Huiyuan
Cheng†
abc,
Yaomiao
Feng†
abc,
Yu
Fu
bd,
Yifan
Zheng
ce,
Yuchuan
Shao
*ce and
Yang
Bai
*abd
aAustralian Institute for Bioengineering and Nanotechnology, The University of Queensland, Brisbane, St Lucia, Queensland 4072, Australia. E-mail: y.bai@uq.edu.au
bFaculty of Materials Science and Engineering/Institute of Technology for Carbon Neutrality, Shenzhen Institute of Advanced Technology, Chinese Academy of Sciences, Shenzhen 518055, China. E-mail: y.bai@siat.ac.cn
cKey Laboratory of Materials for High-Power Laser, Shanghai Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Shanghai 201800, China. E-mail: shaoyuchuan@siom.ac.cn
dShenzhen Key Laboratory of Energy Materials for Carbon Neutrality, Shenzhen 518055, China
eSchool of Physics and Optoelectronic Engineering, Hangzhou Institute for Advanced Study, UCAS, Hangzhou 310024, China
First published on 14th July 2022
Light-emitting diodes (LEDs) based on metal halide perovskites have shown great promise for next-generation display technology as they offer high color purity, satisfy Rec. 2020, and have low-cost solution processability. Moreover, metal halide perovskites exhibit extraordinary optoelectronic properties such as high photoluminescence quantum yields (PLQYs), feasible spectral tunability, narrow emission and high charge-carrier mobility, which has led to a rapid increase in the external quantum efficiencies (EQEs) of up to 28% for perovskite light-emitting diodes (PeLEDs) over the past few years. Nevertheless, further increase in efficiency is impeded in these state-of-the-art devices due to the presence of non-radiative recombination losses, which also limits their operational stability. In this review, we provide a fundamental analysis of the predominant pathways that induce non-radiative recombination losses in PeLEDs, followed by a discussion on what and how reliable characterization techniques could be used to evaluate such losses. We also summarize and critically assess the most recent advances in suppressing non-radiative recombination in PeLEDs. Finally, we discuss the remaining challenges and outline future directions that aim to minimize non-radiative recombination losses and boost the efficiency of PeLEDs towards their radiative limit.
Despite such impressive advances, the electroluminescence performance of PeLEDs, particularly blue PeLEDs, still lags behind those of commercial LED devices, such as OLEDs.5 One of the key reasons is that the presence of non-radiative recombination is still prominent in PeLEDs, which impedes further efficiency enhancement.6,7 The predominant pathways that induce non-radiative recombination losses mainly include trap-assisted recombination, Auger recombination and interface-induced recombination.6
Unlike the perovskite films for photovoltaic application, the emissive perovskite films employed in LEDs need to be very thin (several ten nanometres) and are usually comprised of considerably smaller sized polycrystals or nanocrystals, which aims to spatially confine the charge carriers and increase the radiative bimolecular recombination rates in the emitter layers. Nevertheless, this results in significantly increased surface area-to-volume ratio and larger density of defects in perovskite emitters.8,9 Several types of point defects such as halide and A-site vacancies, Pb-halide anti-sites and iodine interstitials as well as the defects at the device interfaces7 have proven to be associated with trap states that induce the trap-assisted non-radiative recombination,2,10,11 which is the key reason for the luminous efficiency loss at relatively low charge-carrier densities. While Auger recombination becomes dominant in PeLEDs at higher charge-carrier densities causing efficiency roll-off, despite that three-dimensional perovskites have been demonstrated to be promising for high-brightness LEDs.12 In addition, the non-optimal energy level alignment together with unbalanced charge injection within PeLEDs leads to carrier accumulation at the interfaces, which increases the probability of Auger recombination and deteriorate the device performance.13–16 Furthermore, the presence of abundant defects and non-radiative recombination in PeLEDs will also impair their operational stability. The defects, acting as ionic shuttles, were shown to boost the migration of ions such as halides and thus cause the shift of emission wavelength, which reduce the color purity and long-term spectra stability of PeLEDs.17,18 Non-radiative recombination, particularly Auger recombination, generate significant amount of Joule heat, which also accelerates the device degradation.19,20
Therefore, the control of non-radiative recombination processes in PeLEDs is crucial for achieving high electroluminescence performance and stability. Tremendous efforts have recently been devoted to exploring effective strategies for suppressing non-radiative recombination losses in PeLEDs, including crystal size manipulation,21–24 exciton binding energy tuning,25,26 defect passivation,27,28 interface engineering,29,30 and so forth. Yet there remains a significant gap in terms of the luminous efficiency and device lifetime to meet the requirements of commercialization. It is thus in urgent need of a systematic and in-depth overview focusing on the understanding and latest advances in the mitigation of non-radiative recombination in PeLEDs, which serves as a guide for their further development towards practical applications.
In this review, we first present a fundamental understanding of the origin and nature of the non-radiative recombination losses in PeLEDs, which mainly covers the trap-assisted recombination, Auger recombination and interface-induced non-radiative recombination. Then, advanced characterization techniques used to quantify these non-radiative recombination losses in PeLEDs are discussed, and recent advances in suppressing such losses are also critically assessed. Finally, we give our perspectives on future work that potentially could minimize non-radiative recombination losses and further improve the luminous efficiency and stability of PeLEDs to meet the needs of commercialization.
(i) Charge carrier balance. The ratio of electrons and holes injected into emitter layer should be closed to 1.
(ii) Leakage current. Charge carriers across the device without forming a correlated electron–hole pairs should be minimized.
(iii) Recombination. Non-radiative recombination processes should be suppressed.
(iv) Light extraction. Outcoupling of generated photons from device to free area should be maximized.
Therefore, the EQE of the PeLEDs can be defined as31
(1) |
(2) |
Trap-assisted recombination, also known as Shockley–Read–Hall (SRH) recombination (Fig. 1(b)), is a two-step transition via a defect within the band gap.34 Firstly, an electron (or a hole) is trapped by an energy state in the forbidden region which is introduced through defects in the crystal lattice. Then, if a hole (or an electron) moves up to the same energy state before the electron is thermally re-emitted into the conduction band, the trap-assisted recombination occurs. The trap-assisted non-radiative recombination discussed here refer to that induced only by the bulk or surface defects of perovskite emitter layer including polycrystalline thin films and colloidal nanocrystals.
Similar to conventional semiconductors, the typical defects in perovskite can be classified as point defects (including vacancies, interstitials, anti-sites) and structural defects (including dislocations, grain boundaries).35 Defects can cause trap states near the band edge (shallow defects) or within the bandgap (deep defects) to capture charge carriers. Among conceivable defects, interstitial and anti-site defects can form deep trap states in the electronic structure. Deep trap is the main pathway for the loss of charge carriers and non-radiative recombination in perovskite materials.10 However, these defects are normally absence in the perovskite lattice since they have high formation energies.36 Only vacancy defects with sufficiently low formation energies are exclusively observed.9 Fortunately, these shallow defects usually occur on the surface and can be easily passivated by simple chemical modification, such as the use of ligands in nanocrystals and additives in polycrystalline film. Normally, the lattice parameter decreases with halide changing from I− to Br− to Cl− in metal halide perovskite. This leads to stronger interactions between Pb2+ dangling bonds when halide vacancies form, which gives rise to deeper trap level for the halide vacancy when combined with lower electron affinity. Thus, Cl-based perovskite are intended to form deep trap and low PLQY, while Br-based and I-based perovskite primarily form shallow traps and high PLQY.35 Therefore, much attentions have been paid to decreasing the trap density and reducing non-radiative losses in recently years, which will be discussed further in the following sections.
Auger recombination is another unavoidable non-radiative recombination, during which the energy released by electron–hole recombination is transferred to other charge carrier, rather than being released as photons or heat energy.37 Auger recombination is a multiexcitons decay process (Fig. 1(c)) and highly dependent on the density of charge carriers, which usually leads to significant photoluminescence quenching at high carrier concentration (>1017 cm−3).20,38 The simplest charged state is a trion, which comprises a neutral exciton and either an extra hole (positive trion, X+) or an extra electron (negative trion, X−).39 Meanwhile, Auger decay can also generate neutral multiexcitons such as biexcitons. All these Auger processes could cause non-radiative recombination and thus deteriorate the PLQY and PL lifetime, which have a paramount impact on the performance and stability of PeLEDs.
Reducing the Auger recombination is recognized as one of the primary approaches for improving PeLED performance at high carrier concentration. However, the Auger recombination rate of halide perovskites (≈10−28 cm6 s−1) is much faster than that of traditional III–V (GaAs, ≈10−30 cm6 s−1)40 or II–VI (CdTe, ≈10−32 cm6 s−1)41 semiconductor. The unexpectedly strong Auger recombination rate in perovskite is attributed to a coincidental resonance of the bandgap with states that are roughly one bandgap away from the band edges.42 The strong Coulomb interaction between the hole and electron leads to a non-uniform carrier distribution, because a carrier is more likely to be surrounded by the opposite charged carriers, which strengthens Auger recombination at the same position.43,44 In addition, the distortions in metal-halide lattice affects the state distribution, and contributes significantly to the high Auger rate.42 All these factors together contribute to ultra-strong Auger recombination in perovskite materials. In addition, unbalanced injection of electrons and holes at interfaces can cause Auger recombination, which will be described in the next section.
Interface-induced recombination is usually caused either by the presence of interfacial defects or the mismatched energy-level alignment among different layers. In an ideal case as exhibited in Fig. 1(d), the perfect selective layers are expected to have a smooth morphology that leads to a defect-free interface with emitter layer, and in the meantime, allow either electrons or holes to be injected through while block the opposite charged carriers at the interface. Nevertheless, in reality, there are usually numerous defects at the interfaces between different layers, causing non-radiative recombination (Fig. 1(e)).13,45 The formation of such interfacial defects is mostly correlated with the poor affinity of each layer at the interfaces, which increases the difficulties in film deposition and thus may lead to film discontinuity with pinholes.46 Such interfacial defects not only exacerbate non-radiative recombination,7,45 but also impose an adverse impact on the efficient injection of charge carriers.47 The losses associated with interface-induced recombination could be mitigated by defect passivation and interfacial engineering.31,46 In addition to the interface defects, the imbalanced injection of electrons and holes due to the non-optimal energy level alignment is another key issue that can cause interface-induced non-radiative recombination, particularly Auger recombination in PeLEDs.48,49 Recent studies indicated that even a slight imbalance between electron and hole injection can lead to a serious efficiency roll-off.6,50 A typical example is illustrated in Fig. 1(f), where there is an energy barrier present at the interface of electron transport layer (ETL) and emitter layer. Under a forward bias, the holes could pass through the interface of hole transport layer (HTL) and emitter layer easily since the energy barrier is negligible. The holes then accumulate at the ETL/emitter layer interface due to the presence of a large energy barrier in between. In the meantime, the injected electrons also accumulate at the interface of ETL and emitter layer due to emitter the presence of a significant energy barrier. The simultaneous accumulation of electrons and holes at the ETL/emitter layer interface facilitates the occurring of Auger recombination, which enables the third electrons with sufficiently high energy to inject into emitter layer and radiatively recombine with holes in the same layer.48 This also applies to the situation where there is a considerable energy barrier at the HTL/emitter layer interface.51 Therefore, device engineering is of paramount to suppress interface-induced recombination by achieving an optimal energy level alignment and balanced carrier injections.
(3) |
The non-radiative recombination can be suppressed at low temperature, as the traps are frozen out in such situation. The fluctuation concentration of defects may be responsible for reversible PL intensity loss with elevated temperatures. Therefore, Eb can represent the activation energy for charge carrier trapping.55 If the fitted trapping activation energy matches the non-radiative recombination defect energy level, it can reflect the behaviour of this defect thermal quenching in emission.55,56 By fitting the reversible PL loss of the CsPbBr3 nanocrystals, the activation energy of Br vacancies defects can be obtained (246 meV), which is consistent with density functional theory (DFT) calculations.55 The agreement between experimental data and calculated trap energies leads us to identify the mechanism of reversible PL loss in CsPbBr3 nanocrystals as thermal electron occupation of halogen vacancy centers.27 By passivating the nanocrystals with didodecyl dimethylammonium fluoride (DDAF), the activation energy for trapping value rose with the increasement of DDAF adding, indicating that the surface bromine vacancies are effectively passivated.55 Moreover, it should be noted that the sub-bandgap and phase state often changes with the temperature transformation, owing to the fragile nature of perovskite. As shown in Fig. 2(a), band-tail PL emission dominates at very low temperatures and phase transition occurs at around 150 K. The trap energies and densities are slightly different in these phases, which also influence the charge carrier recombination process.57,58 Explicit interpretation could be obtained when the analysis considers the phase transition or take into consideration of charge-carrier dynamics in certain temperature regimes.59
Fig. 2 (a) False-colour plot of PL spectra of FA0.95Cs0.05PbI3 thin film at low fluence 3.6 nJ cm−2 as a function of temperature between 4 and 295 K. A phase transition occurs at 150 K. Reproduced with permission from ref. 54. Copyright © 2020, Wiley-VCH GmbH; (b) excitation-intensity-dependent PLQY of the multiple quantum wells perovskite films under continuous-wave laser excitation. Reproduced with permission from ref. 12. Copyright © 2018, The author(s), Springer Nature; (c) TRPL dynamics of polycrystalline perovskite, and (d) nanocrystals obtained at different excitation densities. The insets show the PL dynamics recorded at high excitation densities where the dependence of the lifetime on the excitation density is opposite than the one observed at low excitation densities. The arrows indicate an increase in the excitation density. Reproduced with permission from ref. 67. Copyright © 2018, American Chemical Society; (e) diagram illustrated the feature of TA spectra during the hot carrier cooling process. (f) TA spectra at different delay times for the quasi-2D perovskite film. (g) Schematic diagram of charge carrier behaviour after excitation. The carrier recombination progress can be divided into five stages: I, carrier formation; II, exciton transfer; III, charge transfer; IV, reverse charge transfer; V, continuous charge transfer and recombination. (f) and (g) were reproduced with permission from ref. 89. Copyright © 2020, The Author(s), Springer Nature. |
(4) |
The PLQY of different samples can be used to quantify the recombination losses in the bulk, at the interfaces, and/or metal contacts.61 For instance, the PLQY of perovskite thin film increases from 12% to 78% after adding dual-additive (18-crown-6 and poly(ethylene glycol) methyl ether acrylate), which indicates the dual additives containing C–O–C bonds can suppress non-radiative recombination in the bulk of perovskite.4,62 Moreover, the comparison of PLQY values between as-synthesized colloidal nanocrystals and as-deposited nanocrystal thin film could reflect the non-radiative loss caused by nanocrystal assembly.
In general, the PLQY depends on a few factors, including the quality of the perovskite polycrystalline film or nanocrystals, the value of Eb,63 the efficiency energy transfer among different phases,20 the energy level of transfer layers, and additional recombination pathways at the perovskite/transfer layers.
(5) |
The average lifetime can be calculated as followed
(6) |
In case of neat perovskite film, the PL decays follow single exponential behaviour under sufficiently low excitation fluence, which has been assigned to a first-order loss process, such as trap-assisted recombination. The lifetime is relatively low in this situation because the trap-assisted non-radiative recombination is dominant. By increasing the fluences to a higher level, the saturation of trap states deactivates the non-radiative relaxation pathways, giving rise to an increase in PL at first. And then bimolecular radiative and Auger recombination enhance and successively become dominant in the PL dynamics. At high excitation density, the amplitude of these fast component increases, resulting in the decreasing of effective lifetime. Accordingly, multi-exponential fitting is applied to analyse the carrier dynamics under high excitation.20 That is the reason why, as seen in the Fig. 2(c and d), the dynamic slows down at first and subsequently speeds up as the increasement of excitation density, both in polycrystalline and nanocrystals samples.67
Normally, the fast delay (A1, τ1) is related to trap-assisted recombination and the slow delay (A2, τ2) associates with the radiative recombination. The variation of these parameters before and after the modification of samples can reflect the behaviour of charge carrier recombination. The lessened A1 indicates less trap-assisted recombination at grains’ boundaries/interfaces or nanocrystals’ surface, while increased A2 manifests more radiative recombination inside the perovskite, ultimately contributing to electroluminescence efficiency.66,68 Therefore, the longer lifetime indicates stronger radiative recombination, while the shorter lifetime related to the non-radiative recombination of trap capture or Auger process.17 Normally, the faster PL decay time and lower PLQY in the case of chloride-based nanocrystals suggest that they exhibit higher non-radiative rates as compared to the iodide and bromine-based nanocrystals.69,70 The mixed anion nanocrystals shows dynamics processes between two pure-anion perovskite.71
The TRPL can not only reflect the traps-assisted and Auger recombination, but it can also reflect the non-radiative loss at the interfaces.15,72 By comparing the TRPL spectra of the perovskite with and without HTL/ETL, we can measure the non-radiative recombination events at the two contact interfaces. If the adding of charge-transport layers does not influence lifetime of carriers, it can indicate non-radiative recombination events at the two contact interfaces are insignificant.62 For example, the TRPL of pure perovskite films with the polymer component (poly(2-hydroxyethyl methacrylate)) and such film on the CTL are almost the same, indicating the interfacial luminescence quenching processes are insignificant.
TRPL is a sensitive measurement for analysing the carriers’ behaviour. But it is difficult to measure the process of exciton formation or decay to free-carrier dynamics, because the TRPL is sensitive to the presence of both free carriers and excitons.73,74 Also, PL lifetime for each bandgap species (n = 1, 2, 3,…) in quasi-two-dimensional (quasi-2D) perovskite cannot extract from TRPL to illustrate energy transfer kinetics because of the limitation of this instrument.75 Therefore, to analyse the carrier dynamic process in perovskites, TRPL should be combined with other precision measurements.
For perovskite nanocrystals, the early time (Δt < τcool) TA spectra exhibit a typical trait of Coulomb-interaction-induced red shift of the band edge transition, whose shape look as an asymmetric derivative-like feature with PA at lower energies and GSB at higher as shown in Fig. 2(e).71,76 This process is in order of sub ps or even fs, which is shorter than excitons cooling and exciton recombination.31 After Δt > τcool, the derivative feature in the TA spectra disappears due to state filling, in which PA signal eventually decays and a very strong band-edge bleaching signal occurs. Further, the charge carriers are located at the band-edge after hot-carrier cooling. During this time, Auger decay is domain over radiative recombination if a nanocrystal has more than one exciton.77 After Auger processes, band to band radiative recombination and trap-assisted non-radiative recombination can occur simultaneously.64 Hence, in order to observe different state of carrier dynamic process, the delay time of TA should be carefully determined. Tracking GSB peak after the carrier cooling process as function of time and fitting the data with the eqn (5) can get the information of the carrier recombination process. Consistent with the TRPL decay, the fast decay and long decay components are associated with non-radiative and radiative recombination, respectively. Normally, post treating the nanocrystal can prolong the long decay, which indicates the passivation of defects.19
For quasi-2D perovskite thin films, the peaks of species (n = 1, 2, 3,…) could be easily distinguished in TA spectrum (Fig. 2(f)), because the peak position of GSB peaks are in agreement with peaks in the steady-state absorption spectrum.75,78 In general, several peaks can be identified in the spectra, suggesting the quasi-2D perovskite films are not single-phase, but a mixture of n values. If the major peak directly mirrors the lowest bandgap component in the steady-state PL spectrum, it might imply that photoexcitation transfer is occurring in these materials. Therefore, the TA spectra can reflect the process of cascade energy transfer from smaller-n phases to larger-n phases in quasi-2D perovskites (Fig. 2(g)),75,79 or from quasi-2D to 3D phase perovskite in hybrid systems.66
Other transient characterization measurements based on frequency, such as terahertz photoconductivity spectra (THz), can also monitor charge carrier recombination in a sub-femtosecond timescale.80 In addition, photoinduced time-resolved microwave conductivity (TRMC) can be applied as a complementary technique to probe charge carrier lifetimes.81 Despite many theories that underpin these studies, the time delayed collection field aim to identify the carrier dynamic process after the excitation pulse, which can provide fundamental information of radiative and non-radiative recombination.
Apart from the methods mentioned above, other electrical ways used in QLEDs and PSCs can also be applied to PeLEDs. Among these methods, space charge limited current (SCLC) has been widely used in probing the trap density and carrier mobility in perovskite.75,86 This measurement records the current density with varying the applied voltage in the hole-only or electron-only device. The trap density can be calculated from the lower value of trap-filled limit voltage (VTFL). Thermal admittance spectroscopy (TAS)87 is another widely used technique to quantify the trap density of state (tDOS) by measuring the device capacitance and conductance as a function of frequency and temperature. Additionally, it could provide an energy level profile of different traps. Time-resolved transient electroluminescence (TREL)88 has been widely used in QLEDs to investigate the pertinent kinetic processes in operational devices. This method can also be utilized in understanding the dynamics in operational PeLEDs to optimize the performance.
Fig. 3 (a) Schematic illustrations decrease the grain size of MAPbBr3 to enhance luminescent properties. Reproduced with permission from ref. 8. Copyright © 2015, American Association for the Advancement of Science; (b) Schematic illustration of the crystal growth and surface passivation of CsPbI3−xBrx polycrystalline thin films regulated by poly(vinylidenefluoride-co-hexafluoropropylene) and L-arginine. Reproduced with permission from ref. 23. Copyright © 2020 Wiley-VCH GmbH; (c) processing and structural illustration of the 2D/3D perovskite films. The films were prepared by one-step spin-coating a precursor solution containing guanidine hydrobromide (GABr) as an inducer to subsidiary generate oriented 2D perovskite phase. Right is the crystal structure, where GABr as the “spacer” occupies interlayer position similar to PEABr. Reproduced with permission from ref. 66, Copyright © 2020 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim; (d) scheme of PEA2(FAPbBr3)n−1PbBr4 phases, which means that the addition of insert-layer molecules can form quasi-2D perovskite phases. Reproduced with permission from ref. 25. Copyright © 2018, The author(s), Springer Nature. |
Converting 3D into quasi-2D perovskite below the Bohr diameter can increase Eb.25,26 Moreover, the quasi-2D perovskites are rising as potential luminescent materials since the cascade energy landscape empower efficient exciton transfer from wide band to narrow band, which then leads to radiative recombination. Recently, Snaith's group53 (Fig. 3(c)) discovered that the incorporation of butylammonium (BA) into the 3D lead halide perovskite can form quasi-2D and 3D grains, accompanied by the enhancement of PL intensity. Fig. 3(d) shows another example of the formation of quasi-2D perovskites by introducing phenylethylammonium bromide (PEABr) in the precursor, which results in an increased PLQY of 57.3%.25
Despite small grain size can spatially confine the excitons for radiative recombination, the amplified carrier density can also lead to enhanced Auger recombination.20 In addition, high Eb would speed up the Auger process due to the enhanced Coulomb electron–hole interaction.20,93 Such interaction can cause an uneven distribution of carriers in space, thus increases the probability of energy transfer to a third carrier and accelerates the Auger process.94 Therefore, Eb should be judiciously controlled in order to maximize radiative recombination while suppress the Auger recombination loss.95 For example, in quasi-2D perovskites of PEA2(FA0.7Cs0.3)n−1PbnBr3n+1 (n = 2, 3,…,∞), Yuan's group found that the low n value perovskite with high Eb cannot reserve the high PLQY and achieve a high brightness under high current density due to the Auger recombination, while high n value with lower Eb perovskite exhibits lower PLQY roll-off because of reduced Auger recombination rate. As a result, the devices-based n = 3 achieved the highest EQE under low current density. The EQE of LEDs declined with the increase of n values (n > 3), and the n = 10 device showed a peak luminance of 91650 cd m−2.95 As the Auger recombination rate is proportional to the materials’ exciton binding energy, another work in the same group have found that introducing high-polar organic cation, p-fluorophenethylammonium (p-FPEA+), into the “A-site” of the quasi-2D perovskites resulted in a reduced Eb. To compensate for PLQY decline induced by the decreased first-order exciton recombination rate, authors have explored a molecular passivation agent, CF3KO3S, to reduce trap-assisted recombination rate. The modified devices exhibited a peak EQE of 20.36% with a record luminance of 82480 cd m−2 thanks to the suppressed Auger recombination at high current density.20
Fig. 4 (a) Schematic illustrated the rearrangement of phase distribution by adding Na+ in quasi-2D perovskites. Reproduced with permission from ref. 97, Copyright © 2020 American Chemical Society; (b) schematic illustration of the addition 4-(2-aminoethyl)benzoic acid (ABA) into perovskite can enhance the interaction between neighbouring perovskite layers. Reproduced with permission from ref. 101, Copyright © 2020 Wiley-VCH GmbH; (c) schematic illustration of the preparation process of adding ethoxylated trimethylolpropane triacrylate (ETPTA) into antisolvent to passivate the defects. Reproduced with permission from ref. 7, Copyright © 2021 Wiley-VCH GmbH; (d) schematic depicting the surface halide vacancies of CsPbBr3 perovskite films and passivated by organic molecules of PEABr and inorganic molecules of LiBr. Reproduced with permission from ref. 27. Copyright © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. |
Meanwhile, the addition of interlayer ligand in quasi-2D perovskite should be rationally chosen, preventing the formation of isolated domains of organic phase and perovskite phase. Such phase separation would weaken the dielectric confinement effects. The dielectric confinement arises from the different dielectric constants between the perovskite phase and organic components, which also contribute to the enhancement of PLQY.21,98 The low dielectric constant of the adjacent organic layers provides poor shielding of the electrons and holes in the inorganic layers.99 If the amount of interlayer molecules of quasi-2D perovskite or the organic modified additives is higher than the threshold values, the organic phase may form, which could weaken the dielectric confinement effect. Taking the case of adding of PEABr into the CsPbBr3 as an example, small ratio (40%) of PEABr impedes perovskite crystal growth and leads to smaller crystallites, while the higher PEABr concentrations promote the formation of PEABr organic phase, which weakens the dielectric confinement effects and serves as a non-radiative recombination channel.21
The introduction of certain small molecule additives can not only promote the uniform phase distribution, but also play an important role in defects passivation. A recent work from Sargent's group showed that the incorporation of fluorinated triphenylphosphine oxide additive (TFPPO) plays a bifunctional role in a reduced-dimensional perovskites (RDP). The fluorine atoms in TFPPO could form a hydrogen bond with large cation PEA, limiting their diffusion during RDP film deposition and consequently resulting in a monodispersed quantum well. Besides, phosphine oxide (PO) could also heal the traps by coordinating with unsaturated Pb2+ sites. These RDP thin films with narrowband emission and high PLQY achieved a champion EQE of 25.6%.100 Further comprehensive discussion about the defect passivation by introducing molecular additives will be provided in Section 4.1.4.
Similarly, modulating the transport property of carriers in quasi-2D/3D hybrids structure could boost the recombination efficiency of electrons and holes, which in return enhances the radiative recombination. In other words, creating a facile cascade channel to induce energy transfer from the wide into narrow bandgap in combined quasi-2D and 3D hybrid structures can inhibit free charge diffusion and thus increase the proportion of radiative recombination.66 Equally importantly, creating stable and efficient heterostructures between quasi-2D and 3D domains can consequently inhibit undesired trap-assisted recombination and further boost the PL intensity.53
Additive engineering is one of the most effective approaches to passivate the undesirable defects and suppress the non-radiative recombination within perovskite polycrystalline film. For convenience, most additives are directly dissolved in the precursor of perovskite,22,27,104 which can mitigating both shallow and deep defects in the bulk, at the surface, or on the grain boundaries.10 In addition, in order to reduce the interfacial defects, the additives can be introduced into antisolvent and deposited on the perovskite layers (Fig. 4(c)).7 Also, during the antisolvent process, small organic molecules could penetrate into the perovskite film along the grain boundaries to restrain pinholes, further reducing non-radiative recombination.7 Abundant additives have been utilized to passivate the defects and successfully suppressed the non-radiative recombination centres. For example, the introduction of bifunctional group including amino and carboxylic molecular can remove metallic Pb defect, which is a major non-radiative recombination center.101 By suppressing such trap state, the PLQY was increased from 45.5% to 63.5% and a high-performance PeLEDs with EQE of 10.11% was achieved. The adding of LiBr can create a Br-rich environment to prevent the aggregation of free Cs+ ions on the grain surface, and thus passivate Br vacancies to reduce non-radiative recombination centres (Fig. 4(d)).27 As a result, a peak EQE of 16.2%, as well as a high maximum brightness of 50270 cd m−2, are achieved. The sulfonic group of zwitterion 3-aminopropanesulfonic acid (APS) can simultaneously passivate deep and shallow level defects in perovskites via coordinate and hydrogen bonding, which improve the EQE from 9% to 19.2%.105 In summary, these additives can be classified as organic and inorganic, which can passivate shallow defects and even deep level defects.4,20
Apart from the prevalent method of additive engineering, other methods are also developed to remove defects, which include vapor-assisted crystallization technique,106,107 doping metal ions (Rb),107 cross linked strategy,28 and so forth.
In typical perovskite nanocrystals synthesis, lead halides always serve as both cation and anion precursor. The molar ratio of Pb-to-Cs is typically set as ∼3.76 for getting pure crystal phase distribution with good morphology.109 Although this “two-precursor” method can produce perovskite nanocrystals with PLQY among 60–90%,110 it still hinders the possibility of studying the desired ion stoichiometry to get high PLQY. In order to overcome such restrictions, the “three-precursor” hot injection approach has been developed (Fig. 5(a)),112 in which the source of lead and halide respectively come from PbO and small halide molecular, including ammonium salts,111,112n-bromosuccinimide halide,113 benzoyl halides,114 and phenacyl halide.115 Such method can provide halide-rich environment during the synthesis, which can effectively remove the non-radiative recombination process at halide-vacancies defects. Therefore, control the ratio between different precursor randomly is essential to get high luminescent nanocrystals.
Fig. 5 (a) Schematic representation of the reaction giving near unity quantum yields in CsPbX3 nanocrystals. Digital photographs of CsPbCl3 (blue), CsPbBr3 (green), and CsPbI3 (red) nanocrystals dispersed solution under irradiation (365 nm). Absorption (black line) and corresponding PL spectra (coloured) of CsPbCl3, CsPbBr3, and CsPbI3 nanocrystals obtained in our approach showing near unity quantum yields. Excitation wavelength was 350 nm. Reproduced with permission from ref. 112. Copyright © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim; (b) diagram showing Ni doping in CsPbCl3 nanocrystals can boost the PLQY to near unity. Reproduced with permission from ref. 118. Copyright © 2018 American Chemical Society; (c) the schematic diagram of the synthesis method with liquid nitrogen cooling and the relationship between cooling rate and crystal quality. Reproduced with permission from ref. 131. Copyright © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. |
The as-synthesized nanocrystals were either halide-deficient or uncoordinated Pb(II) in the lattice and on the surface, which are the main recombination centres of non-radiative.116 The utilization of extra metal halide has been offered as a popular way for removing these flaws.117–120 With excess ZnBr2, the PLQY of CsPbBr3 from blue to green emitting are among 80 to 95%.117 Also, CuCl2 can also help the Mn-doped CsPbI3 nanocrystals to improve the PLQY.119 In both cases, the defects are repaired by adding excess halides without incorporation of Zn2+ or Cu2+ into the lattice. In some cases (Fig. 5(b)), the adding of metal halide can incorporate into perovskite to increase short-range order of the lattice, and simultaneously eliminate the halide vacancies by increasing formation energy of defects to boost the luminescence.118,121 Some trivalent ion (such as Er3+ and Fe3+) can even suppress intrinsic deep traps in CsPbCl3 to increase PLQY.122,123 Intriguingly, Auger recombination can be suppressed by introducing different types of metal halide into nanocrystals, such as chloride into iodide.124 Besides metal halide, different kind of chemical species that can strong adhere to the nanocrystals are adapted to improve the luminescent efficiency, including alkylphosphine,125 quaternary alkylammonium,126 crosslinking component.127
Nanocrystals with approximately unity high PLQY are generally manufactured at high temperatures.119,128 However, a phase transition may be noticed very instantaneously, which cannot be terminated even with ice bath cooling.129 Take cubic CsPbCl3 and orthorhombic CsPbBr3 as examples, they would transfer to corresponding non-emitting tetragonal phases during annealing (>180 °C).129,130 Therefore, alkylammonium halides are used to restrict these phase changes and the obtained nanocrystals can endure the annealing for hours.127 Furthermore, after injecting Cs or another precursor, the nucleation and growth processes might occur simultaneously in a matter of seconds (less than 10 s), requiring rapid cooling to maintain the light-emitting phase of the nanocrystals.116,131 Recent research reveals that ultrafast cooling rate (50 K s−1) by liquid nitrogen can effectively lower the reaction thermodynamic energy of the system below the threshold (Fig. 5(c)). Thus, it can rapidly hinder the further growth of nanocrystals and avoids the additional nucleation, achieving uniformity grain size and better crystallinity.131 The CsPbBrxCl3−x nanocrystals synthesized through such method have been achieved the PLQY of 98%.131
Purified the nanocrystal with poor solvent (such as methyl acetate, ethyl acetate, pentanol, or ever acetone) has been identified as a practical method to remove the insulting ligands.134 However, the excessive detachment of ligands leads to insufficient passivation of the surface defects, resulting in poor stability and facilitating non-radiative recombination. What is worse, recent research indicated that the left uncoordinated halide ions on the surface can even evolve into deep defects.135 And the reduced barrier among the nanocrystals promotes the aggregation and regrowth of perovskite.136 Therefore, additional ligands or additives can be introduced in this process or after the first cycle of purification to passivate the defects and further prevent non-radiative recombination.137
The ligands can be classified as L, X, and Z types by the number of electrons (2, 1, and 0, respectively) that the ligand donates to the surface atoms (Fig. 6(a)).138 L-type ligands usually share a lone electron pair with Pb and act as Lewis bases; while Z-type ligands accept a lone electron pair from the halide anions acting as the Lewis acids. Common X-type ligands with one function group can either passivate the halide anion or Pb site. The zwitterionic/bidentate ligands in X type with two functional groups in positive and negative charges can simultaneously bind to two sites. Due to the ionic nature, the ligands used for capping nanocrystals and maintaining/improving the bright luminescent are fall into the X- and L-type categories.137
Fig. 6 (a) Nanocrystal ligand binding motifs according to the covalent bond classification method. Reproduced with permission from ref. 138. Copyright © 2018 American Chemical Society; (b) comparison of the steady-state absorption and PL spectra for the starting colloid and the same colloid subjected to several treatments, which suggests the synergistic passivation of DDAB and PbBr2 can improve and keep the PLQY. Reproduced with permission from ref. 139. Copyright © 2018 American Chemical Society; (c) relationship between PLQY and different chemical potentials of thiocyanate etchant. Reproduced with permission from ref. 146. Copyright © 2019 American Chemical Society. |
Halide vacancies is one of main defects in causing non-radiative recombination, which can be effectively passivated by organic halogens in X-type. The excess halide-rich environment can produce a halide-rich surface layer, which can form a quantum-well-like band alignment since the bandgap of surface is larger than the core nanocrystals, guaranteeing the excitons generation and high-rate radiative recombination.70 The A site vacancies is another non-radiative recombination centres, which normally can be refilled with the ammonium cation and zwitterionic ligands. As these two defects appear simultaneously during the purification process, there is a need in combining the properties of different ligands to get high PLQY. The coexisting of didodecyl dimethyl ammonium bromide (DDAB) and lead bromide in post treatment process can endow CsPbBr3 nanocrystals with high PLQY of 95–98%, even retaining after three to four cycles of washing (Fig. 6(b)).139 In some cases, the defects can induce Auger recombinations140 and treated with ligands can remove the defects and suppress Auger processes effectively.141
Due to intrinsically ionic and highly dynamic nature of perovskite, the ligands are in a highly dynamic equilibrium between free and bound state.142 Ligands can easily detach from the surface and left unbonded lead or halide ions on the surface, causing more vacancies. Even in the fresh sample, the weak affinity of the ligand can lead coalescence of the nanocrystals, and become worse with aging.143 The binding affinity between the ligands and the surface atoms should be strengthened. According to a recent study, defects caused in purification process can be avoided by increasing the coordination ability between the Pb2+ and the thiol group of ligands, resulting in nanocrystals with a PLQY of 98%.144 In high affinity system, the post treatment with octyl-ammonium sulfate can produce a CsPbX3/PbSO4 core/shell heterostructure, which can improve the radiative recombination and exhibit excellent optical properties.145
Etching the surface selectively represents a promising pathway for mitigating the optical defects in perovskite nanocrystals.146,147 By adjusting concentration or strength of the chemical etchant to change the etchant chemical potential, we can systematically study the etching extent varying from only a few lead atoms to nearly all the lead atoms of the nanocrystal (Fig. 6(c)). It is difficult to completely eliminate non-radiative recombination defects to produce near-unity PLQY if the chemical potential is too low. When the chemical potential is too high, the etching reaction becomes nonselective. Only by choosing an appropriate concentration of an etchant with an appropriate lead binding strength can we selectively remove the non-radiative recombination traps on the surface of nanocrystals, and then improve the PLQY reliably.146
The mismatch of energy level between perovskite layer and CTLs could cause the accumulation of charge carriers at the interfaces, which reduces the possibility of radiative recombination within the emitter layer while increases Auger recombination at the interfaces. To optimize the injection of charges into the perovskite layer, the energy barrier for CTLs/emitting layer should be minimized. In addition, the desired CTLs also should possess electron/hole-blocking abilities for HTLs and ETLs, respectively. Limited by the energy levels of commonly employed CTLs, the modification of transfer layers to align perovskite emitter is a prevalent method in PeLEDs.150 As shown in Fig. 7(a), the addition of MXenes can modified the working function of transfer layer and creates more balanced charge carrier transfer in PeLEDs.150 In the typical device structures of ITO/poly(3,4-ethylenedioythiphene):polystyrene sulfonate (PEDOT:PSS)/perovskite/2,2′,2′′-(1,3,5-benzinetriyl)-tris(1-phenyl-1-H-benzimidazole) (TPBi)/LiF/Al, the efficiency is limited by a lower work-function of the widely used PEDOT:PSS (4.9 eV) compared to perovskite emitting material. In order to tackle this problem, poly(styrene sulfonate) sodium salt (PSS-Na) was incorporated in the PEDOT:PSS solution to increase the work function of HTLs and better aligned with conduction band of perovskite. Besides, the upper insulting polymer increase the carrier quenching blocking ability in the interface of HTL/perovskite, leading to a longer lifetime of perovskite film on modified PEDOT:PSS than that on pristine one. Accompanied with the passivating treatment, the green and blue LEDs showed a champion EQE of 22.5% for green perovskite LEDs and 12.1% for blue LEDs, respectively.7,104
Fig. 7 (a) Device energy-level diagram for MXene modified transfer layer in the PeLEDs (left) and charge carrier injection and recombination mechanisms for 0% (middle), and 10% (right) MXenes based devices. ZTC in the diagram represents ZnO–Ti3C2. Reproduced with permission from ref. 150. Copyright © 2020 The Authors. Published by Wiley-VCH GmbH; (b) schematic illustration of the interface defect passivation mechanism of perovskite films via ammonium thiocyanate modification. Reproduced with permission from ref. 155. Copyright © 2021, The Royal Society of Chemistry. |
The capability for conveying charges varying in light of p/n-type semiconductors. Normally, electrons in ETL transfer much faster than that of the holes in HTL, causing non-uniform distributions of electron and hole in the emitter layer both in spatial and time scale.15 This would result in a higher ratio of Auger recombination accompany with a smaller radiative recombination rate, especially at the interface of emitter layer/CTLs.151 When holes and electrons have a significant transfer rating mismatch, they can recombine within the transfer layers, lowering the possibility of radiative recombination in the emitter layer. In order to tackle this issue, many factors should be taken into consideration for designing the devices, including the thickness of the transfer layer,152 improving/decreasing the transfer ability.150 In the architecture of ITO/PEDOT:PSS/poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine] (PTAA)/perovskite/TPBi/LiF/Al, more holes are injected into nanocrystals and overflow into the ETL of TPBi due to different carrier mobility between PTAA and TPBi, resulting in Auger recombination and interface non-radiative recombination. The introduction of the bilayered electron transport layers including a low mobility ETL (TPBi) and a high mobility ETL (2,4,6-tris[3-(diphenylphosphinyl)phenyl]-1,3,5-triazine) (PO-T2T) provides a feasible route to offset the ability difference in charge mobility, leading to a champion EQE of 21.63% coupled with a significant improved device stability.16
Modifying the carrier transfer layers can influence the interface, especially the one adjusted to the emitter layer. Here we use the most popular HTL, PEDOT:PSS, as example to illustrate such effects. Owing to distinguished hole mobility (1.23 × 10−2 cm2 V−1 S−1) and transparency property, PEDOT:PSS has been widely used in PeLEDs.7,55,101 Nevertheless, the morphologies and crystal quality of perovskite films spin-coated on convention PEDOT:PSS layer is uncontrollable, resulting in interface-induced non-radiative defect.153,154 To solve this issue, additives were added into PEDOT:PSS to manipulate crystallinities, morphology, energy level, and carrier transfer properties. For example (Fig. 7(b)), the adding of organic additive ammonium thiocyanate into PEDOT:PSS can decrease the contact angle, indicating the improved hydrophilicity and wetting properties of the precursor solution dropped on the modified PEDOT:PSS substrate. Moreover, through the scanning electron microscopy, polycrystalline perovskite with clusters transformed into uniform and dense cubic perovskite crystals. As a result, the modified PEDOT:PSS significantly improved light intensity and EQE achieved 460% enhancement.155 Moreover, conducting polymer,7,156 small organic molecules,157 inorganic salt158 are also used in modifying the PEDOT:PSS to remove non-radiative recombination centres in interface between hole injecting layers and perovskite layers, and further improve the PL intensity dramatically. Similarly, the methods used to the most common HTL can also be used to other transfer layers.
Apart from the modification of carrier transfer layers, the modification can also be applied in the emitter layers itself. The surface passivating species, such as alkali metal halide additives, not only can influence the growth of the polycrystalline, but also can significantly smooth the film with less pinholes and further reduce non-radiative recombination at the surface.159 Some amino additives can even both bind with the emitter layer and the charge transfer layer at the same time, thus enhanced the exciton recombination.29
Fig. 8 Future opportunities in minimizing non-radiative recombination towards high-efficiency and long-lifetime PeLEDs. |
The field of PeLEDs is still burgeoning and will continue to flourish in the near future. We envisage that future advances in both understanding and mitigation of non-radiative recombination losses in PeLEDs will soon be translated into further improvement of electroluminescence efficiency and operational stability approaching that of commercial devices, which paves the way for the development of next-generation lighting and display technologies.
Footnote |
† These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2022 |