Synthesis and characterizations of I-III-VI semiconductor quantum dot
CGS core QDs were synthesized and assembled into QD solids as HTL in PeLEDs as depicted in Fig. 1. Supplementary Fig. 1a shows the synthetic scheme and material characterizations of chalcopyrite QD prepared in this work. From high-resolution transmission electron microscopy (HR-TEM) images, CGS QD exhibit uniform spherical shape with average size of 4.4 nm (Supplementary Fig. 1b), comparable to the Bohrs exciton radius (~ 4 nm) of chalcopyrite nanocrystals 47. X-ray diffraction (XRD) patterns of CGS QD present dominant (112) planes attributed to chalcopyrite phase. Fast Fourier Transform (FFT) analysis reveal single-crystal character of CGS QDs, consistent with chalcopyrite structure in the [1–10] direction.
Recent reports reveal that doping with silver (Ag) into chalcopyrite prompts uniform nucleation and reduces deep-level lattice defects 48. Moreover, due to orbital overlap between Ag 3d electrons and Cu 4d electrons, the valence band maximum (VBM) level is lowered, which would facilitate reduced hole transport barrier with quasi-2D perovskite. Herein, Ag doping was conducted by incorporating an appropriate amount of silver iodide (AgI) into metallic precursors. The purified ACGS QD exhibits similar spherical shape without affecting size distributions. Energy-dispersive X-ray spectroscopy (STEM-EDX) show ACG (([Ag]+[Cu])/[Ga]) ratio in the final product is 1.22 as the feed ACG ratio in the precursor is 1:4 (Supplementary Fig. 1c & Supplementary Table 1). Phase evolution from AgGaS2 to CuGaS2 by varying the [Ag]/([Ag]+[Cu]) (AAC) ratio was assessed by XRD diffraction (Supplementary Fig. 1d). Significant influences of chemical composition on the bandgap energy (Eg) of ACGS QD solids were reflected in ultra-violet visible (UV-vis) absorption spectrum. Eg increases from 2.17 eV to 2.78 eV as a deficiency in mono-cation ([Cu]/[Ga]) is increased from 1:1 to 1:8 (AAC = 0) ; and increases conspicuously from 2.58 eV to 2.74 eV as AAC increases from 0 to 10% (ACG = 1:4) (Supplementary Fig. 1e-f). Besides, the appearance of excitonic peak indicates mono-dispersion and homogeneous size distribution 49. Improved crystallinity on Ag doping can be concluded from observable yellow luminescence under UV excitation at 365 nm, shown in the inset picture of Supplementary Fig. 1g.
It is notable that the monocation-poor precursor condition (ACG ratio or [Cu]/[Ga] ≤ 1:2) is necessary to maintain quaternary/ternary chalcopyrite phase in the final QD. Otherwise, the final product is pure chalcocite Cu2S as confirmed by both TEM and XRD measurement. (Supplementary Fig. 2) Detailed investigations into the nucleation and growth process of QD affirmed the formation of crystalline Cu2S/Ag2S at early stages of synthesis. (Supplementary Fig. 3) As temperature rises, the long-wavelength absorption disappears accompanied by a blueshift of the excitonic peak, suggesting a phase transition from binary low-gap Cu2S/ Ag2S, to quaternary wide-gap AgCuGaS2, due to Ga in-diffusion and Ag out-diffusion (Supplementary Table. 2).
Functional surface ligand engineering and improved hole transport in QD solids
Prior to use ACGS QD solids as HTL in PeLEDs, we developed hybrid ligand exchange technology to replace the long-bulky ligand (RNH2, RS−) on the surface of QDs, by short, atomistic ligand 50, as illustrated in Fig. 2a.
At the beginning, ZnCl2 was intentionally selected as solution-phase ligand treatment due to similar lattice structure between chalcopyrite and zinc-blende 51. Specifically, it has been widespread reported about passivation effects on uncoordinated Pb2+ sites with Lewis base through the formation of Lewis adducts 21,25,52–54. Conceptually, electronegative Lewis base Cl− as X-type ligands provide covalent bond formation with electrophilic surface sites of ACGS QDs (Cu+, Ga3+) 55. In addition, the incorporation of Lewis acids into mixed-halide quasi-2D perovskite could reduce the detrimental effects of chlorine on deep-level defects, and avoid phase separation 13,56,57. Simultaneously, ZnCl2 as Z-type ligands could bind to electron-rich Lewis basic sites of ACGS QDs through Lewis acids Zn2+ as two-electron acceptors. The effects of QD surface ligand is representatively depicted in Fig. 1b.
ZnCl2 dissolved in ethanol was added into roughly-purified QD solutions to initiate ligand interactions (see Methods for more detail). Before ligand exchange, pristine ACGS QD exhibit distinguished lattice distortions and stacking faults due to steric hindrances, which could be effectively recovered with improved crystallinities characterized by obvious decrement of inter-dot spacing and uniform distribution after ZnCl2 treatment, which would promote stronger electrical coupling among dots (Fig. 2b and Supplementary Fig. 4a-c). The crystallography characterizations by XRD presents no significant variations in the diffraction pattern after ligand exchange, indicating the process is a surface-limited interaction, instead of alloying or doping into the core (Supplementary 4d). To understand microscopic origins of Cl− passivation, we used density functional theory (DFT) to calculate the binding energies (Eb) of amines (RNH2) and thiolates (RS−) and Cl− ion on the CGS surface. DFT calculations unveil that Cl− could bind to surface Cu+ and Ga3+ with strong Lewis acidity, reconstructing ordered cubic structure from distorted crystal surface due to large steric hindrances of aliphatic ligand (Fig. 2c). Furthermore, from density of states (DOS) calculations, the mid-gap trap states are remarkably suppressed and Eb is significantly improved after ZnCl2 treatment (Fig. 2d and Supplementary 5b). X-ray photoelectron spectroscopy (XPS) studies were conducted to characterize the chemical environment changes of elements in ACGS QD on ligand exchange. As expected, Zn 2p and Cl 2p were distinctly identified in ZnCl2-treated ACGS QD (Supplementary Fig. 6). In addition, both Cu 2p 1/2 and Ga 3d emission peaks, as well as corresponded auger emissions (Cu LMM, Ga LMM) slightly-shift to higher binding energy regime, due to bonding with electron-rich Cl− ions. The metal nitride peaks, evidenced from resolved Ga 3d (396.8 eV) and N 1s (25.68 eV) core-level emission, are efficiently eliminated in ACGS QDs after ZnCl2 treatment (Supplementary Table 3). Quantitative analysis from XPS spectra indicate reduced atomic concentration of nitrogen (Supplementary Table. 4). These results clearly indicate effective exchange of surface amines by Cl− ions. Nevertheless, ZnCl2 cannot sufficiently remove residue thiolates ligand from 1-DDT used to lower the reactivity of Cu+, which can be recognized from thermogravimetric analysis (TGA) and fourier transform infrared spectroscopy (FTIR) analysis, in which the methyl and methylene vibrations are still present in ACGS-ZnCl2 QD. (Supplementary Fig. 7a-b). To fully stripe the un-exchanged alkyl ligand on ACGS QD, we explored solid-phase ligand exchange process to produce homogenous QD solids using 1,2-ethanedithiol (EDT) containing double RS− group 58. Likewise, DFT simulations show that EDT exhibits enhanced coordination with surface gallium, leading to increased Eb from 1.049 eV to 9.747 eV, and fully-passivated defect states. High-resolution XPS spectrum suggest trivial influences on chemical bonding and atomic concentrations of elements in ACGS QDs. Meanwhile, the Cu and Ga concentration is declined after ZnCl2 treatment, as summarized Supplementary Table 3. With observations of decrement of QD size and blue-shifted absorption spectrum of ACGS-ZnCl2 QD (Supplementary Fig. 7c), it is reasonable to address that the crystallinity of ACGS QD is mainly improved by chlorine ions via exchange of surface long-bulky ligand and desorption of the surface metallic complexes 59. As ZnCl2 removes ligand excessively, ACGS QDs tend to aggregate in the solution, detected by dynamic light-scattering (DLS) measurement (Supplementary Fig. 7d). The influences of ligand density and film washing with EDT on surface morphology of QD solids were monitored through scanning electron microscopy (SEM), and atomic force microscopy (AFM) (Supplementary Fig. 8). As the ligand to QD ratio exceeds 100:1, the deposited films become rough and fuzzy, which is detrimental for light out-coupling. The optimized ligand to QD ratio, is limited to be 100:1 for ZnCl2 treatment (see Methods). Consequent solid-phase exchange is optimized to be single coating with EDT (1 vol% in acetonitrile). The as-synthesized and exchanged ACGS QDs by ZnCl2 and EDT is named as ACGS-pristine, ACGS-ZnCl2, ACGS-ZnCl2-EDT, respectively.
Next, we built up a thin-film transistor model with bottom-gate-top-contact structure using QD solids as transport channel (QD-TFT) to quantitatively assess the carrier mobility and concentration. Figure 2e depicts the device structure and the picture of SiO2 substrate as well as SiO2/ACGS QDs under optical microscope is shown in Supplementary Fig. 9a. By altering the concentration of QD dispersions, the thickness of QD solids can be practically controlled (Supplementary Fig. 9b).
Hole-dominant transport can be surly confirmed from transfer characteristics (IDS-VGS) in QD-TFT devices (Fig. 2f). IDS collected from QD-TFT based on ACGS-pristine is significantly increased from 0.1 nA to 20 nA after ZnCl2 treatment, and further elevated to hundreds of nano-ampere (nA) after EDT treatment. Corresponding output characteristics (IDS-VDS) of QD-TFT based on ACGS-ZnCl2-EDT QD solids are depicted in Fig. 2g, signifying saturation region while VDS exceeds − 2 V. The influences of chemical composition on transfer characteristics of QD-TFT based on ACGS QD are reflected in Supplementary Fig. 10a-c. In QD-TFT based on CGS QD, IDS increases as [Cu]/[Ga] ratio increases. This can be explained as the formation of large amounts of nano-crystalline CuxS phase in CGS QD as [Cu]/[Ga] ratio increases. Contrarily, for ACGS QD, IDS decreases as AAC ratio increases, which can be explained as reduced shallow, acceptor-like VCu (copper vacancy) defects therefore decreasing the hole current.
The corresponded hole mobility (µh) in the linear working region was calculated based on the transfer characteristics (see Methods). For comparison, the transfer characteristics of NiOx, TFB, and PVK were also measured and summarized (Supplementary Fig. 10d & Fig. 2h). The calculated µh in ACGS (AAC = 10%), and CGS QD solids could reach 0.191 cm2 V− 1s1 and 0.546 cm2 V− 1s1, respectively, compared to that of 10− 5 for organic HTL and 0.127 cm2V− 1s1 for NiOx thin-films. With µh calculated, the hole concentration (ph) was predicted through current-voltage (J-V) characteristics of the two-terminal device structure: glass/ITO/ACGS QD/MoO3/Ag (Supplementary Fig. 10e). Supplementary Table 5 summarizes µh and ph for chalcopyrite QD solids after hybrid ligand exchange prepared in this work. ACGS QDs exhibits nominal hole mobility but relatively lower hole concentration than organic HTL.
Apart from quantitative analysis of carrier mobility and concentration, the band structure of ACGS QD solids was obtained through ultra-violet spectroscopy (UPS) and kelvin probe force microscopy (KPFM). The effects of chemical stoichiometry, surface chemistry, dimension of QD on work function (WF) and valence band maximum (VBM) were shown in Supplementary Fig. 11a-c. Firstly, Ag doping lowers the VBM energy level due to reduced hole concentration. Specifically, as ACG ratio is 1:2 and the AAC ratio increases from 0 to 10%, the VBM deepens from − 4.53 eV to -4.93 eV. Secondly, increased Cu deficiency leads to remarkably-deepened Fermi-level position (EF) and the VBM. For instance, as AAC ratio is 0, and the [Cu]/[Ga] ratio decreases from 1:2 to 1:8, the VBM is lowered from − 4.93 eV to -5.37 eV. Then, the quantum confinement effects also affect the band structure; when we synthesized larger ACGS QDs (average size ~ 5.37 nm) using chloride precursors (CuCl, GaCl3) remaining other parameters unchanged (ACG = 1:2, AAC = 10%), the VBM is upshifted to -4.50 eV compared to -4.93 eV for smaller QDs (Supplementary Fig. 11d-e). Lastly, the surface ligand treatment also elevates VBM as compared to ACGS-pristine. Thus, it is notable that the band structure of chalcopyrite QD solids could be feasibly manipulated through composition engineering, ligand engineering and quantum confinement effects (Supplementary Fig. 11f). Next, the microscopic WF level of quasi-2D perovskite films as well as various HTL were estimated through surface potential mapping measured with kelvin probe microcopy (KPFM) (see Methods). It is calculated that EF of ACGS QD lies between PEDOT:PSS and perovskite, signifying its critical role as an energy cascade to assist hole injection from PEDOT:PSS to perovskite (Supplementary Fig. 12).
Buried interface passivation and regulated crystallizations of quasi-2D perovskite
After thorough investigations of ACGS QDs and their electrical properties, the influences of ACGS QDs on the interfacial passivation effects on recombination lifetime and crystallizations of quasi-2D perovskite were investigated through time-resolved photoluminescence (TRPL) and XPS. Excitons trapped by defect states attributed to uncoordinated Pb2+ are subject to non-radiative recombination 25,60.
Herein, evidence of chemical coordination between ACGS QDs and Pb2+ were assessed by high-resolution XPS on Pb 4f, Br 3d and Cl 2p photoemissions (Fig. 3a-c). The perovskite deposited on PEDOT:PSS substrate is indicated as “control” sample, compared to that deposited on ACGS-ZnCl2-EDT QD solids. From Fig. 3a, the core-level peaks of Pb 4f shift to higher binding energies, due to increased electron-affinity of Cl− and RS− possessing odd number of valence-shell electrons 60. Similarly, the photoemission peaks of Cl 2p 1/2 (199.48 eV) and Cl 2p 3/2 (197.88 eV), as well as Br 3d 1/2 (69.26 eV) and Br 3d 1/2 (68.33 eV) both shift to higher binding energies by 0.3 eV, due to charge disturbance from electron donations of Cl− and RS−, and chemical bonding with Zn2+ which subsequently results in the change of static interactions with Pb2+ cations. In addition, the Zn2+ cations attached to the surface of ACGS QDs were detected to diffuse into the perovskite during the annealing process (Supplementary Fig. 13). Therefore, it is concluded that the incorporation ACGS QD solids facilitate in eliminating uncoordinated Pb2, leading to the formation of “perovskite crystal seeds” at buried HTL/perovskite interface, favoring ordered and homogenous crystallizations of quasi-2D perovskite 61–63. Furthermore, the stabilization of halide interstitials via Zn2+ incoporation could alleviate vacancy defect formation which tends to provide major ionic migration pathways in mixed-halide perovskite 32, 64,65. Similar observations have been reported for alkaline salts, Rb+ and K+ to suppress the detrimental effects of chlorine in photo-induced phase separation 13,66. Due to suppressed interfacial exciton quenching and better defect passivation, the overall steady-state photoluminescence (PL) intensity of quasi-2D perovskite films crystalized on ACGS-ZnCl2-EDT QD solids is significantly improved compared to those on ACGS-pristine, as well as PEDOT:PSS substrate (Fig. 3d).
The charge carrier dynamics at the perovskite/HTL hetero-junction sites and within the quasi-2D perovskite were examined by time-resolved PL (TRPL) analysis, as shown in Fig. 3e. We extracted an average recombination lifetime of 6.46 ns, 1.26 ns an 1.76 ns for perovskite deposited on PEDOT:PSS (control), PEDOT:PSS/ACGS-pristine and PEDOT:PSS/ACGS-ZnCl2-EDT QD solids, respectively. The extracted PL decay lifetimes (τ1, τ2) and their proportionality coefficients (A1, A2) are presented in Supplementary Table 6 using second-order exponential functions (see Methods). In typical quasi-2D perovskite, the fast process (τ1) is usually considered to be the lifetime of excitonic emission 67. The proportionality coefficient of fast-decay component (A1), increases to 0.81 for perovskite deposited on ACGS-ZnCl2-EDT substrate, compared to 0.50 and 0.75 for that deposited on PEDOT:PSS, and ACGS-pristine substrate, respectively, indicating strengthened radiative recombination in perovskite. In addition,τ2 fitted for perovskite deposited on ACGS-ZnCl2-EDT QD solids substrate (3.98 ns) is strongly improved than that on ACGS-pristine QD (1.88 ns), reflect suppressed non-radiative recombination due to the formation of traps within the perovskite. We also performed TRPL analysis on quartz glass substrate to exclude the effects of charge transport (Supplementary Fig. 14 & Supplementary Table 7). As expected, the overall lifetime is shortened from 9.19 ns (quartz glass) to 1.54 ns (ACGS-pristine) and 4.13 ns (ACGS-ZnCl2-EDT), due to effective charge extraction of ACGS QD solids. Enhanced proportionality of fast-decay component (A1), as well elevated slow-decay lifetime (τ2) for perovskite deposited on ACGS-ZnCl2-EDT QD solids substrate, suggests accelerated excitonic radiative recombination, and suppressed non-radiative recombination due to regulated crystallization kinetics of quasi-2D perovskites on ACGS-ZnCl2-EDT QD solids. As hydrophilicity of ACGS-ZnCl2-EDT QD solids is also increased compared to PEDOT:PSS from contact angle measurement (Supplementary Fig. 15), leading to homogeneous film coverage and uniformly-accelerated nucleation process thus significantly-enlarged crystal dimension of quasi-2D perovskite (Fig. 3f ) 68. The reduced grain boundaries also contributed to suppressed ionic migration and stable perovskite structure.
UV-vis spectra showed distinct absorption peaks attributed to the small-n (n = 1, 2, 3) phases (Supplementary Fig. 16a). XRD diffraction pattern shows the diffraction peaks at 15.6o and 31.3o corresponding to (100) and (200) plane of three-dimensional CsPbBr3 perovskites on both PEDOT:PSS and ACGS QD substrate. While, diffraction peks at 12.8o and 27.8o indicating quasi-2D phases (Supplementary Fig. 16b). The negligible influences of ACGS QD solids on phase distribution of quasi-2D perovskite substantiate that the improved photoluminescence and exciton recombination primarily come from ordered crystallizations and passivated interface with charge transport layer.
Defect physics and highly-efficient, stable blue Pe-LED
To corroborate advantageous effects of ACGS QD solids on carrier transport and interfacial passivation in PeLEDs based on quasi-2D perovskite, blue PeLEDs devices were prepared with the structure: glass/ indium tin oxide (ITO)/PEDOT:PSS/ACGS QD/perovskite/1,3,5-tris(1-phenyl-1H-benzimidazol-2-yl)benzene(TPBi)/LiF/Al (device A), glass/ ITO/PEDOT:PSS/PVK/perovskite/TPBi/LiF/Al (device B) and glass/ ITO/ACGS QD/PVK/perovskite/TPBi/LiF/Al (device C) in which ACGS QD is short for ACGS-ZnCl2-EDT QD solids. Motivated from recent success of “self-assembled monolayer (SAM)” as HTL in highly-efficient perovskite photovoltaics 69, we sought to incorporate a thin-layer of ACGS QD solids on top of PEDOT:PSS. This interfacial layer is supposed to play the role of selective charge transport agent; also be able to suppress the formation of bulk defects through surface modifications.
The device architecture (device A) was optimized based on the band landscape as schematically drawn in Fig. 4a, and the corresponding energy band level of quasi-2D perovskite were obtained from UPS measurement (Fig. 4b & Supplementary Fig. 17). Figure 4c-d & Supplementary Table 9 demonstrate electroluminescent performances of device A, B and C. From Fig. 4c, device C exhibit shunting diode and low EQE. Compared to device B, device A preserves rapidly-increased current density after reaching turn-on voltage (3.3 V), and suppressed efficiency roll-off due to increased hole conductivities of ACGS QD solids, preventing energy loss in HTL and enhancing hole injection into the perovskite emissive layer, thus leading to a maximum EQE and luminance of 10.85%, and 1450 cd/m2, respectively. Figure 4e shows the EL spectrum of our champion blue PeLEDs based on ACGS QDs. The EL emission peak at 471 nm with narrow emission bandwidth of 21 nm matches well PL spectrum (Fig. 3d), indicating balanced charge transport dynamics. The spectral stabilities of EL emission strongly support the critical advantages in preventing exciplex emission and stability enhancement of ACGS QD solids. A histogram of the maximum EQE values for 20 devices (Fig. 4f) shows an average EQE of 9.71%, compared to that of 4.04% for device B with PVK as HTL. Figure 4g-h display the device performances of device A with varied AAC ratio of ACGS QD since the applied AAC ratio during QD synthesis is of vital importance in manipulating hole mobility and hole transport barrier of QD solids. To achieve the best EQE, an optimized AAC ratio between 15% and 20% could be verified, above which dramatically-dropped hole mobility is responsible for current and EQE losses for PeLEDs. To highlight the stability enhancement of PeLEDs due to ACGS QDs, we also prepared NiOx-based PeLEDs (device D), with the structure of glass/ indium tin oxide (ITO)/NiOx/PVK/perovskite/TPBi/LiF/Al (Supplementary Fig. 18). The measured operational lifetime when the luminance dropped to half of the initial value at 100 cd/m2 (T50@100nit) of device A, is significantly extended to 78 min, which is more than four times longer than device D (18 min) (Fig. 4i).The enhanced stabilities indicate optimal charge management and effective blocking effects of impurities from ITO and PEDOT:PSS substrate. The problematic operating lifetime for NiOx has been reported due to weak attachment to PVK and disordered molecular interactions, which result in significant interfacial defects and exciton quenching 68. The overall device performances achieved herein were summarized in supplementary Table 8, and compared to other reported highly-efficient pure blue PeLEDs reported in the literature, manifesting as one of the world top-class devices (Supplementary Table 9).
Despite of comprehensive performance enhancement in blue PeLEDs on introducing ACGS QD solids as HTL, we aim to quantitatively validate their universal effects on hole transport efficiency and defect passivation of quasi-2D perovskite, from device-level electrical diode analysis. Thermal admittance spectroscopy (TAS) and capacitance-profiling (C-V) were conducted on LED devices and hole-only devices (HOD) based on the architecture of device A and B. From raw capacitance-frequency (C-f) characteristics provided in Supplementary Fig. 19, The activation energy (EA) of deep-level defect could be extracted, which is 0.41 eV (device A) and 0.60 eV (device B), respectively (Fig. 5a). Moreover, the calculated density of states for these deep-level traps is 5.42×1016 cm− 3·eV− 1 (device A), and 6.16×1017 cm− 3·eV− 1 (device B), respectively (Fig. 5b). Supplementary Fig. S20 plot the DOS distribution, from which energetically-narrow defect distribution and reduced maximum defect density revealed for device A suggests substantially-weakened trap-assisted non-radiative recombination. The detailed calculations on EA and DOS are provided in Methods.
From C-V profiling of HOD in Fig. 5c and Mott-Schottky (MS) analysis in Supplementary Fig. S21a, the built-in potential (Vbi) of HOD based on ACGS QD solids is 0.21 V, profoundly-lower than that based on PVK (0.75 V), due to higher WF level of ACGS QDs resulting in reduced Fermi-level splitting with perovskite (Fig. 5d). Thus, ACGS QD facilitates in overall reduced band-bending and mitigated hole transport barrier. Consistently, the capacitance value of HOD based on ACGS QD (device A) rises dramatically to 3.4 nF compared to that with PVK (device B), which stay almost constant around 1.75 nF until the bias of 3 V, featuring effective charge injection from ACGS QD solids to perovskite. Similar results are also determined in C-V measurement for PeLED devices, certifying lower Vbi for device A (1.8 V) compared to device B (3.0 V). (Supplementary Fig. S21b-e) Consequently the hole current collected from HOD based on ACGS QD (device A) is ten-fold higher than that based on PVK (device B) (Supplementary Fig. S21f) It is thus ascertain that ACGS QD solids HTL provide effective passivation of uncoordinated Pb2+ and stabilize halide interstitial defect in quasi-2D perovskite. As a result, the interfacial nucleation kinetics is regulated through microscopic defect control which enhance interfacial charge transport and prevent interfacial recombination.