobilities and lifetimes in organolead trihalide perovskites. Adv Mater, 26, 1584(2014).

    [6] W Hu, H Cong, W Huang et al. Germanium/perovskite heterostructure for high-performance and broadband photodetector from visible to infrared telecommunication band. Light: Sci Appl, 8, 106(2019).

    [7] V D'Innocenzo, G Grancini, M J P Alcocer et al. Excitons versus free charges in organo-lead tri-halide perovskites. Nat Commun, 5, 3586(2014).

    [8] Q Lin, A Armin, R C R Nagiri et al. Electro-optics of perovskite solar cells. Nat Photon, 9, 106(2014).

    [9] H H Fang, F Wang, S Adjokatse et al. Photoexcitation dynamics in solution-processed formamidinium lead iodide perovskite thin films for solar cell applications. Light: Sci Appl, 5, e16056(2016).

    [10] J H Noh, S H Im, J H Heo et al. Chemical management for colorful, efficient, and stable inorganic–organic hybrid nanostructured solar cells. Nano Lett, 13, 1764(2013).

    [11] H Bian, D Bai, Z Jin et al. Graded bandgap CsPbI2+xBr1–x perovskite solar cells with a stabilized efficiency of 14.4%. Joule, 2, 1500(2018).

    [12] S D Stranks, G E Eperon, G Grancini et al. Electron-hole diffusion lengths exceeding 1 micrometer in an organometal trihalide perovskite absorber. Science, 342, 341(2013).

    [13] H Wang, H Bian, Z Jin et al. Synergy of hydrophobic surface capping and lattice contraction for stable and high-efficiency inorganic CsPbI2Br perovskite solar cells. Solar RRL, 2, 1800216(2018).

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

    [15] Y C Zhao, W K Zhou, X Zhou et al. Quantification of light-enhanced ionic transport in lead iodide perovskite thin films and its solar cell applications. Light: Sci Appl, 6, e16243(2017).

    [16] C Xiao, Z Li, H Guthrey et al. Mechanisms of electron-beam-induced damage in perovskite thin films revealed by cathodoluminescence spectroscopy. J Phys Chem C, 119, 26904(2015).

    [17] A F Akbulatov, S Y Luchkin, L A Frolova et al. Probing the intrinsic thermal and photochemical stability of hybrid and inorganic lead halide perovskites. J Phys Chem Lett, 8, 1211(2017).

    [18] W Zhou, Y Zhao, X Zhou et al. Light-independent ionic transport in inorganic perovskite and ultrastable cs-based perovskite solar cells. J Phys Chem Lett, 8, 4122(2017).

    [19] Q Wang, X Zhang, Z Jin et al. Energy-down-shift CsPbCl3:Mn quantum dots for boosting the efficiency and stability of perovskite solar cells. ACS Energy Lett, 2, 1479(2017).

    [20] Z Jin, J Yan, X Huang et al. Solution-processed transparent coordination polymer electrode for photovoltaic solar cells. Nano Energy, 40, 376(2017).

    [21] J Jiang, Z Jin, J Lei et al. ITIC surface modification to achieve synergistic electron transport layer enhancement for planar-type perovskite solar cells with efficiency exceeding 20%. J Mater Chem A, 5, 9514(2017).

    [22] R E Beal, D J Slotcavage, T Leijtens et al. Cesium lead halide perovskites with improved stability for tandem solar cells. J Phys Chem Lett, 7, 746(2016).

    [23] X Jia, C Zuo, S Tao et al. CsPb(IxBr1−x)3 solar cells. Sci Bull, 64, 1532(2019).

    [24] X Zhang, Z Jin, J Zhang et al. All-ambient processed binary CsPbBr3-CsPb2Br5 perovskites with synergistic enhancement for high-efficiency Cs-Pb-Br-based solar cells. ACS Appl Mater Interfaces, 10, 7145(2018).

    [25] J Zhang, D Bai, Z Jin et al. 3D–2D–0D interface profiling for record efficiency all-inorganic CsPbBrI2 perovskite solar cells with superior stability. Adv Energy Mater, 8, 1703246(2018).

    [26] D Bai, J Zhang, Z Jin et al. Interstitial Mn2+-driven high-aspect-ratio grain growth for low-trap-density microcrystalline films for record efficiency CsPbI2Br solar cells. ACS Energy Lett, 3, 970(2018).

    [27] Y Y Zhang, S Chen, P Xu et al. Intrinsic instability of the hybrid halide perovskite semiconductor CH3NH3PbI3. Chin Phys Lett, 35, 036104(2018).

    [28] C H Kang, I Dursun, G Liu et al. High-speed colour-converting photodetector with all-inorganic CsPbBr3 perovskite nanocrystals for ultraviolet light communication. Light: Sci Appl, 8, 94(2019).

    [29] G Liu, C Zhou, F Wan et al. Dependence of power conversion properties of perovskite solar cells on operating temperature. Appl Phys Lett, 113, 3501(2018).

    [30] G Liu, B Yang, B Liu et al. Irreversible light-soaking effect of perovskite solar cells caused by light-induced oxygen vacancies in titanium oxide. Appl Phys Lett, 111, 3501(2017).

    [31] J F Wang, D X Lin, Y B Yuan. Recent progress of ion migration in organometal halide perovskite. Acta Phys Sin, 68, 158801(2019).

    [32] W Ahmad, J Khan, G Niu et al. Inorganic CsPbI3 perovskite-based solar cells: a choice for a tandem device. Solar RRL, 1, 1700048(2017).

    [33] P Wang, X Zhang, Y Zhou et al. Solvent-controlled growth of inorganic perovskite films in dry environment for efficient and stable solar cells. Nat Commun, 9, 2225(2018).

    [34] X Zhang, Q Wang, Z Jin et al. Stable ultra-fast broad-bandwidth photodetectors based on α-CsPbI3 perovskite and NaYF4:Yb,Er quantum dots. Nanoscale, 9, 6278(2017).

    [35] J A Steele, H D Jin, I Iurii et al. Thermal unequilibrium of strained black CsPbI3 thin films. Science, 365, 679(2019).

    [36] J B Hoffman, A L Schleper, P V Kamat. Transformation of sintered CsPbBr3 nanocrystals to cubic CsPbI3 and gradient CsPbBrxI3–x through halide exchange. J Am Chem Soc, 138, 8603(2016).

    [37] Q Wang, Z Jin, D Chen et al. μ-graphene crosslinked CsPbI3 quantum dots for high efficiency solar cells with much improved stability. Adv Energy Mater, 8, 1800007(2018).

    [38] H Zhao, J Xu, S Zhou et al. Preparation of tortuous 3D γ-CsPbI3 films at low temperature by CaI2 as dopant for highly efficient perovskite solar cells. Adv Funct Mater, 29, 1808986(2019).

    [39] A S Dayan, B E Cohen, S Aharon et al. Enhancing stability and photostability of CsPbI3 by reducing its dimensionality. Chem Mater, 30, 8017(2018).

    [40] T Ye, B Zhou, F Zhan et al. Below 200 °C fabrication strategy of black phase CsPbI3 film for ambient-air-stable solar cells. Solar RRL, 10(2019).

    [41] S Xiang, W Li, Y Wei et al. Natrium doping pushes the efficiency of carbon-based CsPbI3 perovskite solar cells to 10.7%. iScience, 15, 156(2019).

    [42] G E Eperon, G M Paternò, R J Sutton et al. Inorganic caesium lead iodide perovskite solar cells. J Mater Chem A, 3, 19688(2015).

    [43] K Wang, Z Jin, L Liang et al. All-inorganic cesium lead iodide perovskite solar cells with stabilized efficiency beyond 15%. Nat Commun, 9, 4544(2018).

    [44] Y Wang, T Zhang, M Kan et al. Bifunctional stabilization of all-inorganic α-CsPbI3 perovskite for 17% efficiency photovoltaics. J Am Chem Soc, 140, 12345(2018).

    [45] A Swarnkar, V K Ravi, A Nag. Beyond colloidal cesium lead halide perovskite nanocrystals: analogous metal halides and doping. ACS Energy Lett, 2, 1089(2017).

    [46] A Marronnier, G Roma, S Boyer-Richard et al. Anharmonicity and disorder in the black phases of cesium lead iodide used for stable inorganic perovskite solar cells. ACS Nano, 12, 3477(2018).

    [47] D Bai, H Bian, Z Jin et al. Temperature-assisted crystallization for inorganic CsPbI2Br perovskite solar cells to attain high stabilized efficiency 14.81%. Nano Energy, 52, 408(2018).

    [48] M A Green, A Ho-Baillie, H J Snaith. The emergence of perovskite solar cells. Nat Photon, 8, 506(2014).

    [49] G E Eperon, S D Stranks, C Menelaou et al. Formamidinium lead trihalide: a broadly tunable perovskite for efficient planar heterojunction solar cells. Energy Environ Sci, 7, 982(2014).

    [50] J Zhang, G Hodes, Z Jin et al. All-inorganic CsPbX3 perovskite solar cells: progress and prospects. Angew Chem Int Ed, 58, 15596(2019).

    [51] Y Huang, W J Yin, Y He. Intrinsic point defects in inorganic cesium lead iodide perovskite CsPbI3. J Phys Chem C, 122, 1345(2018).

    [52] R J Sutton, M R Filip, A A Haghighirad et al. Cubic or orthorhombic? revealing the crystal structure of metastable black-phase CsPbI3 by theory and experiment. ACS Energy Lett, 3, 1787(2018).

    [53] J K Sun, S Huang, X Z Liu et al. Polar solvent induced lattice distortion of cubic CsPbI3 nanocubes and hierarchical self-assembly into orthorhombic single-crystalline nanowires. J Am Chem Soc, 140, 11705(2018).

    [54] V K Ravi, G B Markad, A Nag. Band edge energies and excitonic transition probabilities of colloidal CsPbX3 (X = Cl, Br, I) perovskite nanocrystals. ACS Energy Lett, 1, 665(2016).

    [55] C C Stoumpos, M G Kanatzidis. The renaissance of halide perovskites and their evolution as emerging semiconductors. Acc Chem Res, 48, 2791(2015).

    [56] C Katan, L Pedesseau, M Kepenekian et al. Interplay of spin–orbit coupling and lattice distortion in metal substituted 3D tri-chloride hybrid perovskites. J Mater Chem A, 3, 9232(2015).

    [57] L Zheng, D Zhang, Y Ma et al. Morphology control of the perovskite films for efficient solar cells. Dalton Trans, 44, 10582(2015).

    [58] C M M Soe, C C Stoumpos, B Harutyunyan et al. Room temperature phase transition in methylammonium lead iodide perovskite thin films induced by hydrohalic acid additives. ChemSusChem, 9, 2656(2016).

    [59] A Sharenko, C Mackeen, L Jewell et al. Evolution of iodoplumbate complexes in methylammonium lead iodide perovskite precursor solutions. Chem Mater, 29, 1315(2017).

    [60] D K Mohamad, B G Freestone, R Masters et al. Optimized organometal halide perovskite solar cell fabrication through control of nanoparticle crystal patterning. J Mater Chem C, 5, 2352(2017).

    [61] F Haque, M Wright, M A Mahmud et al. Effects of hydroiodic acid concentration on the properties of CsPbI3 perovskite solar cells. ACS Omega, 3, 11937(2018).

    [62] Y G Kim, T Y Kim, J H Oh et al. cesium lead iodide solar cells controlled by annealing temperature. Phys Chem Chem Phys, 19, 6257(2017).

    [63] D Y Heo, S M Han, N S Woo et al. Role of additives on the performance of CsPbI3 solar cells. J Phys Chem C, 122, 15903(2018).

    [64] Y Wei, W Li, S Xiang et al. Precursor effects on methylamine gas-induced CH3NH3PbI3 films for stable carbon-based perovskite solar cells. Solar Energy, 174, 139(2018).

    [65] F Wang, H Yu, H Xu et al. HPbI3: a new precursor compound for highly efficient solution-processed perovskite solar cells. Adv Funct Mater, 25, 1120(2015).

    [66] S Pang, Y Zhou, Z Wang et al. Transformative evolution of organolead triiodide perovskite thin films from strong room-temperature solid-gas interaction between HPbI3–CH3NH2 precursor pair. J Am Chem Soc, 138, 750(2016).

    [67] X Ding, H Chen, Y Wu et al. Triple cation additive NH3+C2H4NH2+C2H4NH3+-induced phase-stable inorganic α-CsPbI3 perovskite films for use in solar cells. J Mater Chem A, 6, 18258(2018).

    [68] S Xiang, Z Fu, W Li et al. Highly air-stable carbon-based α-CsPbI3 perovskite solar cells with a broadened optical spectrum. ACS Energy Lett, 3, 1824(2018).

    [69] Y Wang, T Zhang, F Xu et al. A Facile low temperature fabrication of high performance CsPbI2Br all-inorganic perovskite solar cells. Solar RRL, 2, 1700180(2018).

    [70] Q Ye, Y Zhao, S Mu et al. Cesium lead inorganic solar cell with efficiency beyond 18% via reduced charge recombination. Adv Mater, 31, e1905143(2019).

    [71] W Ke, I Spanopoulos, C C Stoumpos et al. Myths and reality of HPbI3 in halide perovskite solar cells. Nat Commun, 9, 4785(2018).

    [72] Y Pei, Y Liu, F Li et al. Unveiling property of hydrolysis-derived DMAPbI3 for perovskite devices: composition engineering, defect mitigation, and stability optimization. iScience, 15, 165(2019).

    [73] Y Wang, M I Dar, L K Ono et al. Thermodynamically stabilized β-CsPbI3–based perovskite solar cells with efficiencies >18%. Science, 365, 591(2019).

    [74] Y Wang, X Liu, T Zhang et al. the role of dimethylammonium iodide in CsPbI3 perovskite fabrication: additive or dopant. Angew Chem Int Ed, 58, 16691(2019).

    [75] H Meng, Z Shao, L Wang et al. Chemical composition and phase evolution in DMAI-derived inorganic perovskite solar cells. ACS Energy Lett, 5, 263(2020).

    [76] H Bian, H Wang, Z Li et al. Unveiling the effects of hydrolysis-derived DMAI/DMAPbIx intermediate compound on performance of CsPbI3 solar cells. Adv Sci, 10, 1902868(2019).

    [77] A Dutta, N Pradhan. Phase-stable red-emitting CsPbI3 nanocrystals: successes and challenges. ACS Energy Lett, 4, 709(2019).

    [78] T Zhang, M I Dar, G Li et al. Bication lead iodide 2D perovskite component to stabilize inorganic α-CsPbI3 perovskite phase for high-efficiency solar cells. Adv Sci, 3, e1700841(2017).

    [79] Y Wang, T Zhang, M Kan et al. Efficient α-CsPbI3 photovoltaics with surface terminated organic cations. Joule, 2, 2065(2018).

    [80] T Wu, Y Wang, Z Dai et al. Efficient and stable CsPbI3 solar cells via regulating lattice distortion with surface organic terminal groups. Adv Mater, 31, e1900605(2019).

    [81] Y Fu, M T Rea, J Chen et al. Selective stabilization and photophysical properties of metastable perovskite polymorphs of CsPbI3 in thin films. Chem Mater, 29, 8385(2017).

    [82] P Becker, J A Márquez, J Just et al. Low temperature synthesis of stable γ-CsPbI3 perovskite layers for solar cells obtained by high throughput experimentation. Adv Energy Mater, 9, 1900555(2019).

    [83] B Zhao, S Jin, S Huang et al. Thermodynamically stable orthorhombic γ-CsPbI3 thin films for high-performance photovoltaics. J Am Chem Soc, 140, 11716(2018).

    [84] C Liu, Y Yang, X Xia et al. Soft Template-controlled growth of high-quality CsPbI3 films for efficient and stable solar cells. Adv Energy Mater, 10, 1903751(2020).

    [85] L Liang, L Zhizai, F Zhou et al. Humidity-insensitive fabrication of efficient CsPbI3 solar cells in ambient air. J Mater Chem A, 7, 26776(2019).

    [86] H Wang, H Bian, Z Jin et al. Cesium lead mixed-halide perovskites for low-energy loss solar cells with efficiency beyond 17%. Chem Mater, 31, 6231(2019).

    [87] K Wang, Z Jin, L Liang et al. Chlorine doping for black γ-CsPbI3 solar cells with stabilized efficiency beyond 16%. Nano Energy, 58, 175(2019).

    [88] Z Yao, Z Jin, X Zhang et al. Pseudohalide (SCN-)-doped CsPbI3 for high performance solar cells. J Mater Chem C, 7, 13736(2019).

    [89] H Bian, Q Wang, S Yang et al. Nitrogen-doped graphene quantum dots for 80% photoluminescence quantum yield for inorganic γ-CsPbI3 perovskite solar cells with efficiency beyond 16%. J Mater Chem A, 7, 5740(2019).

    [90] L Liang, M Liu, Z Jin et al. Optical Management with nanoparticles for a light conversion efficiency enhancement in inorganic γ-CsPbI3 solar cells. Nano Lett, 19, 1796(2019).

    [91] H Bian, Q Wang, L Ding et al. Light management via tuning the fluorine-doped tin oxide glass haze-drives high-efficiency CsPbI3 solar cells. Phys Status Solidi A, 216, 1900602(2019).

    [92] Q Wang, X Zheng, Y Deng et al. Stabilizing the α-phase of CsPbI3 perovskite by sulfobetaine zwitterions in one-step spin-coating films. Joule, 1, 371(2017).

    [93] Z Jin, M Yuan, H Li et al. Graphdiyne: an efficient hole transporter for stable high-performance colloidal quantum dot solar cells. Adv Funct Mater, 26, 5284(2016).

    [94] Z Jin, A Wang, Q Zhou et al. Detecting trap states in planar PbS colloidal quantum dot solar cells. Sci Rep, 6, 37106(2016).

    [95] H Yao, F Zhou, Z Li et al. Strategies for improving the stability of tin-based perovskite (ASnX3) solar cells. Adv Sci, 10, 1903540(2020).

    [96] Z Jin, Q Zhou, Y Chen et al. Graphdiyne:ZnO nanocomposites for high-performance UV photodetectors. Adv Mater, 28, 3697(2016).

    [97] Y Jiang, J Yuan, Y Ni et al. Reduced-dimensional α-CsPbX3 perovskites for efficient and stable photovoltaics. Joule, 2, 1356(2018).

    [98] K Wang, Z Li, F Zhou et al. Ruddlesden–popper 2D component to stabilize γ-CsPbI3 Perovskite phase for stable and efficient photovoltaics. Adv Energy Mater, 9, 1902529(2019).

    [99] H Wang, S Xiang, W Li et al. Skillfully deflecting the question: a small amount of piperazine-1,4-diium iodide radically enhances the thermal stability of CsPbI3 perovskite. J Mater Chem C, 7, 11757(2019).

    [100] A Dutta, S K Dutta, S Das Adhikari et al. Phase-stable CsPbI3 nanocrystals: the reaction temperature matters. Angew Chem Int Ed, 57, 9083(2018).

    [101] J Xi, C Piao, J Byeon et al. Rational core-shell design of open air low temperature in situ processable CsPbI3 quasi-nanocrystals for stabilized p –i –n solar cells. Adv Energy Mater, 9, 1901787(2019).

    [102] B Li, Y Zhang, L Fu et al. Surface passivation engineering strategy to fully-inorganic cubic CsPbI3 perovskites for high-performance solar cells. Nat Commun, 9, 1076(2018).

    [103] J Gan, J He, R L Z Hoye et al. α-CsPbI3 colloidal quantum dots: synthesis, photodynamics, and photovoltaic applications. ACS Energy Lett, 4, 1308(2019).

    [104] X Zhang, R Munir, Z Xu et al. Phase transition control for high performance ruddlesden-popper perovskite solar cells. Adv Mater, 30, 1707166(2018).

    Zhizai Li, Zhiwen Jin. HI hydrolysis-derived intermediate as booster for CsPbI3 perovskite: from crystal structure, film fabrication to device performance[J]. Journal of Semiconductors, 2020, 41(5): 051202
    Download Citation