• Journal of Semiconductors
  • Vol. 44, Issue 8, 080201 (2023)
Zhimin Fang1, Jie Sun2, Shengzhong (Frank) Liu1、*, and Liming Ding2、**
Author Affiliations
  • 1Key Laboratory of Applied Surface and Colloid Chemistry (MoE), School of Materials Science and Engineering, Shaanxi Normal University, Xi’an 710119, China
  • 2Center for Excellence in Nanoscience (CAS), Key Laboratory of Nanosystem and Hierarchical Fabrication (CAS), National Center for Nanoscience and Technology, Beijing 100190, China
  • show less
    DOI: 10.1088/1674-4926/44/8/080201 Cite this Article
    Zhimin Fang, Jie Sun, Shengzhong (Frank) Liu, Liming Ding. Defects in perovskite crystals[J]. Journal of Semiconductors, 2023, 44(8): 080201 Copy Citation Text show less

    Abstract

    Introducing hydrogen bonding is also an effective method to passivate defects of perovskite films[1821]. Li et al. demonstrated that (4-aminobutyl) phosphonic acid hydrochloride (4-ABPACl) can heal the grain boundaries through strong hydrogen bondings between –PO(OH)2 and –NH3+ terminals and the perovskite (O–H···I and N–H···I) (Fig. 2(c))[18]. The modified perovskite gave higher PCEs. In addtion to these molecules with hydroxyl and amino groups, fluorides are also used as passivators by forming hydrogen bonding between fluorine atoms and MA+ or FA+[21]. These interactions can not only modulate crystallization to enhance crystallinity but also improve structural stability of perovskite films.

    In short, different defects in perovskite layer are discussed, and typical passivation strategies are highlighted. More investigation on the generation, location and types of defects would be helpful for making highly efficient and stable PSCs.

    Considering the structural feature and ionic nature of perovskite materials, most trap states are point defects and inevitably charged. The positively charged defects include uncoordinated Pb2+ ions, halogen vacancies, interstitial cations, AX, BX and BA anti-sites, while negatively charged defects are uncoordinated I ions, cation vacancies, interstitial halogen ions, XA, XB and AB anti-sites (Fig. 1(b))[4]. For anti-site defects, AX means A cation occupies X site. In addtition to charged defects, metal Pb clusters can also form in the perovskite films due to the reduction of uncoordinated Pb2+. The generation of a defect is directly determined by its formation energy. Once a defect forms, it will introduce additional transition energy level (Fig. 1(c))[5]. The transition levels near the conduction band minimum (CBM) or valence band maximum (VBM) edge are called shallow-level trap states, like vacancies and interstitial ions, whose formation energy is relatively low. The trapped carriers in shallow-level defects have high de-trapping probability to fall back into the CBM or VBM, thereby contributing little to nonradiative recombination. But defect migration can lead to undesirable local band bending, phase segregation, and current–voltage (JV) hysteresis, as well as degradation of perovskite films due to interfacial reactions[6]. Transition energy levels located within the middle third of forbidden band are identified as deep-level trap states, like uncoordinated ions, anti-sites and Pb clusters, which have higher formation energy. Different from shallow-level trap state, once a deep-level trap state is formed, it will serve as charge recombination centre and induce nonradiative recombination, thus decreasing the PCE. In order to diminish negative impact of various defects, the passivators should have following functions: (1) providing chemical bonding to deactivate the defects; (2) forming wide-bandgap phases or type-I heterojunction at grain boundaries and interfaces to suppress nonradiative recombination; (3) optimizing crystal orientation to reduce structural defects.

    Lewis acid/base coordination chemistry is widely employed to passivate uncoordinated defects. Lewis acid and Lewis base can accept and donate electrons, which means that they can passivate electron-rich and hole-rich defects via forming Lewis adducts, respectively. Fullerene (C60) and its derivatives PCBM are the most popular Lewis acids due to their excellent electron-accepting ability, making them powerful to passivate Lewis base type defects. Combining theoretical calculation and experimental characterizations, Xu et al. revealed that PCBM can effectively passivate IPb antisite and uncoordinated I to suppress nonradiative recombination and promote electron extraction (Fig. 2(a))[7]. In addition, a PCBM passivation layer can also inhibit anion migration that may result in JV hysteresis[8]. Compared with Lewis acid, Lewis base type passivators are more widely employed. Most Lewis bases base on nitrogen, oxygen, and sulphur atoms containing lone pair of electrons. Noel et al. first reported that orgainc molecules like thiophene and pyridine can serve as Lewis base to passivate uncoordinated Pb2+ (Fig. 2(b))[9]. Thiophene and pyridine can donate electrons to Pb2+ and form a coordinate covalent bond, neutralizing the excess positive charge of uncoordinated Pb2+. As a result, the treated perovskite films demonstrated suppressed nonradiative recombination and enhanced carrier lifetimes. Later, Lewis bases bearing amines, hydroxyl and carbonyl groups show effective passivation on uncoordinated Pb2+ defects[1013]. Moreover, some conjugated small molecules or polymers also show excellent passivation on Lewis acid traps. These molecules usually contain functional groups like thiophene, pyridine, amine, thiocyanato and so on, thus providing multiple passivation sites[1417].

    Organic-inorganic hybrid perovskite materials are ionic crystals with typical ABX3 perovskite structure[3], where A is monovalent cations like CH3NH3+ (MA+), NH2CHNH2+ (FA+) and Cs+, B is divalent cations (Pb2+ or Sn2+), and X is halogen anions (I, Br or Cl) (Fig. 1(a)). Owing to the diversity of compositions, various defects like vacancies, interstitial ions, and anti-sites can be generated in the solution-processed polycrystalline films. These defects mainly exist at grain boundaries and interfaces, which significantly influence the efficiency and stability of PSCs. Therefore, most works have focused on passivating defects to improve the device performance.

    (Color online) (a) Pb-I antisite defects passivated by PCBM. Reproduced with permission[7], Copyright 2015, Nature Publishing Group. (b) Uncoordinated Pb2+ passivated by thiophene. Reproduced with permission[9], Copyright 2014, American Chemical Society. (c) Hydrogen bondings (O–H···I and N–H···I) between 4-ABPACl and perovskite film. Reproduced with permission[18], Copyright 2015, Nature Publishing Group. (d) Halogen bonding between IPFB and halogen anion. Reproduced with permission[21], Copyright 2014, American Chemical Society. (e) Halogen vacancy passivated by K+. Reproduced with permission[28], Copyright 2018, Nature Publishing Group. (f) PEAI post-treatment on perovskite film. Reproduced with permission[29], Copyright 2019, Nature Publishing Group. (g) MA and I vacancies passivated by 3-(decyldimethylammonio)-propane-sulfonate zwitterion. Reproduced with permission[31], Copyright 2018, Wiley-VCH.

    Figure 2.(Color online) (a) Pb-I antisite defects passivated by PCBM. Reproduced with permission[7], Copyright 2015, Nature Publishing Group. (b) Uncoordinated Pb2+ passivated by thiophene. Reproduced with permission[9], Copyright 2014, American Chemical Society. (c) Hydrogen bondings (O–H···I and N–H···I) between 4-ABPACl and perovskite film. Reproduced with permission[18], Copyright 2015, Nature Publishing Group. (d) Halogen bonding between IPFB and halogen anion. Reproduced with permission[21], Copyright 2014, American Chemical Society. (e) Halogen vacancy passivated by K+. Reproduced with permission[28], Copyright 2018, Nature Publishing Group. (f) PEAI post-treatment on perovskite film. Reproduced with permission[29], Copyright 2019, Nature Publishing Group. (g) MA and I vacancies passivated by 3-(decyldimethylammonio)-propane-sulfonate zwitterion. Reproduced with permission[31], Copyright 2018, Wiley-VCH.

    Recently, halogen bonding has been used to enhance the efficiency of PSCs[2225]. Snaith et al. identified that the uncoordinated I at perovskite surface are responsible for the charge accumulation and consequent recombination losses. They modified perovskite surface by using iodopentafluorobenzene (IPFB), which can form halogen bonding with uncoordinated I to passivate these defects[22]. In IPFB, the fluorine atoms withdraw electrons out of the benzene ring, then the benzene ring in turn withdraws electrons from I, thereby leaving a partially positive charge around iodine atom (Fig. 2(d)). This electropositive region would interact with halogen anions to form strong halogen bonding. Similar results were achieved by using 4-iodo-2,3,5,6-tetrafluorobenzoic acid (I-TFBA)[25].

    (Color online) (a) SEM image for perovskite film containing PbI2 phases. (b) Type-I energy level alignment. Reproduced with permission[34], Copyright 2014, American Chemical Society. (c) PbI2 as charge-transport barrier at grain boundaries. Reproduced with permission[35], Copyright 2016, American Chemical Society. (d) 2D perovskites establish type-I heterojunction. Reproduced with permission[36], Copyright 2017, Nature Publishing Group. (e) DFT calculation for trap states on the surface of different crystallographic facets. Reproduced with permission[40], Copyright 2020, Nature Publishing Group.

    Figure 3.(Color online) (a) SEM image for perovskite film containing PbI2 phases. (b) Type-I energy level alignment. Reproduced with permission[34], Copyright 2014, American Chemical Society. (c) PbI2 as charge-transport barrier at grain boundaries. Reproduced with permission[35], Copyright 2016, American Chemical Society. (d) 2D perovskites establish type-I heterojunction. Reproduced with permission[36], Copyright 2017, Nature Publishing Group. (e) DFT calculation for trap states on the surface of different crystallographic facets. Reproduced with permission[40], Copyright 2020, Nature Publishing Group.

    Perovskite solar cell (PSC) is now a shining star in photovoltaics field[1]. Benefiting from excellent optoelectronic properties of perovskite materials, the certified power conversion efficiency (PCE) of PSCs has reached 26.0%[2], showing great potential for commercialization. In essence, the efficiency of solar cells is determined by the radiative and nonradiative recombination of photogenerated charge carriers. The unfavorable nonradiative recombination mainly assisted by the trap states leads to severe charge carrier loss and thus unsatisfactory efficiency.

    Ionic passivation bases on electrostatic interaction between passivators and the charged defects in perovskite films. In earlier studies, Na+ and K+ cations were selected to passivate defects like MA vacancies and uncoordinated halogen ions via ionic bonding[26]. Owing to small ionic radius, Na+ and K+ cations are unfavorable to occupy A-sites, but they can occupy the interstitial sites in perovskite lattice to suppress ion migration and eliminate hysteresis[27]. Some research suggested that Na+ and K+ cations can also assemble at grain boundaries and surface, forming KI or KBr to passivate defects and improve carrier lifetimes (Fig. 2(e))[28]. Since You et al. reported highly efficient PSCs through phenylethylammonium iodide (PEAI) post-treatment (Fig. 2(f))[29], organic ammonium halides have been intensively used as effective passivators. The halogen ions can fill in halogen vacancies, while NH3+ terminal can interact with negatively charged defects in perovskite films, suppressing nonradiative recombination[30]. Beyond these materials, zwitterions can simultaneously passivate positively and negatively charged defects (Fig. 2(g))[31]. Usually, zwitterions contain positively charged ammonium group (NH3+) and negatively charged carboxylate (COO), phosphate (PO42) or sulfonic (SO3) groups[3133].

    Introducing wide-bandgap phases to form a type-I heterojunction at grain boundaries and interfaces can effectively suppress nonradiative recombination. The most common approach is to introduce excess PbI2 to perovskite films (Fig. 3(a))[34]. The type-I energy alignment (Fig. 3(b)) between PbI2 and perovskite can hold back charge transfer at grain boundaries (Fig. 3(c)), thereby suppressing nonradiative recombination[35]. Excess PbI2 might also induce film degradation. In view of this, converting excess PbI2 to two-dimensional (2D) perovskites by using long-chain ammoniums becomes a frontier technique. The 2D perovskites can establish type-I heterojunction to suppress nonradiative recombination (Fig. 3(d))[36]. Moreover, the strong hydrophobicity of 2D perovskites would protect 3D perovskite from water erosion. Based on this technique, scientists have made highly efficient and stable PSCs. It should be noted that the issues for 2D or quasi-2D perovskites are crystal orientation and layer-layer distance. The undesirable crystal orientation and long layer-layer distance would definitely block charge transport in the film and at interfaces, which might sacrifice Jsc and FF[37]. Therefore, optimizing crystal orientation and layer-layer distance of 2D or quasi-2D perovskites is critical.

    Though using passivators is promising to deactivate various defects, the initial generation of defects should not be ignored. Perovskite layers made from solution are polycrystalline films that inevitably present grain boundaries. Since grain boundaries contain various defects, increasing grain size to reduce grain boundaries is vital for high-performance PSCs. Moreover, the optoelectronic anisotropy of perovskite films is quite important. Different crystal facets might present different trap states[38, 39]. Single-crystal perovskites show less defects. Zheng et al. investigated the dependence of defect density on crystallographic facets via density functional theory (DFT) calculations based on a cubic FAPbI3 model[40]. They revealed that there are no defects on (100) facet, while the (110), (111) and (211) facets present defects (Fig. 3(e)). By incorporating a trace amount of surface-anchoring alkylamine ligands (AALs) into perovskite precursor, they achieved (100)-oriented crystal growth with lower defect density and enhanced carrier mobility. Similar phenomenon was observed in Pb-Sn mixed perovskites[41, 42]. Optimizing crystal orientation and exploring facet-dependent defect variation have attracted interests[4345].

    (Color online) (a) Crystal structure for perovskite materials. Reproduced with permission[3], Copyright 2019, the Royal Society of Chemistry. (b) Calculated transition energy levels of different defects in MAPbI3. The corresponding formation energies are shown in parentheses. Reproduced with permission[4], Copyright 2014, Wiley-VCH. (c) Charge recombination pathways in perovskite films. Reproduced with permission[5], Copyright 2018, the Royal Society of Chemistry.

    Figure 1.(Color online) (a) Crystal structure for perovskite materials. Reproduced with permission[3], Copyright 2019, the Royal Society of Chemistry. (b) Calculated transition energy levels of different defects in MAPbI3. The corresponding formation energies are shown in parentheses. Reproduced with permission[4], Copyright 2014, Wiley-VCH. (c) Charge recombination pathways in perovskite films. Reproduced with permission[5], Copyright 2018, the Royal Society of Chemistry.

    References

    [1] L X Zhang, X Y Pan, L Liu et al. Star perovskite materials. J Semicond, 43, 030203(2022).

    [2] NREL . "Best Research-cell Efficiency Chart, " www. nrel. gov/pv/ cell- effic iency. html Accessed: Jun 2023.

    [3] Y Y Zhou, Y X Zhao. Chemical stability and instability of inorganic halide perovskites. Energy Environ Sci, 12, 1495(2019).

    [4] W J Yin, T T Shi, Y F Yan. Unique properties of halide perovskites as possible origins of the superior solar cell performance. Adv Mater, 26, 4653(2014).

    [5] C X Ran, J T Xu, W Y Gao et al. Defects in metal triiodide perovskite materials towards high-performance solar cells: origin, impact, characterization, and engineering. Chem Soc Rev, 47, 4581(2018).

    [6] B Chen, P N Rudd, S Yang et al. Imperfections and their passivation in halide perovskite solar cells. Chem Soc Rev, 48, 3842(2019).

    [7] J X Xu, A Buin, A H Ip et al. Perovskite-fullerene hybrid materials suppress hysteresis in planar diodes. Nat Commun, 6, 7081(2015).

    [8] Y C Shao, Z G Xiao, C Bi et al. Origin and elimination of photocurrent hysteresis by fullerene passivation in CH3NH3PbI3 planar heterojunction solar cells. Nat Commun, 5, 5784(2014).

    [9] N K Noel, A Abate, S D Stranks et al. Enhanced photoluminescence and solar cell performance via lewis base passivation of organic-inorganic lead halide perovskites. ACS Nano, 8, 9815(2014).

    [10] F Wang, W Geng, Y Zhou et al. Phenylalkylamine passivation of organolead halide perovskites enabling high-efficiency and air-stable photovoltaic cells. Adv Mater, 28, 9986(2016).

    [11] S Yang, J Dai, Z H Yu et al. Tailoring passivation molecular structures for extremely small open-circuit voltage loss in perovskite solar cells. J Am Chem Soc, 141, 5781(2019).

    [12] W Z Li, H P Dong, X D Guo et al. Graphene oxide as dual functional interface modifier for improving wettability and retarding recombination in hybrid perovskite solar cells. J Mater Chem A, 2, 20105(2014).

    [13] Y Cai, J Cui, M Chen et al. Multifunctional enhancement for highly stable and efficient perovskite solar cells. Adv Funct Mater, 31, 2005776(2021).

    [14] Y Z Lin, L Shen, J Dai et al. π-conjugated lewis base: Efficient trap-passivation and charge-extraction for hybrid perovskite solar cells. Adv Mater, 29, 1604545(2017).

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

    [16] L J Zuo, H X Guo, D W deQuilettes et al. Polymer-modified halide perovskite films for efficient and stable planar heterojunction solar cells. Sci Adv, 3, e1700106(2017).

    [17] D Q Bi, C Y Yi, J S Luo et al. Polymer-templated nucleation and crystal growth of perovskite films for solar cells with efficiency greater than 21%. Nat Energy, 1, 16142(2016).

    [18] X Li, M I Dar, C Y Yi et al. Improved performance and stability of perovskite solar cells by crystal crosslinking with alkylphosphonic acid ω-ammonium chlorides. Nat Chem, 7, 703(2015).

    [19] X Y Meng, J B Lin, X Liu et al. Highly stable and efficient FASnI3-based perovskite solar cells by introducing hydrogen bonding. Adv Mater, 31, 1903721(2019).

    [20] R Wang, J J Xue, K L Wang et al. Constructive molecular configurations for surface-defect passivation of perovskite photovoltaics. Science, 366, 1509(2019).

    [21] C Gong, C Zhang, Q X Zhuang et al. Stabilizing buried interface via synergistic effect of fluorine and sulfonyl functional groups toward efficient and stable perovskite solar cells. Nanomicro Lett, 15, 17(2022).

    [22] A Abate, M Saliba, D J Hollman et al. Supramolecular halogen bond passivation of organic-inorganic halide perovskite solar cells. Nano Lett, 14, 3247(2014).

    [23] P Metrangolo, L Canil, A Abate et al. Halogen bonding in perovskite solar cells: A new tool for improving solar energy conversion. Angew Chem Int Ed, 61, e202114793(2022).

    [24] G H Ren, W B Han, Q Zhang et al. Overcoming perovskite corrosion and de-doping through chemical binding of halogen bonds toward efficient and stable perovskite solar cells. Nano-Micro Lett, 14, 175(2022).

    [25] C Y Zhang, X Q Shen, M J Chen et al. Constructing a stable and efficient buried heterojunction via halogen bonding for inverted perovskite solar cells. Adv Energy Mater, 13, 2203250(2023).

    [26] N X Li, S X Tao, Y H Chen et al. Cation and anion immobilization through chemical bonding enhancement with fluorides for stable halide perovskite solar cells. Nat Energy, 4, 408(2019).

    [27] J Cao, S X Tao, P A Bobbert et al. Interstitial occupancy by extrinsic alkali cations in perovskites and its impact on ion migration. Adv Mater, 30, 1707350(2018).

    [28] M Abdi-Jalebi, Z Andaji-Garmaroudi, S Cacovich et al. Maximizing and stabilizing luminescence from halide perovskites with potassium passivation. Nature, 555, 497(2018).

    [29] Q Jiang, Y Zhao, X W Zhang et al. Surface passivation of perovskite film for efficient solar cells. Nat Photon, 13, 460(2019).

    [30] R X Lin, J Xu, M Y Wei et al. All-perovskite tandem solar cells with improved grain surface passivation. Nature, 603, 73(2022).

    [31] X P Zheng, Y H Deng, B Chen et al. Dual functions of crystallization control and defect passivation enabled by sulfonic zwitterions for stable and efficient perovskite solar cells. Adv Mater, 30, 1803428(2018).

    [32] X P Zheng, B Chen, J Dai et al. Defect passivation in hybrid perovskite solar cells using quaternary ammonium halide anions and cations. Nat Energy, 2, 17102(2017).

    [33] J H Kim, Y R Kim, B Park et al. Simultaneously passivating cation and anion defects in metal halide perovskite solar cells using a zwitterionic amino acid additive. Small, 17, 2005608(2021).

    [34] Q Chen, H P Zhou, T B Song et al. Controllable self-induced passivation of hybrid lead iodide perovskites toward high performance solar cells. Nano Lett, 14, 4158(2014).

    [35] T J Jacobsson, J P Correa-Baena, E H Anaraki et al. Unreacted PbI2 as a double-edged sword for enhancing the performance of perovskite solar cells. J Am Chem Soc 2016, 138, 1(1033).

    [36] Z P Wang, Q Q Lin, F P Chmiel et al. Efficient ambient-air-stable solar cells with 2D–3D heterostructured butylammonium-caesium-formamidinium lead halide perovskites. Nat Energy, 2, 17135(2017).

    [37] D N Yu, Q Wei, H S Li et al. Quasi-2D bilayer surface passivation for high efficiency narrow bandgap perovskite solar cells. Angew Chem Int Ed, 61, e202202346(2022).

    [38] S Y Leblebici, L Leppert, Y B Li et al. Facet-dependent photovoltaic efficiency variations in single grains of hybrid halide perovskite. Nat Energy, 1, 16093(2016).

    [39] Z M Fang, N Yan, S Z Liu. Modulating preferred crystal orientation for efficient and stable perovskite solar cells—From progress to perspectives. InfoMat, 4, e12369(2022).

    [40] X P Zheng, Y Hou, C X Bao et al. Managing grains and interfaces via ligand anchoring enables 22.3%-efficiency inverted perovskite solar cells. Nat Energy, 5, 131(2020).

    [41] N Yan, X D Ren, Z M Fang et al. Ligand-anchoring-induced oriented crystal growth for high-efficiency lead-tin perovskite solar cells. Adv Funct Mater, 32, 2201384(2022).

    [42] J G Li, N Yan, Z M Fang et al. Alkyl diamine-induced (100)-preferred crystal orientation for efficient Pb–Sn perovskite solar cells. ACS Appl Energy Mater, 5, 6936(2022).

    [43] C Q Ma, F T Eickemeyer, S H Lee et al. Unveiling facet-dependent degradation and facet engineering for stable perovskite solar cells. Science, 379, 173(2023).

    [44] C Q Ma, M Grätzel, N G Park. Facet engineering for stable, efficient perovskite solar cells. ACS Energy Lett, 7, 3120(2022).

    [45] C Luo, G H J Zheng, F Gao et al. Facet orientation tailoring via 2D-seed-induced growth enables highly efficient and stable perovskite solar cells. Joule, 6, 240(2022).

    Zhimin Fang, Jie Sun, Shengzhong (Frank) Liu, Liming Ding. Defects in perovskite crystals[J]. Journal of Semiconductors, 2023, 44(8): 080201
    Download Citation