Yuanbin Cheng, Qian Li, Mengyuan Chen, Fei Chen, Zhenghui Wu, Huaibin Shen. High-brightness green InP-based QLEDs enabled by in-situ passivating core surface with zinc myristate[J]. Materials Futures, 2024, 3(2): 025201. DOI: 10.1088/2752-5724/ad3a83
Citation: Yuanbin Cheng, Qian Li, Mengyuan Chen, Fei Chen, Zhenghui Wu, Huaibin Shen. High-brightness green InP-based QLEDs enabled by in-situ passivating core surface with zinc myristate[J]. Materials Futures, 2024, 3(2): 025201. DOI: 10.1088/2752-5724/ad3a83
Paper •
OPEN ACCESS

High-brightness green InP-based QLEDs enabled by in-situ passivating core surface with zinc myristate

© 2024 The Author(s). Published by IOP Publishing Ltd on behalf of the Songshan Lake Materials Laboratory
Materials Futures, Volume 3, Number 2
  • Received Date: February 03, 2024
  • Revised Date: March 24, 2024
  • Accepted Date: April 02, 2024
  • Available Online: April 06, 2024
  • Published Date: April 22, 2024
  • AbstractThe performance of red InP and blue ZnTeSe-based quantum dots (QDs) and corresponding QD light emitting diodes (QLEDs) has already been improved significantly, whose external quantum efficiencies (EQEs) and luminances have exceeded 20% and 80 000 cd m−2, respectively. However, the inferior performance of the green InP-based device hinders the commercialization of full-color Cd-free QLED technology. The ease of oxidation of the highly reactive InP cores leads to high non-radiative recombination and poor photoluminescence quantum yield (PL QY) of the InP-based core/shell QDs, limiting the performance of the relevant QLEDs. Here, we proposed a fluoride-free synthesis strategy to in-situ passivate the InP cores, in which zinc myristate reacted with phosphine dangling bonds to form Zn-P protective layer and protect InP cores from the water and oxygen in the environment. The resultant InP/ZnSe/ZnS core/shell QDs demonstrated a high PL QY of 91%. The corresponding green-emitting electroluminescence devices exhibited a maximum EQE of 12.74%, along with a luminance of over 175 000 cd m−2 and a long T50@100 cd m−2 lifetime of over 20 000 h.
  • Due to cadmium’s toxicity and carcinogenic effects, InP based quantum dots (QDs) are the material most favored for new QD displays [1-10]. By changing size, their spectra can cover the entire visible region. However, compared with Cd-based QDs, InP-based QDs tend to have lower photoluminescence quantum yields (PL QYs) and broader full-width at half-maximum (FWHM). In particular, bare InP core has extremely low PL QY (<1%) due to the ease of oxidation of their highly reactive surface even in insert reaction chambers. It has been reported that the defects from oxidative species are probably the reasons for the non-radiative recombination and poor PL QY of InP-based core/shell QDs [11-14]. Therefore, eliminating the surface defects to improve the optoelectronic properties of InP-based QDs is critical to incentivizing the transition to less toxic QD materials.

    Over the past decade, extensive studies have been applied to improve the optoelectronic properties of InP-based QDs for display and lighting applications [9, 10, 14-20]. The most common method is growing shells composed of wide-bandgap semiconductors (ZnS, ZnSe, ZnSeS) on InP cores [21-24]. The shell composition and structure have been carefully optimized to reduce defects and improve radiative recombination. However, the surface defects still exist even after the shell growth. Therefore, eliminating surface defect states of the InP core before and during the shell coating is necessary to enhance the optoelectronic characteristics of core/shell QDs. So far, the most widespread and effective method for synthesizing InP-based QD with PL QY as high as 90% for consumer displays is the addition of hydrofluoric acid (HF) before shell growth [25-28]. The introduction of fluoride ions not only can eliminate surface oxides, but also passivate the dangling bonds around phosphorus atoms and fill the vacancies on the InP core. Nevertheless, HF has severe toxicity and corrosivity, which makes it unsuitable for large-scale synthesis. Hence, Li et al reported red InP/ZnSe/ZnS core/shell QDs with a near-unity PL QY by treated with inorganic salt ZnF2. The corresponding QD light emitting diodes (QLEDs) achieved a maximum external quantum efficiency (EQE) of 22.2% and a peak luminance of over 110 000 cd m−2. Particularly, the device showed a T95 lifetime of over 32 000 h. This strategy has shown the highest efficiency in electroluminescence devices to date [14]. However, the generation of HF at high temperature can rapidly increase the pressure in the container, posing a concealed risk. Therefore, a safe and easy synthesis strategy is necessary for growing high-quality InP-based QDs.

    Here, we demonstrate a fluoride-free synthesis strategy in which zinc myristate reacts with phosphine dangling bond to form Zn-P protective layer and protect InP core from environmental influences. Eventually, it enhances the optoelectronic properties of the cores and the corresponding core/shell structures. X-ray photoelectron spectroscopy (XPS) tests suggested the efficient removal of defects. The resultant InP/ZnSe/ZnS core/shell QDs exhibit a PL peak of 534 nm, an average size of 8.1 nm and a high PL QY of 91%. The corresponding green-emitting electroluminescence device demonstrates a high EQE of 12.74% and a peak current efficiency of 53.31 cd A−1. Especially, the device obtains a luminance of over 175 000 cd m−2 and a long T50@100 cd m−2 (which characterizes how long it takes for the 100 cd m−2 luminance to decrease to its half) lifetime of over 20 000 h. As we know, the luminance is the state-of-the-art value among those for the reported green InP-based QLEDs.

    To synthesize the green InP cores, 0.2 mmol In(Ac)3, 0.1 mmol Zn(Ac)2, 0.8 mmol MA, and 10 mL of ODE were mixed and placed in a 50 mL flask, and then degassed at 150 °C for 30 min. Then, phosphorus precursors were promptly injected into the above mixture under nitrogen atmosphere, and the mixture was heated to 270 °C to grow InP cores for 3 min. In turn, 1 mL ZnMy2 was added at this temperature and reacted for 5 min. Then, 1 mL Se-ODE was injected and kept for 15 min at 270 °C. After completion, the reaction dropped to room temperature. The cooled solution was purified one time using hexane and ethanol, and then redispersed into hexane for shell coating.

    The InP/ZnSethin cores, 2 mmol Zn(St)2, 0.55 mmol Zn(Ac)2 and 10 mL ODE were mixed and placed in a 50 mL flask, followed by degassing at 150 °C for 10 min, and then 1 mL Se-ODE was added with increasing reaction temperature to 200 °C. For the ZnSe shell growth, the 0.5 mL Se-ODE was added every 10 min, and the mixture was heated to 310 °C at each 10 °C. The Se-ODE maintained a total addition of 10 mL. Then, we cooled the mixture down to room temperature. Following, for further ZnSe shell growth, 2 mmol Zn(St)2 and 0.55 mmol Zn(Ac)2 were added to the mixture, and then raised the reaction temperature to 310 °C. Furthermore, the 0.5 mL Se-ODE was added every 10 min until the total additional amount was 15 mL. Finally, the ZnS shell was grown by adding 0.5 mL S-ODE-TOP every 10 min three times. The cooled solution was purified by hexane and ethanol three times, and then redispersed into n-octane.

    Mg(OAc)2 · 4H2O and Zn(OAc)2 · 2H2O with a ratio of 1:7 by mass were dissolved in DMSO to form 15 mL solution at a concentration of 0.50 mol L−1. TMAH was dissolved in ethanol to form 5 mL solution with a concentration of 0.55 mol L−1. Then, the above two solutions were mixed and then stirred for 1 h. Finally, the solution was purified by n-hexane and redispersed in ethanol. The concentration of ZnMgO NC was maintained at 30 mg mL−1.

    In the synthesis, myristic acid, indium acetate and ODE were mixed in a 100 mL flask. Then, the (TMS)3P precursors were injected quickly at 150 °C under nitrogen atmosphere and maintained for 3 min to promote the InP core growth. However, the extreme reactivity of (TMS)3P would lead to the fast depletion of (TMS)3P in solution, which seriously affected the subsequent growth process of InP core. Herein, we added ZnMy2 at 270 °C and adjusted its concentration (figure S1, supporting information), which could adjust the nucleation and eliminate surface oxidation. Choi et al reported that the introduction of ZnMy2 might produce Zn-P complex, which acted as phosphine reservoir to improve the monodisperse in the growth process [29]. As shown in figure 1(a), compared with the InP cores without ZnMy2, the InP cores with ZnMy2 have an increased absorption valley/peak ratio (Vd) of 0.5 and a narrower FWHM of 35 nm, indicating their more uniform size distribution. What is more, the InP core with ZnMy2 exhibits brighter green light emission even without shell formation (figure 1(b)). Correspondingly, the PL QY increases from 13% for the InP core without ZnMy2 to 36% for the InP core with ZnMy2. From the inset of figure 1(b), we can see the tailing emission in the InP core without ZnMy2, which may be attributed to their surface defect states caused by oxidation. To verify this hypothesis, XPS spectra were also analyzed (figure 1(c)). In the XPS spectra for P2p, the peak near 133.5 eV is attributed to the oxidized phosphorus due to the formation of P-O bond. The area ratio of P-O decreases from 33% for the InP core without ZnMy2 to 14% for that with ZnMy2, indicating that ZnMy2 can react with the surface of InP core and inhibit the development of the InPOx surface oxide layer. In order to further analyze the interaction between ZnMy2 and the surface of InP core, high-resolution XPS signals of In3d and Zn2p spectra were tested. As shown in figure 1(d), due to the weaker electronegativity of P than that of Se, the peak of Zn2p shifts to lower energy by 0.4-0.6 eV when the Zn-Se bond is partially replaced by Zn-P complex. 1H NMR spectra (figure S2 in supporting information) also supported that Zn atoms from ZnMy2 were transferred to the vicinity of phosphine atoms and formed Zn-P complex. The signal around 2.20-2.25 ppm in 1H NMR spectra corresponds to the -CH2- around the carboxy group in myristic acid. The 1H NMR spectra of ZnMy2 did not show any peaks around 2.20-2.25 ppm, since there was no carboxy group. However, the peaks around 2.20-2.25 ppm appear again after mixing ZnMy2 and phosphine precursor, which indicates that Zn atoms were removed from ZnMy2 and a small amount of myristic acid was obtained. The free Zn2+ ions combined with phosphine atoms to form Zn-P complex. In addition, the 1H NMR signal of phosphine in 0.89-0.93 ppm shifted to the left after it mixed with ZnMy2 and formed Zn-P complex. This meant that the valence state of H atoms became less positive and that of phosphine atoms became less negative in (TMS)3P, leading to a less chemical reactivity of phosphine precursor. The formation of Zn-P complex will protect the InP core from the influence of water and oxygen in the environment, while it improves the monodisperse through the phosphine reservoir on the other hand. The peak of In3d shifted to lower energy by 0.3 eV, which further verified the reduction of InPOx.

    Figure  1.  (a) UV-vis absorption and (b) PL spectra of InP cores without and with ZnMy2. High-resolution XPS signals of (c) P spectra and (d) In and Zn spectra for InP cores without and with ZnMy2. The inset in figure 1(b) shows photographs of the InP cores without (bottom) and with ZnMy2 (top) solution under UV irradiation.

    What is more, the quasi-ZnSe shell was introduced to avoid the InP core Ostwald ripening at high temperature [30]. As can be seen from figure S3 in supporting information, the exciton peak position of the InP core with ZnMy2 in absorption spectra is red-shifted by 30 nm, indicating that the size of InP/ZnSethin with ZnMy2 is larger. The corresponding TEM images (figure S4, supporting information) also verified this conclusion, and the average size of the cores is 2.8 nm (with ZnMy2) and 2.0 nm (without ZnMy2), respectively. The above results show that the ZnMy2 is degraded to form Zn2+ and myristic acid after mixing. Zn2+ combine with phosphine to form Zn-P complex, while myristic acid enters the solvent. When Se-ODE is added, Zn in Zn-P complex can act as a Zn source, forming a thin ZnSe shell layer. Furthermore, the Zn-P complex can also act as phosphine reservoir to control the monodisperse growth process, and act as a protective layer to avoid the exposure of InP core to water and oxygen. Therefore, the more uniform size distribution and better surface passivation of InP cores will bring about improved PL QY, which are also beneficial for subsequent shell growth.

    Subsequently, we coated the ZnMy2-treated InP core with ZnSe/ZnS shell and studied the effect of ZnMy2 addition on shell growth. Figure 2(a) displays the route for synthesizing InP/ZnSe core/shell QDs and InP/ZnSe/ZnS core/shell/shell QDs. The evolution of spectra and morphologies are shown in figures 2(b)-(d), respectively. Both samples show red shifts of the first excitonic absorption and the PL peaks after ZnSe coating. This is because the offset between the conduction bands of InP core and ZnSe inner shell is small. To further grow the ZnS outer shell, electrons are well confined because there is a larger band offset between ZnS and InP, limiting the shift of emission peaks. Finally, InP/ZnSe/ZnS QDs were synthesized with a similar PL peak located at 534 nm. As verified by TEM in figure 2(d), the average size for the as-prepared core/shell/shell QD with ZnMy2 is estimated to be 8.1 ± 1.0 nm with better size distribution, while the average size for the QD without ZnMy2 is smaller (7.8 ± 1.2 nm) (figures S5(a) and (b), supporting information). The better size distribution might be attributed to the more uniform effective growth of the shell layer due to the minor defects in the InP/ZnSethin cores. QDs with ZnMy2 show high crystallinity as suggested by the high-resolution transmission electron microscopy image. The (111) plane shows a lattice spacing of ∼0.31 nm (inset), confirming that the ZnS outer shell is epitaxially grown successfully. This result is consistent with the x-ray diffraction (XRD) patterns, as shown in figure 2(e). The three peaks at 26.6°, 44.2°, and 52.2° in the XRD patterns correspond to (111), (220), and (311) planes, respectively, which indicates that the InP core is a zinc-blende crystal structure. Continue to grow the ZnSe inner shell and ZnS outer shells, the three peaks are moved to higher angles and gradually become narrower, indicating the successful growth of ZnSe and ZnS shells. What’s more, the addition of ZnMy2 has almost no effect on the lattice structure of QD (figure S5(c), supporting information). By virtue of the passivation of the core, the core/shell QDs with ZnMy2 demonstrate an improved optical property with a high PLQY of 91% and an FWHM of 44 nm. In contrast, the relatively low PL QY (76%) and broadened emission (48 nm) of QDs without ZnMy2 can be ascribed to the imperfect passivation of defect states in the InP/ZnSethin core. The PL decays of both samples are shown in figure 2(f). The time-resolved PL decays is fitted with a double exponential, where A1 (A2) and τ1 (τ2) are the fraction ratios and lifetimes of the exciton (trapped exciton) emissions [31, 32]. The exciton emissions not only have increased PL lifetime (form 23.26 ns to 25.05 ns) but also have increased fraction ratios (from 48% to 55%), while the PL lifetime of trapped exciton emissions decreased from 61.26 ns to 58.98 ns with the introduction of ZnMy2. The average PL lifetime of QDs solution without ZnMy2 is 43.02 ns, while the lifetime of ZnMy2-treated InP QDs decreases to 40.32 ns. The above result suggests that the treatment of ZnMy2 not only can reduce the exciton quenching, but also can enhance the radiative recombination rate.

    Figure  2.  (a) Synthetic route of InP/ZnSe/ZnS QDs. Absorption and PL spectra of InP/ZnSe/ZnS QDs (b) without and (c) with ZnMy2. The photographs of both the QDs solutions under UV irradiation are shown in the insets. (d) TEM pictures of InP/ZnSe/ZnS QDs without (top) and with ZnMy2 (bottom). (e) XRD patterns of InP/ZnSethin, InP/ZnSe and InP/ZnSe/ZnS QDs. (f) Time-resolved PL decays of InP/ZnSe/ZnS QDs without and with ZnMy2.

    We next fabricated QLEDs with the benchmark structure of ITO/PEDOT:PSS/PF8Cz/QDs/ZMO/Al using InP/ZnSe/ZnS QDs as the emitting layer (figure 3(a)). They exhibit a bright green color, and their Commission Internationale de l’Eclairage chromaticity coordinates are (0.304, 0.675) (figure S6, supporting information). The fabrication process of QLEDs is detailed in the supporting information. The energy levels of the charge transport layer were acquired from literature [33-35], and the energy level of QDs was calculated through a combination of ultraviolet photoelectron spectroscopy and optical measurements. As can be seen from figure S7 in supporting information, according to the formula of work function = 21.22-cutoff edge of secondary electrons, the valence bands of QDs without and with ZnMy2 are 5.33 eV and 5.16 eV, respectively. Taking the similar bandgap of the two QDs (2.28 eV) into account, their conduction bands are 3.05 eV and 2.88 eV, respectively. The upward shift of the valence band and conduction band not only facilitates hole injection but also suppresses the injection of electrons, thereby facilitating the radiative recombination of injected charge carriers. To verify the conclusion, we fabricated the hole-only-device (HOD: ITO/PEDOT:PSS/TFB/QDs/MoO3/Al) and the electron-only-device (EOD: ITO/ZMO/QDs/ZMO/Al) (see figure S8 in supporting information for the diagrams of energy level), and the J-V curves are exhibited in figure 3(b). The QLEDs with ZnMy2-treated QDs show improved hole injection and decreased electron injection at any voltages, indicating more balanced charge injection and more effective radiative recombination of injected carriers. As can be seen from figure S9 in supporting information, the result was confirmed further by capacitance-voltage curves. The device with ZnMy2 has smaller peak capacitances than that of the device without ZnMy2, suggesting reduced charge accumulations and more effective radiative recombination.

    Figure  3.  (a) The energy levels of QLEDs (Its structure is ITO/PEDOT:PSS/PF8Cz/QD/ZnMgO/Al). (b) The current density-voltage (J-V) characteristics of hole-only-device (HOD: ITO/PEDOT:PSS/PF8Cz/QD/MoO3/Al) and electron-only-device (EOD: ITO/ZnMgO/QD/ZnMgO/Al). (c) The curves of current density and luminance with voltage variation for the device without and with ZnMy2. (d) The histogram distribution of luminance for 25 devices treated with ZnMy2. (e) The curves of EQE and current efficiency with luminance variation for the QLEDs without and with ZnMy2. (f) The operational lifetimes of QLEDs without and with ZnMy2.

    The J-L-V characteristics of the devices without and with ZnMy2 are demonstrated in figure 3(c). The device with ZnMy2 exhibits a smaller turn-on voltage (2.0 V) than that of the device without ZnMy2 (2.2 V), which may be related to the improved hole injection. In addition, with only a slightly larger current density (around 20% larger), the QLED based on ZnMy2-treated QDs shows much higher luminance within the entire voltage range, and the peak luminance reaches 175 084 cd m−2, which is improved by 67% compared to the device without ZnMy2 (105 127 cd m−2). As far as we know, the luminance is the state-of-the-art value among those of the previously reported green InP-based QLEDs [36, 37]. Importantly, devices can achieve high brightness even at lower voltages (e.g. 26 321 cd m−2@ 3.0 V and 83 834 cd m−2@ 4.0 V), which is beneficial for the reduction of joule heat, thereby improving the device stability. The histogram of the luminance of 25 devices indicates excellent fabrication reproducibility (figure 3(d)), and an average luminance of over 169 152 ± 842 cd m−2. The dramatically higher luminance but just slightly larger current density in the device with ZnMy2 means more effective radiative recombination efficiency of carriers. According to the EQE-current efficiency-luminance characteristics in figure 3(e), the peak EQE and current efficiency reach 12.74% and 54.56 cd A−1 at the luminance of 6643 cd m−2 for the device with ZnMy2, which is more than 38% higher than the maximum EQE of 9.22% and current efficiency of 36.72 cd A−1 for the QLEDs without ZnMy2. Finally, we compare the T50 lifetimes of the QLEDs with and without ZnMy2-treated QDs (figure 3(f)). When the initial luminance is 6987 cd m−2, the device with ZnMy2-treated QDs obtains a T50 lifetime of 9.6 h. According to the formula L0nT50 = constant (L0: initial luminance, n: acceleration factor, T50: half lifetime), the T50@100 cd m−2 lifetime of QLEDs with ZnMy2-treated QDs is calculated to be 20 044 h with n = 1.80. However, the T50@100 cd m−2 was only about 6078 h for the QLEDs without ZnMy2. The device performance is also shown in table 1.

    Table  1.  Summary of turn-on voltage (Von), EQE, current efficiency (ηA), maximum luminance (Lmax) and lifetime of our QLEDs.
    Von (V)EQE (%)ηA (cd A−1)Luminance (cd m−2)T50 lifetime (h)@100 cd m−2
    Without ZnMy22.29.2236.72105 1276078
    With ZnMy22.012.7454.56175 08420 044
     | Show Table
    DownLoad: CSV

    In summary, by in-situ passivating the InP core with ZnMy2, the resultant InP/ZnSe/ZnS QDs exhibit a PL peak of 534 nm and a high PL QY of 91%. Additionally, the corresponding green-emitting QLEDs show enhanced performance with an EQE of 12.74%, a peak current efficiency of 53.31 cd A−1, along with a breakthrough luminance of 175 084 cd m−2 and a long T50@100 cd m−2 lifetime of more than 20 000 h. The XPS and 1H NMR spectra demonstrated that the introduction of ZnMy2 not only acts as phosphine reservoir to control the monodisperse growth process, but also acts as a protective layer to protect the InP core from being affected by the water and oxygen in the environment.

    With the acceleration of the commercialization of QDs in terms of display and lighting, the performance enhancement of environmentally friendly QDs and corresponding electroluminescent devices has become an urgent task. To realize the commercialization of full-color environmentally friendly QLED technology, the performance of green InP-based devices still requires further effort by researchers. We demonstrated a strategy to in-situ passivate the InP cores, in which zinc myristate reacted with P dangling bonds to form Zn-P protective layer and protect InP cores from the water and oxygen in the environment. This proposed a safe and easy operating route for high-performance green InP-based devices.

    Funding

    This work was supported by the National Natural Science Foundation of China (Grant Nos. 62204078 and U22A2072) and the Natural Science Foundation of Henan Province for Excellent Youth Scholar (Grant No. 232300421092).

    Author contributions

    F Chen, Z Wu and H Shen conceived and supervised the project. Y Cheng and Q Li contributed equally. Y Cheng and Q Li synthesized and characterized the materials. M Chen and F Chen fabricated and characterized the devices. All authors participated in the scientific discussion and the manuscript modification.

    Conflict of interest

    The authors declare that they have no competing financial interests.

    These authors have made equal contributions to this work.

  • [1]
    Yadav R, Kwon Y, Rivaux C, Pierre C S, Ling W L, Reiss P 2023 Narrow near-infrared emission from InP QDs synthesized with indium (I) halides and aminophosphine J. Am. Chem. Soc. 145 5970-81 DOI: 10.1021/jacs.2c13834
    [2]
    Dumbgen K C, Leemans J, De Roo V, Minjauw M, Detavernier C, Hens Z 2023 Surface chemistry of InP quantum dots, amine-halide co-passivation, and binding of Z-type ligands Chem. Mater. 35 1037-46 DOI: 10.1021/acs.chemmater.2c02960
    [3]
    Zhou X, Ren J, Cao W, Meijerink A, Wang Y 2023 Narrow-band blue-emitting indium phosphide quantum dots induced by highly active Zn precursor Adv. Opt. Mater. 11 2202128 DOI: 10.1002/adom.202202128
    [4]
    Lee Y, Jo D Y, Kim T, Jo J H, Park J, Yang H, Kim D 2022 Effectual interface and defect engineering for Auger recombination suppression in bright InP/ZnSeS/ZnS quantum dots ACS Appl. Mater. Interfaces 14 12479-87 DOI: 10.1021/acsami.1c20088
    [5]
    Sung Y M, et al 2021 Increasing the energy gap between band-edge and trap states slows down picosecond carrier trapping in highly luminescent InP/ZnSe/ZnS quantum dots Small 17 2102792 DOI: 10.1002/smll.202102792
    [6]
    Dou Y, Wang L, Wang Y, Wu Q, Cao F, Wang S, Huang Q, Ma Y, Yang X 2023 Coordinating solvent synthesis of InP quantum dots with large sizes and suppressed defects for yellow light-emitting diodes Adv. Opt. Mater. 11 2300133 DOI: 10.1002/adom.202300133
    [7]
    Shin S, et al 2023 Fluoride-free synthesis strategy for luminescent InP cores and effective shelling processes via combinational precursor chemistry Chem. Eng. J. 466 143223 DOI: 10.1016/j.cej.2023.143223
    [8]
    Huang P, Liu X, Jin G, Liu F, Shen H, Li H 2023 Deep-red InP core-multishell quantum dots for highly bright and efficient light-emitting diodes Adv. Opt. Mater. 11 2300612 DOI: 10.1002/adom.202300612
    [9]
    Li H, Bian Y, Zhang W, Wu Z, Ahn T K, Shen H, Du Z 2022 High performance InP-based quantum dot light-emitting diodes via the suppression of field-enhanced electron delocalization Adv. Funct. Mater. 32 2204529 DOI: 10.1002/adfm.202204529
    [10]
    Won Y H, Cho O, Kim T, Chung D Y, Kim T, Chung H, Jang H, Lee J, Kim D, Jang E 2019 Highly efficient and stable InP/ZnSe/ZnS quantum dot light-emitting diodes Nature 575 634-8 DOI: 10.1038/s41586-019-1771-5
    [11]
    Ubbink R F, Almeida G, Iziyi H, du Fossé I, Verkleij R, Ganapathy S, van Eck E R H, Houtepen A J 2022 A water-free in situ HF treatment for ultrabright InP quantum dots Chem. Mater. 34 10093-103 DOI: 10.1021/acs.chemmater.2c02800
    [12]
    Fan X B, et al 2023 InP/ZnS quantum dot photoluminescence modulation via in situ H2S interface engineering Nanoscale Horiz. 8 522-9 DOI: 10.1039/d2nh00436d
    [13]
    Duan X, Ma J, Zhang W, Liu P, Liu H, Hao J, Wang K, Samuelson L, Sun X W 2023 Study of the interfacial oxidation of InP quantum dots synthesized from tris(dimethylamino)phosphine ACS Appl. Mater. Interfaces 15 1619-28 DOI: 10.1021/acsami.2c20138
    [14]
    Li H, Zhang W, Bian Y, Ahn T K, Shen H, Ji B 2022 ZnF2-assisted synthesis of highly luminescent InP/ZnSe/ZnS quantum dots for efficient and stable electroluminescence Nano Lett. 22 4067-73 DOI: 10.1021/acs.nanolett.2c00763
    [15]
    Wang H C, Zhang H, Chen H Y, Yeh H C, Tseng M R, Chung R J, Chen S, Liu R S 2017 Cadmium-free InP/ZnSeS/ZnS heterostructure-based quantum dot light-emitting diodes with a ZnMgO electron transport layer and a brightness of over 10 000 cd m−2 Small 13 1603962 DOI: 10.1002/smll.201603962
    [16]
    Zhang H, et al 2019 High-efficiency green InP quantum dot-based electroluminescent device comprising thick-shell quantum dots Adv. Opt. Mater. 7 1801602 DOI: 10.1002/adom.201801602
    [17]
    Moon H, Lee W, Kim J, Lee D, Cha S, Shin S, Chae H 2019 Composition-tailored ZnMgO nanoparticles for electron transport layers of highly efficient and bright InP-based quantum dot light emitting diodes Chem. Commun. 55 13299-302 DOI: 10.1039/C9CC06882A
    [18]
    Li Y, Hou X, Dai X, Yao Z, Lv L, Jin Y, Peng X 2019 Stoichiometry-controlled InP-based quantum dots: synthesis, photoluminescence, and electroluminescence J. Am. Chem. Soc. 141 6448-52 DOI: 10.1021/jacs.8b12908
    [19]
    Chao W C, Chiang T H, Liu Y C, Huang Z X, Liao C C, Chu C H, Wang C H, Tseng H W, Hung W Y, Chou P T 2021 High efficiency green InP quantum dot light-emitting diodes by balancing electron and hole mobility Commun. Mater. 2 96 DOI: 10.1038/s43246-021-00203-5
    [20]
    Zhang W, et al 2022 High quantum yield blue InP/ZnS/ZnS quantum dots based on bromine passivation for efficient blue light-emitting diodes Adv. Opt. Mater. 10 2200685 DOI: 10.1002/adom.202200685
    [21]
    Yu P, Cao S, Shan Y, Bi Y, Hu Y, Zeng R, Zou B, Wang Y, Zhao J 2022 Highly efficient green InP-based quantum dot light-emitting diodes regulated by inner alloyed shell component Light Sci. Appl. 11 162 DOI: 10.1038/s41377-022-00855-z
    [22]
    Park J, Won Y H, Kim T, Jang E, Kim D 2020 Electrochemical charging effect on the optical properties of InP/ZnSe/ZnS quantum dots Small 16 2003542 DOI: 10.1002/smll.202003542
    [23]
    Chen P, Liu H, Cui Y, Liu C, Li Y, Gao Y, Cheng J, He T 2023 Inner shell influence on the optical properties of InP/ZnSeS/ZnS quantum dots J. Phys. Chem. C 127 2464-70 DOI: 10.1021/acs.jpcc.3c00144
    [24]
    Okamoto A, Bai H, Toda S, Huang M, Kajii H, Kawai K, Murakami H 2023 Controlling thickness of ZnSe intermediate shell narrows FWHM of green-emitting spectra of InP/ZnSe/ZnS multi-shell quantum dots ChemNanoMat 9 e202200534 DOI: 10.1002/cnma.202200534
    [25]
    Kim T G, Zherebetskyy D, Bekenstein Y, Oh M H, Wang L W, Jang E, Alivisatos A P 2018 Trap passivation in indium-based quantum dots through surface fluorination: mechanism and applications ACS Nano 12 11529-40 DOI: 10.1021/acsnano.8b06692
    [26]
    Zhang X, Hudson M H, Castellano F N 2021 Passivation of electron trap states in InP quantum dots with benzoic acid ligands J. Phys. Chem. C 125 18362-71 DOI: 10.1021/acs.jpcc.1c05594
    [27]
    Pu Y C, Fan H C, Chang J C, Chen Y H, Tseng S W 2021 Effects of interfacial oxidative layer removal on charge carrier recombination dynamics in InP/ZnSexS1-x core/shell quantum dots J. Phys. Chem. Lett. 12 7194-200 DOI: 10.1021/acs.jpclett.1c02125
    [28]
    Yang W, Yang Y, Kaledin A L, He S, Jin T, McBride J R, Lian T 2020 Surface passivation extends single and biexciton lifetimes of InP quantum dots Chem. Sci. 11 5779-89 DOI: 10.1039/D0SC01039A
    [29]
    Yoo D, Bak E, Ju H M, Shin Y M, Choi M J 2022 Zinc carboxylate surface passivation for enhanced optical properties of In(Zn)P colloidal quantum dots Micromachines 13 1775 DOI: 10.3390/mi13101775
    [30]
    Wu Q, et al 2022 Quasi-shell-growth strategy achieves stable and efficient green InP quantum dot light-emitting diodes Adv. Sci. 9 2200959 DOI: 10.1002/advs.202200959
    [31]
    Jo J H, Jo D Y, Choi S W, Lee S H, Kim H M, Yoon S Y, Kim Y, Han J N, Yang H 2021 Highly bright, narrow emissivity of InP quantum dots synthesized by aminophosphine: effects of double shelling scheme and Ga treatment Adv. Opt. Mater. 9 2100427 DOI: 10.1002/adom.202100427
    [32]
    Taylor D A, Teku J A, Cho S, Chae W S, Jeong S J, Lee J S 2021 Importance of surface functionalization and purification for narrow FWHM and bright green-emitting InP core-multishell quantum dots via a two-step growth process Chem. Mater. 33 4399-407 DOI: 10.1021/acs.chemmater.1c00348
    [33]
    Xu H, et al 2024 Dipole-dipole-interaction-assisted self-assembly of quantum dots for highly efficient light-emitting diodes Nat. Photon. 18 186-91 DOI: 10.1038/s41566-023-01344-4
    [34]
    Deng Y, et al 2022 Solution-processed green and blue quantum-dot light-emitting diodes with eliminated charge leakage Nat. Photon. 16 505-11 DOI: 10.1038/s41566-022-00999-9
    [35]
    Han C Y, Lee S H, Song S W, Yoon S Y, Jo J H, Jo D Y, Kim H M, Lee B J, Kim H S, Yang H 2020 More than 9% efficient ZnSeTe quantum dot-based blue electroluminescent devices ACS Energy Lett. 5 1568-76 DOI: 10.1021/acsenergylett.0c00638
    [36]
    Zhang T, et al 2023 Electric dipole modulation for boosting carrier recombination in green InP QLEDs under strong electron injection Nanoscale Adv. 5 385-92 DOI: 10.1039/d2na00705c
    [37]
    Kim J, Hong A, Hahm D, Lee H, Bae W K, Lee T, Kwak J 2023 Realization of highly efficient InP quantum dot lightemitting diodes through indepth investigation of excitonharvesting layers Adv. Opt. Mater. 11 2300088 DOI: 10.1002/adom.202300088
  • Related Articles

    [1]Xiang Xu, Yi Wang, Siyu Zhu, Qian Xu, Zulan Liu, Guotao Cheng, Dingpei Long, Lan Cheng, Fangyin Dai. Functionalized bio-spinning silk fiber scaffolds containing Mg2+ with osteoimmunomodulatory and osteogenesis abilities for critical-sized bone defect regeneration[J]. Materials Futures, 2025, 4(1): 015401. DOI: 10.1088/2752-5724/ad9e09
    [2]Amir Motaharinia, Jaroslaw W. Drelich, Safian Sharif, Ahmad Fauzi Ismail, Farid Naeimi, Alexandra Glover, Mahshid Ebrahiminejad, Hamid Reza Bakhsheshi-Rad. Overview of porous magnesium-based scaffolds: development, properties and biomedical applications[J]. Materials Futures, 2025, 4(1): 012401. DOI: 10.1088/2752-5724/ad9493
    [3]Binodhya Wijerathne, Ting Liao, Xudong Jiang, Juan Zhou, Ziqi Sun. Plant-inspired surfaces and interfaces for sustainable technologies[J]. Materials Futures, 2025, 4(1): 012301. DOI: 10.1088/2752-5724/ad93ea
    [4]Yongxin Huang, Yiqing Wang, Xiyue Peng, Tongen Lin, Xia Huang, Norah S Alghamdi, Masud Rana, Peng Chen, Cheng Zhang, Andrew K Whittaker, Lianzhou Wang, Bin Luo. Enhancing performance and longevity of solid-state zinc-iodine batteries with fluorine-rich solid electrolyte interphase[J]. Materials Futures, 2024, 3(3): 035102. DOI: 10.1088/2752-5724/ad50f1
    [5]Pin Tian, Libin Tang, Kar-Seng Teng, Shu-Ping Lau. Graphene quantum dots: preparations, properties, functionalizations and applications[J]. Materials Futures, 2024, 3(2): 022301. DOI: 10.1088/2752-5724/ad08cb
    [6]Yuxin Cui, Ruixin Zhang, Sinian Yang, Lili Liu, Shimou Chen. Research progress on the design of electrolyte additives and their functions for zinc-ion batteries[J]. Materials Futures, 2024, 3(1): 012102. DOI: 10.1088/2752-5724/acef41
    [7]Wanyao Zhang, Yufang Chen, Hongjing Gao, Wei Xie, Peng Gao, Chunman Zheng, Peitao Xiao. Review of regulating Zn2+ solvation structures in aqueous zinc-ion batteries[J]. Materials Futures, 2023, 2(4): 042102. DOI: 10.1088/2752-5724/ace3de
    [8]Yichuan Rui, Zuoming Jin, Xinyi Fan, Weitao Li, Bin Li, Tianpeng Li, Yuanqiang Wang, Liang Wang, Jia Liang. Defect passivation and electrical conductivity enhancement in perovskite solar cells using functionalized graphene quantum dots[J]. Materials Futures, 2022, 1(4): 045101. DOI: 10.1088/2752-5724/ac9707
    [9]Zhaodong Huang, Rong Zhang, Shaoce Zhang, Pei Li, Chuan Li, Chunyi Zhi. Recent advances and future perspectives for aqueous zinc-ion capacitors[J]. Materials Futures, 2022, 1(2): 022101. DOI: 10.1088/2752-5724/ac4263
    [10]Danyang Li, Junjie Wang, Miaozi Li, Biao Guo, Lan Mu, Yu Luo, Yi Xiao, Chaohuang Mai, Jian Wang, Junbiao Peng. Efficient red perovskite quantum dot light-emitting diode fabricated by inkjet printing[J]. Materials Futures, 2022, 1(1): 015301. DOI: 10.1088/2752-5724/ac3568
  • Other Related Supplements

  • Cited by

    Periodical cited type(4)

    1. Lee, S.Y., Sakong, C., Ju, B.-K. et al. Enhancing the reliability of InP-based QD color conversion layer through a uniform organic encapsulation layer via inkjet printing. Organic Electronics, 2024. DOI:10.1016/j.orgel.2024.107136
    2. Luo, Y., Cao, X., Wang, S. et al. Suppression of Interfacial Oxidation in Core/Shell InP Quantum Dots through Solvent Assisted Core-Etching Strategy for Efficient Green Light-Emitting Diodes. Nano Letters, 2024. DOI:10.1021/acs.nanolett.4c05832
    3. Chen, H.-S., Yeh, C.-W., Lee, H.-Y. et al. Preparing Smaller InP Quantum Dots by Suppressing Over-Etch Using Core Protective Layer and Ammonium Fluoride as Alternative Etchant. Advanced Optical Materials, 2024. DOI:10.1002/adom.202402051
    4. Wang, S., Yang, W., Li, Y. et al. Stable and Efficient Red InP-Based QLEDs through Surface Passivation Strategies of Quantum Dots. Nano Letters, 2024. DOI:10.1021/acs.nanolett.4c04580

    Other cited types(0)

Catalog

    Figures(3)  /  Tables(1)

    Article Metrics

    Article views (533) Full Text (256) PDF downloads (45) Cited by(4)
    Related

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return