TY - JOUR AU1 - Shang, Yuequn AU2 - Ning, Zhijun AB - Abstract The application of colloidal quantum dots for light-emitting devices has attracted considerable attention in recent years, due to their unique optical properties such as size-dependent emission wavelength, sharp emission peak and high luminescent quantum yield. Tremendous efforts have been made to explore quantum dots for light-emission applications such as light-emitting diodes (LEDs) and light converters. The performance of quantum-dots-based light-emitting diodes (QD-LEDs) has been increasing rapidly in recent decades as the development of quantum-dots synthesis, surface-ligand engineering and device-architecture optimization. Recently, the external quantum efficiencies of red quantum-dots LEDs have exceeded 20.5% with good stability and narrow emission peak. In this review, we summarize the recent advances in QD-LEDs, focusing on quantum-dot surface engineering and device-architecture optimization. colloidal quantum dots, light-emitting diodes, surface engineering, device architecture INTRODUCTION Semiconductor colloidal quantum dots (QDs), due to their unique optical properties including narrow and finely tunable emission spectra from the ultraviolet (UV) into the infrared (IR), high photoluminescence quantum yield (PLQY) and solution processability [1–6], have attracted intensive interest for solar cells [7–13], biosensors [14,15], photodetectors [16–20], field-effect transistors [21–23], lasers [24,25], light-emitting diodes (LEDs) [26–32] and displays [30,33,34]. Especially, its high color purity and color-tunable emission make quantum-dots-based light-emitting diodes (QD-LEDs) a promising candidate for next-generation displays [35–38]. QDs-based light converters have been applied for displays for TVs and mobile phones. However, the development of the light converter relies on blue LED chips as the back light source. In recent years, the development of active-mode QD-LEDs has received great interest due to its low cost, low energy consumption and variability. In order to improve the performance of QD-LEDs, great efforts have been made to improve the PLQY of QDs film, carrier-injection efficiency from carrier-transporting layers to QDs, as well as to reduce interface carrier recombination. As a result, the external quantum efficiency of QD-LEDs has increased from ∼0.1 to 20.5% [29,39], which is comparable with that of the state-of-the-art LEDs based on organic semiconductors. QDs mainly contain three major categories: II–VI group QDs such as CdSe, PbS; III–V QDs such as InP QDs; and halide perovskite. II–VI group QDs normally show excellent luminescent character, but they are generally toxic. III–V group QDs are less toxic, but their optic characteristics are worse than those in II–VI group QDs. Recently, halide perovskite QDs have aroused great interest for their simple synthesis and high luminescence yield, and they have become one important category [5,38,40–44]. Although QDs generally possess high PLQY in solution, the realization of high PL QY QD film is much more difficult. Firstly, the ligand loss during film fabrication can bring about the formation of defects. Effective surface-ligand passivation is therefore critical for colloidal QDs to keep their quantum confinement character and high luminescence quantum yield. In addition, in film, carrier transport between QDs may deteriorate carrier recombination, as well as non-radiative Auger recombination. Shell growth is an effective strategy to relax the impact of ligand loss [39,45]. In recent years, surface-ligand modification was widely explored to enhance the bonding strength of ligands in order to prevent ligand desorption, improve carrier injection and maintain charge balance [29,46,47]. Recently, in-situ growth of ligands was realized by epitaxy growth of perovskite on QD surface in film, which presented a new platform for reducing defects [48]. Apart from the improvement of active materials, great progress has been made in device architecture. In the early stages, QD-LEDs generally utilized QDs embedded in an organic host as the carrier-transporting matrix [49]. This strategy endows QDs with high PLQY in film. Nevertheless, carrier injection from matrix to QDs was hampered by the alkyl chain ligands. In recent years, much effort has been devoted to enhancing the carrier transport through the introduction of an inorganic or organic/inorganic hybrid matrix [48,50,51]. A sandwiched structure that integrates pure QD films between electrons and hole-transporting layers was explored as well to enhance carrier transport. Meanwhile, much work has been made to optimize carrier-transporting layers including tailoring of the band alignment and the use of carrier-blocking layers [29]. In this review, we will firstly introduce QD-LEDs, the concept and figure of merits for readers. We will then discuss QD surface-ligand engineering and the use of core–shell architecture. Device-architecture improvement, such as active-layer and transporting-layer optimization, will subsequently be presented including the analysis of the pros and cons of each device architecture. Finally, a perspective on development trends will be presented. A BRIEF INTRODUCTION OF QD-LEDs The typical dimension of QDs lies at around 1–10 nm and they are composed of a few hundred to 10 000 atoms [52]. When excitons are confined in three spatial dimensions, discrete electronic states will replace the continuous energy bands of a bulk material, and their electronic band structure becomes strongly size-dependent, as shown in Fig. 1a [53–55]. QDs therefore exhibit unique properties such as band-gap tunability, sharp emission peak and high PLQY. Figure 1b shows the emission range of classic QDs. A summary of commonly used QDs is shown in Table 1. Figure 1. Open in new tabDownload slide (a) Electronic states in a bulk semiconductor (left) and a QD made of the same material (right). Continuous bands of a bulk semiconductor with parabolic dispersion of carrier kinetic energies in the valence bands (VBs) and conduction bands (CBs) transform into discrete atomic-like levels in the case of the QDs. (b) Previous heavy-metal-based QDs covering the ultraviolet (CdS), visible (CdSe, CdTe and perovskite) and infrared (PbSe and PbS) regions have heavy metal-free alternatives: ZnSe for ultraviolet; InP, CuInS2 and Si for visible; InAs for infrared. Adapted with permission from [39]. (c) Energy-band diagram of a typical QD-LED that outlines the two suspected QD excitation mechanisms: charge injection and energy transfer. (d) Progression of QD-LED performance over time in terms of peak EQE. Red ball, red QD-LEDs; green ball, green QD-LEDs; blue ball: blue QD-LEDs. Inset: four types of device structure. Figure 1. Open in new tabDownload slide (a) Electronic states in a bulk semiconductor (left) and a QD made of the same material (right). Continuous bands of a bulk semiconductor with parabolic dispersion of carrier kinetic energies in the valence bands (VBs) and conduction bands (CBs) transform into discrete atomic-like levels in the case of the QDs. (b) Previous heavy-metal-based QDs covering the ultraviolet (CdS), visible (CdSe, CdTe and perovskite) and infrared (PbSe and PbS) regions have heavy metal-free alternatives: ZnSe for ultraviolet; InP, CuInS2 and Si for visible; InAs for infrared. Adapted with permission from [39]. (c) Energy-band diagram of a typical QD-LED that outlines the two suspected QD excitation mechanisms: charge injection and energy transfer. (d) Progression of QD-LED performance over time in terms of peak EQE. Red ball, red QD-LEDs; green ball, green QD-LEDs; blue ball: blue QD-LEDs. Inset: four types of device structure. Table 1. Summary of commonly used QDs. QD type . Lattice structure . Egap . Emission wavelength . PLQY . Ref. . Heavy-metal QDs CdSe, CdSe/CdS Wurtzite/ 1.74 ∼480–700 Up to 100% [56–58] Zinc blende CdS Wurtzite/ 2.49 ∼380–750 69% [59] Zinc blende PbS Rocksalt 0.41 Abs peak 830–2200 20–90% [10,60] PbSe Rocksalt 0.28 Abs peak 848–1896 Up to 43% [10,61] ABX3 A = MA, FA, Cs B = Pb, Sn, BiX = Cl, Br, I Perovskite 1.55–3.11 410–700 Up to 100% [41,42,62–65] Heavy-metal-free QDs InP, InP/ZnS Zinc blende 1.35 480–750 Up to 80% [66–68] CuInS2, CuInS2/ZnS Wurtzite 1.5 550–800 20–80% [69–71] AgInS2, AgInS2/ZnS Wurtzite 1.87 540–800 Up to 60% [72–74] Mn:ZnS, Cu:ZnS Zinc blende 3.7 480–580 Up to 50% [75,76] Silicon Diamond 1.1 420–1000 Up to 40% [77–79] Carbon Layered structure 0 430–650 Up to 54.5% [80,81] QD type . Lattice structure . Egap . Emission wavelength . PLQY . Ref. . Heavy-metal QDs CdSe, CdSe/CdS Wurtzite/ 1.74 ∼480–700 Up to 100% [56–58] Zinc blende CdS Wurtzite/ 2.49 ∼380–750 69% [59] Zinc blende PbS Rocksalt 0.41 Abs peak 830–2200 20–90% [10,60] PbSe Rocksalt 0.28 Abs peak 848–1896 Up to 43% [10,61] ABX3 A = MA, FA, Cs B = Pb, Sn, BiX = Cl, Br, I Perovskite 1.55–3.11 410–700 Up to 100% [41,42,62–65] Heavy-metal-free QDs InP, InP/ZnS Zinc blende 1.35 480–750 Up to 80% [66–68] CuInS2, CuInS2/ZnS Wurtzite 1.5 550–800 20–80% [69–71] AgInS2, AgInS2/ZnS Wurtzite 1.87 540–800 Up to 60% [72–74] Mn:ZnS, Cu:ZnS Zinc blende 3.7 480–580 Up to 50% [75,76] Silicon Diamond 1.1 420–1000 Up to 40% [77–79] Carbon Layered structure 0 430–650 Up to 54.5% [80,81] Open in new tab Table 1. Summary of commonly used QDs. QD type . Lattice structure . Egap . Emission wavelength . PLQY . Ref. . Heavy-metal QDs CdSe, CdSe/CdS Wurtzite/ 1.74 ∼480–700 Up to 100% [56–58] Zinc blende CdS Wurtzite/ 2.49 ∼380–750 69% [59] Zinc blende PbS Rocksalt 0.41 Abs peak 830–2200 20–90% [10,60] PbSe Rocksalt 0.28 Abs peak 848–1896 Up to 43% [10,61] ABX3 A = MA, FA, Cs B = Pb, Sn, BiX = Cl, Br, I Perovskite 1.55–3.11 410–700 Up to 100% [41,42,62–65] Heavy-metal-free QDs InP, InP/ZnS Zinc blende 1.35 480–750 Up to 80% [66–68] CuInS2, CuInS2/ZnS Wurtzite 1.5 550–800 20–80% [69–71] AgInS2, AgInS2/ZnS Wurtzite 1.87 540–800 Up to 60% [72–74] Mn:ZnS, Cu:ZnS Zinc blende 3.7 480–580 Up to 50% [75,76] Silicon Diamond 1.1 420–1000 Up to 40% [77–79] Carbon Layered structure 0 430–650 Up to 54.5% [80,81] QD type . Lattice structure . Egap . Emission wavelength . PLQY . Ref. . Heavy-metal QDs CdSe, CdSe/CdS Wurtzite/ 1.74 ∼480–700 Up to 100% [56–58] Zinc blende CdS Wurtzite/ 2.49 ∼380–750 69% [59] Zinc blende PbS Rocksalt 0.41 Abs peak 830–2200 20–90% [10,60] PbSe Rocksalt 0.28 Abs peak 848–1896 Up to 43% [10,61] ABX3 A = MA, FA, Cs B = Pb, Sn, BiX = Cl, Br, I Perovskite 1.55–3.11 410–700 Up to 100% [41,42,62–65] Heavy-metal-free QDs InP, InP/ZnS Zinc blende 1.35 480–750 Up to 80% [66–68] CuInS2, CuInS2/ZnS Wurtzite 1.5 550–800 20–80% [69–71] AgInS2, AgInS2/ZnS Wurtzite 1.87 540–800 Up to 60% [72–74] Mn:ZnS, Cu:ZnS Zinc blende 3.7 480–580 Up to 50% [75,76] Silicon Diamond 1.1 420–1000 Up to 40% [77–79] Carbon Layered structure 0 430–650 Up to 54.5% [80,81] Open in new tab A classic device architecture of QD-LEDs is shown in Fig. 1c, which consists of an active emission layer, an electron-transporting layer (ETL) and a hole-transporting layer (HTL). Electrons and holes are injected from the ETL and HTL, respectively, and recombine at the active layer to drive light emission. By tuning the size and composition of QDs, the emission wavelength can be conveniently tuned. Figure 1d shows the efficiency development of QD-LEDs in recent decades. Ever since the first demonstration of QD-LEDs in 1994 [82], a considerable improvement in device efficiency and stability have been realized [29,30]. Up to now, the demonstrated record peak external quantum efficiency (EQE) and luminance for QD-LEDs have reached 20.5% and 106 000 Cd/m2 for red [29,83], 14.5% and 218 800 Cd/m2 for green [30,84] and 12.2% and 7600 Cd/m2 for blue [85], respectively. Apart from the single-color device, full-color display has also been demonstrated [33,86,87]. Figures of merit for QD-LEDs Luminance, EQE, luminous efficiency and lifetime are critical figures of merit for judging device performance. The luminance of a device is defined by the derivative \begin{equation} {L_v} = \frac{{{d^2}{\Phi _v}}}{d\sum \,d{\Omega _{\scriptstyle\sum}}\cos {\theta _{\scriptstyle\sum}}},\end{equation} (1) where Lv is the luminance (in Cd/m2), d2Φv is the luminous flux (in lm) output from the area dΣ in any direction contained inside the solid angle dΩΣ, dΣ is an infinitesimal area (in m2) of the source, dΩΣ is an infinitesimal solid angle (sr) and θΣ is the angle between the normal nΣ to the surface. EQE is the generally used parameter for evaluating the performance of LED, which is defined as the ratio of the emitted photons from the device to the injected electrons: \begin{equation}EQE = \frac{1}{{hv}}{P_0}\frac{e}{i}, \end{equation} (2) where i is the electrical current, e is the electron charge, hν is the photon energy and P0 is the output power [88,89]. Luminous efficiency is a parameter weighing the conversion efficiency of visible light emission, which is defined as the ratio of luminous flux to power (lm/W) [90]. Turn-on voltage is the minimum voltage required to drive the current across the diode and start light-emitting. Full width at half maximum (FWHM) is the width of a spectrum range between points on the y-axis that are half the maximum amplitude. A small FWHM value is required for high color purity. QD SURFACE ENGINEERING Surface shell growth The nanoscale QDs have a high surface-to-volume ratio. A significant fraction of atoms are located on the surface, and some atoms are not fully coordinated, which need to be passivated by surface ligands. The loss of ligands leads to the formation of dangling bonds and defects on the surface, which can be evidenced by low PL QY. Overcoating QDs with higher-band-gap inorganic shell material can confine carriers to the central part of QDs (Fig. 2a) [57,91–96], which can decrease the reactivity of the QD surface with oxygen and water, as well as create the possibility for the formation of a dangling bond [94]. Figure 2. Open in new tabDownload slide (a) Schematic illustration of shell growth. Route I: layer-by-layer growth of CdSe–CdS core–shell QDs. Route II: The cation-exchange reaction used to convert core-only PbS QDs into core–shell PbS–CdS QDs. (b) Evolution of PL QY of zinc blende (ZB) and wurtzite (WZ) CdSe/CdS core/shell nanocrystals. Adapted with permission from [56]. (c) Schematic of the chemical composition and electronic energy level within a graded-shell QD. Adapted with permission from [103]. (d) EQE versus current density of QD-LEDs with different types of QDs. The onset of EQE roll-off shifts to higher currents with increasing degree of suppression of Auger recombination. Adapted with permission from [107]. Figure 2. Open in new tabDownload slide (a) Schematic illustration of shell growth. Route I: layer-by-layer growth of CdSe–CdS core–shell QDs. Route II: The cation-exchange reaction used to convert core-only PbS QDs into core–shell PbS–CdS QDs. (b) Evolution of PL QY of zinc blende (ZB) and wurtzite (WZ) CdSe/CdS core/shell nanocrystals. Adapted with permission from [56]. (c) Schematic of the chemical composition and electronic energy level within a graded-shell QD. Adapted with permission from [103]. (d) EQE versus current density of QD-LEDs with different types of QDs. The onset of EQE roll-off shifts to higher currents with increasing degree of suppression of Auger recombination. Adapted with permission from [107]. In addition to the energy-level alignment, the match of the crystal structure and the lattice constant between the core and the shell are important to avoid lattice strain at the interface. The first reported prototype core/shell structure is in CdSe/ZnS QDs, in which the ZnS shell significantly improves the PLQY and stability against photo bleaching. When the shell becomes thick, the strain-induced defects increase [97], resulting in the decrease of PLQY. When the shell thickness was more than three monolayers thick, however, the impact of surface states on QDs was relaxed (Fig. 2b.) Apart from ZnS, CdS [56,98] and ZnSe [99] are generally used shell materials to coat CdSe QDs as well. The lattice mismatches are about 4% and 6%, respectively, resulting in much reduced lattice strain at the interface [45]. Lim et al. studied the influence of shell thickness on the performance of corresponding CdSe/Zn1–xCdxS QD-LEDs. Based on the optimization of shell thickness, they realized the highest EQE over 7% and record-high luminance of 105 870 cd/m2 along with improved device stability. In addition, they revealed that the QD–QD energy transfer, an important source of energy loss, can be inhibited by a thick shell [83]. The Bawendi group studied the effect of shell thickness on emission line widths and photoluminescence blinking. As the shell thickness increases, the emission line width decreases and photoluminescence blinking is suppressed [100–102]. In order to reduce the strain-induced defects in the core–shell interface, a graded, alloyed intermediate shell is introduced to the core–shell structure (Fig. 2c) [103]. The graded shell effectively suppresses the non-radiative decay and Auger recombination, leading to enhanced PLQY and device performance (Fig. 2d) [30,92,95,104–107]. Overall, the use of the core–shell structure reduced surface defects, significantly boosted PLQY and improved the stability of the film, which is important for device fabrication. Surface-ligand engineering QDs, one kind of nanocrystals (NCs) are typically synthesized in a solution containing surface ligands (e.g. oleic acid, oleylamine) with a head-group tethered to the QD surface and a hydrocarbon tail directed away [108]. The surface ligands act as stoppers to control the QD growth and surface-passivating ligands prohibit the formation of dangling bonds [53,109]. The commonly used head-groups of the ligands are carboxyl, amino, thiol, phosphate group, etc. Recently, halide ligands received great attention for their excellent surface-passivation capability [11,110,111]. The Owen group systematically summarized the ligand coordinating types on the QD surface, which can be classified into three main categories. The first type is the L-type ligand, i.e. a neutral donor that datively bonds onto the QD surface. Phosphine and amine are the L-type ligands usually used. The bonds between ligands and QDs are typically weak, and they can desorb in the presence of other ligands in the system. Another kind of ligand is the X-type ligands, which include the carboxyl group, thiol group and halides, etc., which usually bond strongly on the QD surface due to their high electron affinity. Cations on QD surfaces can bond with one or two X-type ligands, depending on their coordinating numbers. Both cations and anions can bond with L-type ligands; however, the bond is generally weak and the ligands are easy to desorb (Fig. 3) [112]. Figure 3. Open in new tabDownload slide (a) Examples of several ligand-exchange reactions are shown. (b) The coordination of different types of ligands in Green's formulation to metal-chalcogenide nanocrystals (such as cadmium selenide) are illustrated. R is analkyl group; Bu is n-butyl. Adapted with permission from [112]. Figure 3. Open in new tabDownload slide (a) Examples of several ligand-exchange reactions are shown. (b) The coordination of different types of ligands in Green's formulation to metal-chalcogenide nanocrystals (such as cadmium selenide) are illustrated. R is analkyl group; Bu is n-butyl. Adapted with permission from [112]. The use of proton solvent in post treatment can cause the desorption of ligands such as carboxyl or amino groups, especially during the solid-state ligand-exchange process. An important strategy to reduce ligand loss is the use of ligands with strong binding characteristics such as thiol ligands. In addition, the long organic alkyl ligands used in synthesis can block carrier injection. Ligand exchange by short ligands is therefore needed to improve carrier transfer and promote the balance transport of electrons and holes. Short alkyl thiol ligands are therefore the most common ligands used in QD surface modification [113–115]. In 2012, Sun et al. utilized a series of ligands with both thiol and carboxylic acids units to replace the long-chain oleate ligands of PbS QDs. The distance between QDs can be tuned by varying the ligand-chain lengths. Based on the optimization of ligand length to balance carrier injection and effective surface passivation, an order-of-magnitude improvement in device efficiency was achieved (Fig. 4a and b) [116]. Figure 4. Open in new tabDownload slide (a) Schematic device structure of PbS QD-LEDs. (b) shows the distance between adjacent PbS colloidal QDs. Adapted with permission from [116]. (c) Solubility of CdSe QD−ligand complexes with different ligands. (The solubility increased significantly for the entropic ligand-coated CdSe QDs.) (d) External power efficiency of the LEDs with the same CdSe–CdS core–shell QDs and device structure but different ligands. Adapted with permission from [47]. Figure 4. Open in new tabDownload slide (a) Schematic device structure of PbS QD-LEDs. (b) shows the distance between adjacent PbS colloidal QDs. Adapted with permission from [116]. (c) Solubility of CdSe QD−ligand complexes with different ligands. (The solubility increased significantly for the entropic ligand-coated CdSe QDs.) (d) External power efficiency of the LEDs with the same CdSe–CdS core–shell QDs and device structure but different ligands. Adapted with permission from [47]. The Peng group reported a solution-processed, high-performance QD-LED based on ligand-exchanged CdSe/CdS QDs. They used 1-dodecanethiol to replace the original oleate ligands,together with device structural engineering. The device showed the highest EQE up to 20.5% and a long operational lifetime of more than 100,000 hours at 100 cd/m2, which is the best-performing solution-processed red QD-LED so far [29]. Ligands engineering was utilized to improve the efficiency of CdZnS–ZnS QD-based blue light-emitting diodes as well. The replacement of oleic acid ligands on the as-synthesized QDs with 1-octanethiol ligands resulted in a double increased electron mobility and greater balanced carrier injection, leading to the highest EQE of 12.2% [46]. Quite recently, branched ligands, namely entropic ligands for QDs, were developed, leading to much improved solubility of QDs by 102−106. High concentration of QDs in solvent is important for large-scale printing techniques, as well as for large-scale application of QD-based optoelectronics (Fig. 4c). Based on branched thiol ligand-coated CdSe–CdScore–shell QDs, a device with better charge transport and higher EQE was achieved (Fig. 4d) [47,117]. Overall, the surface ligands play a crucial role for QD surface passivation, efficient charge transport and carrier balance. Delicate surface-ligand engineering is essential to achieve high efficiency and stable QD-LED devices. THE OPTIMIZATION OF DEVICE ARCHITECTURE Apart from the high PLQY of QD active layers, the use of appropriate device architecture to ensure efficient carrier injection and radiative recombination in QDs plays a critical role for high-performance light-emitting diodes. To date, various device architectures, ETLs and HTLs (Fig. 5) have been employed to enhance the device performance. Meanwhile, the interface states between charge-injection layer and active layer have been carefully tailored to reduce non-radiative recombination on the interface. In this section, we will discuss the modification of charge-transport layers and the organic/inorganic matrix for QD-LEDs. We will divide the structures into two parts: planar architecture and QDs in a matrix. For the matrix part, we will classify it into organic matrix, inorganic matrix and hybrid perovskite. Figure 5. Open in new tabDownload slide The energy-band level of general transport layers. Figure 5. Open in new tabDownload slide The energy-band level of general transport layers. Planar QD-LEDs At present, QD-LEDs mainly use planar architecture with a flat active layer sandwiched between ETLs and HTLs [27–30,84,96,118]. In this structure, the active layer is typically very thin given the poor conductivity of the QDs. The balanced injection of electrons and holes is critical for efficient radiative recombination at the active layer, as well as a long device-operation lifetime. The imbalanced carrier injection gives rise to electron charging on QD layers and consequently non-radiative Auger recombination, as well as carrier recombination at the transporting layers [92,119]. Energy-level engineering of the transport layer through ion dopant or surface modification was introduced to tailor the band structure of the carrier-transporting layer [120–125]. Mg-doped ZnO(Mg:ZnO) nanoparticles (NPs) that possess lower electronic energy levels were employed as the ETL (Fig. 6a). The Mg dopant is beneficial to reduce the carrier-injection barrier and improve the carrier-injection efficiency. Devices with Mg: ZnO NPs ETL show superior performance compared to pure ZnO in terms of luminance and efficiency (Fig. 6b) [126]. Figure 6. Open in new tabDownload slide (a) Proposed energy-band diagram of solution-processed multilayered QD-LEDs. (b) ZnMgO NP ETL-dependent variations of current efficiency and EQE of CIS/ZnS QD-LEDs as a function of driving voltage. Adapted with permission from [126]. (c) Flat-band energy-level diagrams of QD-LEDs illustrate the reduction in the electron-injection barrier between ZnO and QDs due to the presence of the PFN layer. (d) EQE versus current density characteristics of InP@ZnSeS (1.7-nm shell thickness) QLEDs prepared with varied PFN concentrations. Adapted with permission from [128]. Figure 6. Open in new tabDownload slide (a) Proposed energy-band diagram of solution-processed multilayered QD-LEDs. (b) ZnMgO NP ETL-dependent variations of current efficiency and EQE of CIS/ZnS QD-LEDs as a function of driving voltage. Adapted with permission from [126]. (c) Flat-band energy-level diagrams of QD-LEDs illustrate the reduction in the electron-injection barrier between ZnO and QDs due to the presence of the PFN layer. (d) EQE versus current density characteristics of InP@ZnSeS (1.7-nm shell thickness) QLEDs prepared with varied PFN concentrations. Adapted with permission from [128]. Polyethylenimine (PEI) was employed to modify ZnO ETL surfaces, which facilitates the electron injection into CdSe–ZnS QD emissive layers by lowering the work function of the ZnO layer from 3.58 eV to 2.87 eV, resulting in a much reduced interface barrier. Eventually, better charge injection and carrier balance in the emission layer is achieved. As a result, red QD-LEDs with a ZnO/PEI ETL layer exhibit a low turn-on voltage of 2–2.5 V [127]. A solution-processed thin conjugated polyelec-trolyte (poly[(9,9-bis(30-(N,N-dimethylamino) propyl)-2,7-fluorene)-alt-2,7-(9,9-ioctylfluorene)) (PFN) layer was applied in green QD-LEDs based on the InP/ZnSeS QDs. The PFN layer acted as an interfacial dipole layer between ZnO ETLs and QDs (Fig. 6c), leading to a reduced electron-injection barrier, and promoted the charge balance within QD layers, so the performance of the device is therefore improved significantly (Fig. 6d) [128]. In addition to the energy-level engineering, extra carrier-blocking layers have been explored in QD-LED devices [29,30,122,128–131]. The additional carrier-blocking layers help to maintain the charge neutrality of QD emitters and preserve superior emissive properties. Furthermore, an additional buffer layer with a wide band gap can block the reverse transfer of electrons or holes and overshooting of carriers out of the active layer, resulting in better charge balance and suppressed carrier recombination at the active layer. A deoxyribonucleic acid and cetyltrimetylammonium (DNA-CTMA) complex was employed as an additional hole-transporting and electron-blocking layer (EBL) for CdSe/CdS/ZnS QD-LEDs. With the inclusion of this blocking layer, no electroluminescence (EL) emission from the HTLs/ETLs was observed, indicating that electron injection into HTLs was completely blocked by the DNA-CTMA EBLs [130]. Similarly, the incorporation of an insulating polymer buffer layer between ETLs and QD layers can prevent the overshooting of electrons into active layers (Fig. 7a), giving rise to much better charge balance and a much improved device performance (Fig. 7b). Furthermore, the insertion of this insulation layer can reduce charge recombination at the interface between QD layers and ETLs [29]. Based on this principle, PVK was employed as the HTL for blue QD-LEDs (Fig. 7c and d) [30]. To date, the highest EQE for red, green and blue QD-LEDs reached up to 20.5%, 14.5% and 10.7%, respectively, and all of them rely on polymer as a buffer layer [29,30]. In conclusion, the use of energy-level engineering and a carrier-blocking layer plays an important role for high-performance QD-LED devices. Figure 7. Open in new tabDownload slide (a) Device structure of multilayer QD-LED devices with Polymethyl methacrylate (PMMA) as a blocking layer. (b) Current density and luminance versus voltage characteristics for QLEDs without and with the 6-nm PMMA layer. Adapted with permission from [29]. (c) Schematic of blue QD-LED devices structure with poly[(9,9-dioctylfluorenyl-2,7-diyl)-co-(4,4΄-(N-(4-sec y-butylphenyl))diphenylamine)] (TFB) or Poly(N-vinylcarbazole) (PVK) as HTL. (d) Current and EQE efficiencies as a function of luminance of blue QD-LEDs based on PVK or TFB as HTL in devices. Adapted with permission from [30]. Figure 7. Open in new tabDownload slide (a) Device structure of multilayer QD-LED devices with Polymethyl methacrylate (PMMA) as a blocking layer. (b) Current density and luminance versus voltage characteristics for QLEDs without and with the 6-nm PMMA layer. Adapted with permission from [29]. (c) Schematic of blue QD-LED devices structure with poly[(9,9-dioctylfluorenyl-2,7-diyl)-co-(4,4΄-(N-(4-sec y-butylphenyl))diphenylamine)] (TFB) or Poly(N-vinylcarbazole) (PVK) as HTL. (d) Current and EQE efficiencies as a function of luminance of blue QD-LEDs based on PVK or TFB as HTL in devices. Adapted with permission from [30]. QDs in matrix QDs in organic matrix In the early stage of QD-LEDs, QDs normally contained a large number of ligands on its surface, partly due to the problem of a lack of strategies for efficient ligand exchange. In order to improve the conductivity of the film, they were normally dispersed in a conductive organic matrix that acted as a carrier-injection medium [49,132–135]. End group functionalization of conducting organic molecules was explored to enable direct linkage of them on QDs, allowing efficient carrier injection into QDs [49,133,136,137]. In addition, organic molecules grafted onto the surface of QDs can protect them from aggregation and enhance solubility in non-polar solvents, facilitating the fabrication of uniform and stable QD films [133]. To facilitate efficient carrier injection from polymers to QDs, polymers with large band gaps to form type-I band alignment with QDs are employed [138]. Poly(p-methyltriphenylamine) (Poly-TPA) [133], poly(vinyltriphenylamine dimer) (poly-TPD) [138] and their derivatives are common polymer matrices. CdSe/ZnS QDs and conducting polymer hybrid films by grafting blocking copolymers to them were investigated to enhance carrier transport (Fig. 8a and b). The hybrid films showed uniform QD distribution, and the device based on it exhibited relatively low turn-on voltage and a three-fold increase in EQE (Fig. 8c) [133]. More importantly, the stability of the film was dramatically improved when QDs were mixed with polymers (Fig. 8d). Figure 8. Open in new tabDownload slide (a) Schematic presentation of QDs in polymer matrix. (b) The energy-band diagram of QD-LEDs employing QD/Poly-TPA hybrid emissive layers. (c) EQE versus current density. (d) Test for colloidal stability of the QD/polymer hybrid under illumination in comparison with oleic acid and hexandecanethiol-coated QDs in chloroform. Adapted with permission from [133]. Figure 8. Open in new tabDownload slide (a) Schematic presentation of QDs in polymer matrix. (b) The energy-band diagram of QD-LEDs employing QD/Poly-TPA hybrid emissive layers. (c) EQE versus current density. (d) Test for colloidal stability of the QD/polymer hybrid under illumination in comparison with oleic acid and hexandecanethiol-coated QDs in chloroform. Adapted with permission from [133]. Side-chain conjugated polymers were used as matrix materials for CdSe/ZnS QD-based QD-LEDs. The devices exhibited the highest EQE of 6.09%. Moreover, the correlation between the highest occupied molecular orbit (HOMO) level of the polymer and device performance was identified. Polymers with low-lying HOMO levels provide a reduced barrier for hole injection, and therefore lead to improved device performance [138,139]. In conclusion, organic matrix could serve as an effective material for pursuing high-performance QD-LEDs. However, the efficiency of the LED with this structure is relatively low, and the thermal stability is another issue to be addressed. QDs in inorganic matrix Compared with organic matrix, inorganic matrix possesses better thermal stability and faster carrier-transporting properties. QDs such as PbS and CdSe embedded in a matrix of a wide-band-gap semiconductor material (e.g. CdS, ZnS) were developed. QDs were firstly overcoated with several monolayers of wide-band-gap semiconductor materials, yielding a type-I core–shell structure. The core–shell QDs capped with thermally degradable ligands are subsequently spin-coated onto a substrate, and heated to remove the organic ligands. Finally, the pores generated in the film are filled with an additional wide-band-gap semiconductor material through the successive ionic layer adsorption and reaction (SILAR) [51] methods [50,140], as shown in Fig. 9. Figure 9. Open in new tabDownload slide Flow chart showing the key steps involved in the development of semiconductor matrix-encapsulated nanocrystal arrays (SMENA). These stages include colloidal synthesis of PbS NCs (step 1), growth of the CdS shell (step 2) and spin-coating or dip-coating of NC films, exchange of bulky ligands with thermally degradable molecules (MPA, FA) and crystallographic fusion of NCs; all three were performed via layer-by-layer deposition (steps 3 and 4). In the last step, the pores of the PbS/CdS solid are filled with an additional CdS or ZnS material (step 5). Adapted with permission from [139]. Figure 9. Open in new tabDownload slide Flow chart showing the key steps involved in the development of semiconductor matrix-encapsulated nanocrystal arrays (SMENA). These stages include colloidal synthesis of PbS NCs (step 1), growth of the CdS shell (step 2) and spin-coating or dip-coating of NC films, exchange of bulky ligands with thermally degradable molecules (MPA, FA) and crystallographic fusion of NCs; all three were performed via layer-by-layer deposition (steps 3 and 4). In the last step, the pores of the PbS/CdS solid are filled with an additional CdS or ZnS material (step 5). Adapted with permission from [139]. Inorganic matrix was explored to improve QD surface passivation as well. A solution-phase assembly of PbS QDs into CdS inorganic matrix solids was investigated. As a result, the film exhibits bright IR emission and superior thermal stability [50]. Recently, Brovelli et al. developed a dot-in-bulk structure with thick inorganic CdS matrix that shows dual-peak photoluminescence (Fig. 10a and b). Dual-color electroluminescence can be realized as well by controlling the applied bias, which provides a convenient way to develop white light-emitting diodes [141]. Figure 10. Open in new tabDownload slide (a) Left: schematic illustration of CdSe/CdS dot-in-bulk structure and device architecture. Right: Dual-color EL from the dot-in-bulk LEDs. (b) Current density (J; black circles) and luminance (red open circles) versus applied voltage for the QD-LEDs made of CdSe/CdS dot-in-bulk structure. Inset: EQE for emission from the top face of the device as a function of current density. The inset also shows photographs of functioning LEDs biased with 10 V. Adapted with permission from [141]. Figure 10. Open in new tabDownload slide (a) Left: schematic illustration of CdSe/CdS dot-in-bulk structure and device architecture. Right: Dual-color EL from the dot-in-bulk LEDs. (b) Current density (J; black circles) and luminance (red open circles) versus applied voltage for the QD-LEDs made of CdSe/CdS dot-in-bulk structure. Inset: EQE for emission from the top face of the device as a function of current density. The inset also shows photographs of functioning LEDs biased with 10 V. Adapted with permission from [141]. In general, the use of inorganic matrix effectively improved the stability of the device. Relatively poor surface passivation and low PL QY of the film form the bottleneck that limited device performance. QDs in perovskite To address the problem of interface defects and poor carrier transport, a new matrix passivator with high carrier mobility is desirable. Borrowed from the typical heteroepitaxial growth method that is generally used for semiconductor devices, organohalide perovskite were epitaxialy grown on QDs in situ in the film. The film exhibited remarkable PLQY traceable to their atom-scale surface passivation via perovskite. Meanwhile, the excellent carrier-transporting capability of perovskite allows efficient carrier injection into QDs. For type-I QDs/perovskite composites, electrons and holes in the larger-band-gap perovskites were transferred with 80% efficiency to become excitons in the quantum-dot nanocrystals. It hence provides a new protocol to leverage perovskites’ excellent carrier diffusion to produce bright light emission from infrared QDs [48]. By combining the excellent carrier transport of the perovskite matrix and the high radiative efficiency of QDs in film, Gong et al. achieved a record power-conversion efficiency of 4.9% based on the PbS QDs in the MAPbIxBr3–x film (Fig. 11a and d). Owing to the smaller lattice mismatch, QD-in-perovskite composite (Fig. 11b) shows lower defect density at the interface between QD and perovskite, giving rise to enhanced PLQY and carrier-injection efficiency [142]. In addition to high device efficiency, the device shows band edge-derived turn-on voltage (Fig. 11c). QDs in a perovskite structure therefore provide an opportunity to realize both high PLQY and good carrier-transporting capability at the same time. Figure 11. Open in new tabDownload slide (a) The LED device architecture. (b) Illustration of enhanced electroluminescence efficiency in PbS QDs in MAPbX3 (X = I, Br) perovskite QD-LEDs; the left panel illustrates that radiative recombination dominates when QDs and perovskite are lattice matched, whereas lattice mismatch causes interfacial defects (black dashed line) and non-radiative recombination through traps (right panel). (c) Radiance–voltage characteristics. Inset: Zoom-in region of radiance–voltage plots (0.9–1.3 V) show the device turn-on voltages. (d) Perovskite devices (3.6% volume ratio) matrix versus a pure QD device. Adapted with permission from [142]. Figure 11. Open in new tabDownload slide (a) The LED device architecture. (b) Illustration of enhanced electroluminescence efficiency in PbS QDs in MAPbX3 (X = I, Br) perovskite QD-LEDs; the left panel illustrates that radiative recombination dominates when QDs and perovskite are lattice matched, whereas lattice mismatch causes interfacial defects (black dashed line) and non-radiative recombination through traps (right panel). (c) Radiance–voltage characteristics. Inset: Zoom-in region of radiance–voltage plots (0.9–1.3 V) show the device turn-on voltages. (d) Perovskite devices (3.6% volume ratio) matrix versus a pure QD device. Adapted with permission from [142]. CONCLUSION AND PERSPECTIVES In summary, as benefited from QD surface engineering and device-architecture optimization, the performance of QD-LEDs has been significantly improved, and this shows great promise for future commercialization. Surface engineering, such as the use of core–shell QD structures and thiol ligands, effectively reduced defect concentrations and improved PLQY. Device-architecture engineering including planar or matrix structure and the development of new carrier-transporting and blocking layers prompted radiative carrier recombination. The performance of QD-based LEDs such as light intensity and power-conversion efficiency have exceeded current commercial requirements in recent years, indicating the bright future of QD-LEDs for next-generation display applications. To meet commercial requirements, device stability needs to be much improved. When devices are operating, the non-radiative decay may cause device heating, while the injection of carriers may cause oxidation or reduction of QDs, both of which will cause ligand loss and defect formation, incurring aggravated luminescence efficiency and performance decay. The development of a pure inorganic matrix epitaxially growing on the QD surface is one potential solution. On the other hand, to improve device stability, the defect concentrations need to be reduced further; better solutions for ligand exchange, film fabrication and matrix growth are still highly desirable. Post treatment such as atomic layer deposition could be another potential strategy to improve the stability of the device. The development of heavy-metal-free QDs like InP or CuInS2 will be a solution to the toxicity issue. The core–shell QDs based on InP [36,128] and CuInS2 [7] have recently exhibited over 70% PLQY. Based on these advances, the performance of heavy-metal-free QD-LEDs has also rapidly improved. However, their broad emission peak limited their color quality. More effort needs to be made to improve PLQY, and reduce the emission peak width. For large-scale application, more attention needs to be paid to the development of scale-up film-fabrication processes. Transfer-printing (Fig. 12a) [33], inkjet-printing (Fig. 12b) [87] and patterning processes [143] could be ideal ways. In addition, large-scale synthesis using a reactor is an effective strategy to reduce the production cost. Figure 12. Open in new tabDownload slide Schematic representation of the (a) transfer-printing (adapted with permission from [33]) and (b) inkjet-printing QD-LED devices. Figure 12. Open in new tabDownload slide Schematic representation of the (a) transfer-printing (adapted with permission from [33]) and (b) inkjet-printing QD-LED devices. Acknowledgments We thank Hefei Liu at Shanghai Tech University for editing. FUNDING This work was supported by the start-up funding from Shanghai Tech University, the Young 1000 Talents Program, the National Key Research Program (2016YFA0204000), the National Natural Science Foundation of China (21571129), Joint Foundation for Big Instrument by National Natural Science Foundation and Chinese Academy of Sciences (U1632118) the Shanghai key research program (16JC1402100) and the Shanghai International Cooperation Project (16520720700). REFERENCES 1. Brus L . Electronic wave-functions in semiconductor clusters—experiment and theory . J Phys Chem 1986 ; 90 : 2555 – 60 . Google Scholar Crossref Search ADS WorldCat 2. Alivisatos AP . Semiconductor clusters, nanocrystals, and quantum dots . Science 1996 ; 271 : 933 – 7 . Google Scholar Crossref Search ADS WorldCat 3. Lim J , Bae WK, Kwak Jet al. Perspective on synthesis, device structures, and printing processes for quantum dot displays . Opt Mater Express 2012 ; 2 : 594 – 628 . Google Scholar Crossref Search ADS WorldCat 4. Shirasaki Y , Supran GJ, Bawendi MGet al. Emergence of colloidal quantum-dot light-emitting technologies . Nat Photonics 2013 ; 7 : 13 – 23 . Google Scholar Crossref Search ADS WorldCat 5. Protesescu L , Yakunin S, Bodnarchuk MIet al. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): novel optoelectronic materials showing bright emission with wide color gamut . Nano Lett 2015 ; 15 : 3692 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 6. Schmidt LC , Pertegas A, Gonzalez-Carrero Set al. Nontemplate synthesis of CH3NH3PbBr3 perovskite nanoparticles . J Am Chem Soc 2014 ; 136 : 850 – 3 . Google Scholar Crossref Search ADS PubMed WorldCat 7. Ning Z , Zhitomirsky D, Adinolfi Vet al. Graded doping for enhanced colloidal quantum dot photovoltaics . Adv Mater 2013 ; 25 : 1719 – 23 . Google Scholar Crossref Search ADS PubMed WorldCat 8. Ning Z , Dong H, Zhang Qet al. Solar cells based on inks of n-type colloidal quantum dots . ACS Nano 2014 ; 8 : 10321 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 9. Lan XZ , Voznyy O, Kiani Aet al. Passivation using molecular halides increases quantum dot solar cell performance . Adv Mater 2016 ; 28 : 299 – 304 . Google Scholar Crossref Search ADS PubMed WorldCat 10. Carey GH , Abdelhady AL, Ning ZJet al. Colloidal quantum dot solar cells . Chem Rev 2015 ; 115 : 12732 – 63 . Google Scholar Crossref Search ADS PubMed WorldCat 11. Wang RL , Shang YQ, Kanjanaboos Pet al. Colloidal quantum dot ligand engineering for high performance solar cells . Energ Environ Sci 2016 ; 9 : 1130 – 43 . Google Scholar Crossref Search ADS WorldCat 12. Kim GH , Garcia de Arquer FP, Yoon YJet al. High-efficiency colloidal quantum dot photovoltaics via robust self-assembled monolayers . Nano Lett 2015 ; 15 : 7691 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 13. Lan X , Voznyy O, Garcia de Arquer FPet al. 10.6% certified colloidal quantum dot solar cells via solvent-polarity-engineered halide passivation . Nano Lett 2016 ; 16 : 4630 – 4 . Google Scholar Crossref Search ADS PubMed WorldCat 14. Medintz IL , Uyeda HT, Goldman ERet al. Quantum dot bioconjugates for imaging, labelling and sensing . Nat Mater 2005 ; 4 : 435 – 46 . Google Scholar Crossref Search ADS PubMed WorldCat 15. Xu G , Zeng S, Zhang Bet al. New generation cadmium-free quantum dots for biophotonics and nanomedicine . Chem Rev 2016 ; 116 : 12234 – 327 . Google Scholar Crossref Search ADS PubMed WorldCat 16. Konstantatos G , Howard I, Fischer Aet al. Ultrasensitive solution-cast quantum dot photodetectors . Nature 2006 ; 442 : 180 – 3 . Google Scholar Crossref Search ADS PubMed WorldCat 17. Konstantatos G , Levina L, Tang Jet al. Sensitive solution-processed Bi2S3 nanocrystalline photodetectors . Nano Lett 2008 ; 8 : 4002 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 18. Lee JW , Kim do Y, Baek Set al. Photodetectors: inorganic UV-visible-SWIR broadband photodetector based on monodisperse PbS nanocrystals . Small 2016 ; 12 : 1246 . Google Scholar Crossref Search ADS WorldCat 19. Gao JB , Nguyen SC, Bronstein NDet al. Solution-processed, high-speed, and high-quantum-efficiency quantum dot infrared photodetectors . ACS Photonics 2016 ; 3 : 1217 – 22 . Google Scholar Crossref Search ADS WorldCat 20. Qiao K , Deng H, Yang Xet al. Spectra-selective PbS quantum dot infrared photodetectors . Nanoscale 2016 ; 8 : 7137 – 43 . Google Scholar Crossref Search ADS PubMed WorldCat 21. Koh WK , Saudari SR, Fafarman ATet al. Thiocyanate-capped PbS nanocubes: ambipolar transport enables quantum dot based circuits on a flexible substrate . Nano Lett 2011 ; 11 : 4764 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 22. Liu Y , Tolentino J, Gibbs Met al. PbSe quantum dot field-effect transistors with air-stable electron mobilities above 7 cm2 V−1 s−1 . Nano Lett 2013 ; 13 : 1578 – 87 . Google Scholar Crossref Search ADS PubMed WorldCat 23. Adinolfi V , Kramer IJ, Labelle AJet al. Photojunction field-effect transistor based on a colloidal quantum dot absorber channel layer . ACS Nano 2015 ; 9 : 356 – 62 . Google Scholar Crossref Search ADS PubMed WorldCat 24. Adachi MM , Fan F, Sellan DPet al. Microsecond-sustained lasing from colloidal quantum dot solids . Nat Commun 2015 ; 6 : 8694 . Google Scholar Crossref Search ADS PubMed WorldCat 25. Hoogland S , Sukhovatkin V, Howard Iet al. A solution-processed 1.53 mum quantum dot laser with temperature-invariant emission wavelength . Opt Express 2006 ; 14 : 3273 – 81 . Google Scholar Crossref Search ADS PubMed WorldCat 26. Caruge JM , Halpert JE, Wood Vet al. Colloidal quantum-dot light-emitting diodes with metal-oxide charge transport layers . Nat Photonics 2008 ; 2 : 247 – 50 . Google Scholar Crossref Search ADS WorldCat 27. Qian L , Zheng Y, Xue JGet al. Stable and efficient quantum-dot light-emitting diodes based on solution-processed multilayer structures . Nat Photonics 2011 ; 5 : 543 – 8 . Google Scholar Crossref Search ADS WorldCat 28. Mashford BS , Stevenson M, Popovic Zet al. High-efficiency quantum-dot light-emitting devices with enhanced charge injection . Nat Photonics 2013 ; 7 : 407 – 12 . Google Scholar Crossref Search ADS WorldCat 29. Dai X , Zhang Z, Jin Yet al. Solution-processed, high-performance light-emitting diodes based on quantum dots . Nature 2014 ; 515 : 96 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 30. Yang YX , Zheng Y, Cao WRet al. High-efficiency light-emitting devices based on quantum dots with tailored nanostructures . Nat Photonics 2015 ; 9 : 259 – 66 . Google Scholar Crossref Search ADS WorldCat 31. Yang Z , Voznyy O, Liu Met al. All-quantum-dot infrared light-emitting diodes . ACS Nano 2015 ; 9 : 12327 – 33 . Google Scholar Crossref Search ADS PubMed WorldCat 32. Sun Q , Wang YA, Li LSet al. Bright, multicoloured light-emitting diodes based on quantum dots . Nat Photonics 2007 ; 1 : 717 – 22 . Google Scholar Crossref Search ADS WorldCat 33. Kim TH , Cho KS, Lee EKet al. Full-colour quantum dot displays fabricated by transfer printing . Nat Photonics 2011 ; 5 : 176 – 82 . Google Scholar Crossref Search ADS WorldCat 34. Zhu R , Luo Z, Chen Het al. Realizing Rec 2020 color gamut with quantum dot displays . Opt Express 2015 ; 23 : 23680 – 93 . Google Scholar Crossref Search ADS PubMed WorldCat 35. Lee J , Sundar VC, Heine JRet al. Full color emission from II–VI semiconductor quantum dot-polymer composites . Adv Mater 2000 ; 12 : 1102–5 . Google Scholar OpenURL Placeholder Text WorldCat 36. Kim S , Kim T, Kang Met al. Highly luminescent InP/GaP/ZnS nanocrystals and their application to white light-emitting diodes . J Am Chem Soc 2012 ; 134 : 3804 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 37. Chen B , Zhong H, Wang Met al. Integration of CuInS2-based nanocrystals for high efficiency and high colour rendering white light-emitting diodes . Nanoscale 2013 ; 5 : 3514 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 38. Li XM , Wu Y, Zhang SLet al. CsPbX3 quantum dots for lighting and displays: room-temperature synthesis, photoluminescence superiorities, underlying origins and white light-emitting diodes . Adv Funct Mater 2016 ; 26 : 2435 – 45 . Google Scholar Crossref Search ADS WorldCat 39. Pietryga JM , Park YS, Lim JHet al. Spectroscopic and device aspects of nanocrystal quantum dots . Chem Rev 2016 ; 116 : 10513 – 622 . Google Scholar Crossref Search ADS PubMed WorldCat 40. Bai S , Yuan ZC, Gao F. Colloidal metal halide perovskite nanocrystals: synthesis, characterization, and applications . J Mater Chem C 2016 ; 4 : 3898 – 904 . Google Scholar Crossref Search ADS WorldCat 41. Song J , Li J, Li Xet al. Quantum dot light-emitting diodes based on inorganic perovskite cesium lead halides (CsPbX3) . Adv Mater 2015 ; 27 : 7162 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 42. Huang H , Zhao F, Liu Let al. Emulsion synthesis of size-tunable CH3NH3PbBr3 quantum dots: an alternative route toward efficient light-emitting diodes . ACS Appl Mater Inter 2015 ; 7 : 28128 – 33 . Google Scholar Crossref Search ADS WorldCat 43. Pan J , Quan LN, Zhao Yet al. Highly efficient perovskite-quantum-dot light-emitting diodes by surface engineering . Adv Mater 2016 ; 28 : 8718 – 25 . Google Scholar Crossref Search ADS PubMed WorldCat 44. Zhang F , Zhong HZ, Chen Cet al. Brightly luminescent and color-tunable colloidal CH3NH3PbX3 (X = Br, I, Cl) quantum dots: potential alternatives for display technology . ACS Nano 2015 ; 9 : 4533 – 42 . Google Scholar Crossref Search ADS PubMed WorldCat 45. Reiss P , Protiere M, Li L. Core/shell semiconductor nanocrystals . Small 2009 ; 5 : 154 – 68 . Google Scholar Crossref Search ADS PubMed WorldCat 46. Shen H , Cao W, Shewmon NTet al. High-efficiency, low turn-on voltage blue-violet quantum-dot-based light-emitting diodes . Nano Lett 2015 ; 15 : 1211 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 47. Yang Y , Qin HY, Jiang MWet al. Entropic ligands for nanocrystals: from unexpected solution properties to outstanding processability . Nano Lett 2016 ; 16 : 2133 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 48. Ning Z , Gong X, Comin Ret al. Quantum-dot-in-perovskite solids . Nature 2015 ; 523 : 324 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 49. Kwak J , Bae WK, Zorn Met al. Characterization of quantum dot/conducting polymer hybrid films and their application to light-emitting diodes . Adv Mater 2009 ; 21 : 5022 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 50. Moroz P , Liyanage G, Kholmicheva NNet al. Infrared emitting PbS nanocrystal solids through matrix encapsulation . Chem Mater 2014 ; 26 : 4256 – 64 . Google Scholar Crossref Search ADS WorldCat 51. Pathan HM , Lokhande CD. Deposition of metal chalcogenide thin films by successive ionic layer adsorption and reaction (SILAR) method . B Mater Sci 2004 ; 27 : 85 – 111 . Google Scholar Crossref Search ADS WorldCat 52. Bawendi MG , Steigerwald ML, Brus LE. The quantum-mechanics of larger semiconductor clusters (quantum dots) . Annu Rev Phys Chem 1990 ; 41 : 477 – 96 . Google Scholar Crossref Search ADS WorldCat 53. Murray CB , Norris DJ, Bawendi MG. Synthesis and characterization of nearly monodisperse Cde (E = S, Se, Te) semiconductor nanocrystallites . J Am Chem Soc 1993 ; 115 : 8706 – 15 . Google Scholar Crossref Search ADS WorldCat 54. Smith AM , Nie SM. Semiconductor nanocrystals: structure, properties, and band gap engineering . Accounts Chem Res 2010 ; 43 : 190 – 200 . Google Scholar Crossref Search ADS WorldCat 55. Klimov VI , Mikhailovsky AA, Xu Set al. Optical gain and stimulated emission in nanocrystal quantum dots . Science 2000 ; 290 : 314 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 56. Nan W , Niu Y, Qin Het al. Crystal structure control of zinc-blende CdSe/CdS core/shell nanocrystals: synthesis and structure-dependent optical properties . J Am Chem Soc 2012 ; 134 : 19685 – 93 . Google Scholar Crossref Search ADS PubMed WorldCat 57. Dabbousi BO , RodriguezViejo J, Mikulec FVet al. (CdSe)ZnS core-shell quantum dots: synthesis and characterization of a size series of highly luminescent nanocrystallites. J Phys Chem B 1997 ; 101 : 9463 – 75 . Crossref Search ADS 58. Qu LH , Peng ZA, Peng XG. Alternative routes toward high quality CdSe nanocrystals . Nano Lett 2001 ; 1 : 333 – 7 . Google Scholar Crossref Search ADS WorldCat 59. Bansal AK , Antolin F, Zhang Set al. Highly luminescent colloidal CdS quantum dots with efficient near infrared electroluminescence in light-emitting diodes . J Phys Chem C 2016 ; 120 : 1871 – 80 . Google Scholar Crossref Search ADS WorldCat 60. Moreels I , Justo Y, De Geyter Bet al. Size-tunable, bright, and stable PbS quantum dots: a surface chemistry study . ACS Nano 2011 ; 5 : 2004 – 12 . Google Scholar Crossref Search ADS PubMed WorldCat 61. Zhang J , Gao J, Church CPet al. PbSe quantum dot solar cells with more than 6% efficiency fabricated in ambient atmosphere . Nano Lett 2014 ; 14 : 6010 – 5 . Google Scholar Crossref Search ADS PubMed WorldCat 62. Protesescu L , Yakunin S, Bodnarchuk MIet al. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): novel optoelectronic materials showing bright emission with wide color gamut . Nano Lett 2015 ; 15 : 3692 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 63. Leng M , Chen Z, Yang Yet al. Lead-Free, Blue emitting bismuth halide perovskite quantum dots . Angew Chem Int Edit 2016 ; 55 : 15012 – 6 . Google Scholar Crossref Search ADS WorldCat 64. Protesescu L , Yakunin S, Bodnarchuk MIet al. Monodisperse formamidinium lead bromide nanocrystals with bright and stable green photoluminescence . J Am Chem Soc 2016 ; 138 : 14202 – 5 . Google Scholar Crossref Search ADS PubMed WorldCat 65. Jellicoe TC , Richter JM, Glass HFet al. Synthesis and optical properties of lead-free cesium tin halide perovskite nanocrystals . J Am Chem Soc 2016 ; 138 : 2941 – 4 . Google Scholar Crossref Search ADS PubMed WorldCat 66. Xu S , Ziegler J, Nann T. Rapid synthesis of highly luminescent InP and InP/ZnS nanocrystals . J Mater Chem 2008 ; 18 : 2653 – 6 . Google Scholar Crossref Search ADS WorldCat 67. Li L , Reiss P. One-pot synthesis of highly luminescent InP/ZnS nanocrystals without precursor injection . J Am Chem Soc 2008 ; 130 : 11588 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 68. Tessier MD , Dupont D, De Nolf Ket al. Economic and size-tunable synthesis of InP/ZnE (E = S, Se) colloidal quantum dots . Chem Mater 2015 ; 27 : 4893 – 8 . Google Scholar Crossref Search ADS WorldCat 69. Deng DW , Chen YQ, Cao Jet al. High-quality CuInS2/ZnS quantum dots for in vitro and in vivo bioimaging . Chem Mater 2012 ; 24 : 3029 – 37 . Google Scholar Crossref Search ADS WorldCat 70. Li L , Pandey A, Werder DJet al. Efficient synthesis of highly luminescent copper indium sulfide-based core/shell nanocrystals with surprisingly long-lived emission . J Am Chem Soc 2011 ; 133 : 1176 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 71. Zhong HZ , Zhou Y, Ye MFet al. Controlled synthesis and optical properties of colloidal ternary chalcogenide CuInS2 nanocrystals . Chem Mater 2008 ; 20 : 6434 – 43 . Google Scholar Crossref Search ADS WorldCat 72. Hamanaka Y , Ozawa K, Kuzuya T. Enhancement of donor-acceptor pair emissions in colloidal AgInS2 quantum dots with high concentrations of defects . J Phys Chem C 2014 ; 118 : 14562 – 8 . Google Scholar Crossref Search ADS WorldCat 73. Mao BD , Chuang CH, McCleese Cet al. Near-infrared emitting AgInS2/ZnS nanocrystals . J Phys Chem C 2014 ; 118 : 13883 – 9 . Google Scholar Crossref Search ADS WorldCat 74. Ogawa T , Kuzuya T, Hamanaka Yet al. Synthesis of Ag-In binary sulfide nanoparticles-structural tuning and their photoluminescence properties . J Mater Chem 2010 ; 20 : 2226 – 31 . Google Scholar Crossref Search ADS WorldCat 75. Srivastava BB , Jana S, Karan NSet al. Highly luminescent Mn-doped ZnS nanocrystals: gram-scale synthesis . J Phys Chem Lett 2010 ; 1 : 1454 – 8 . Google Scholar Crossref Search ADS WorldCat 76. Geszke-Moritz M , Clavier G, Lulek Jet al. Copper- or manganese-doped ZnS quantum dots as fluorescent probes for detecting folic acid in aqueous media . J Lumin 2012 ; 132 : 987 – 91 . Google Scholar Crossref Search ADS WorldCat 77. Yao L , Yu T, Ba LXet al. Efficient silicon quantum dots light emitting diodes with an inverted device structure . J Mater Chem C 2016 ; 4 : 673 – 7 . Google Scholar Crossref Search ADS WorldCat 78. Sugimoto H , Fujii M, Fukuda Yet al. All-inorganic water-dispersible silicon quantum dots: highly efficient near-infrared luminescence in a wide pH range . Nanoscale 2014 ; 6 : 122 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 79. Cheng XY , Lowe SB, Ciampi Set al. Versatile ‘click chemistry’ approach to functionalizing silicon quantum dots: applications toward fluorescent cellular imaging . Langmuir 2014 ; 30 : 5209 – 16 . Google Scholar Crossref Search ADS PubMed WorldCat 80. Wang YF , Hu AG. Carbon quantum dots: synthesis, properties and applications . J Mater Chem C 2014 ; 2 : 6921 – 39 . Google Scholar Crossref Search ADS WorldCat 81. Wu X , Tian F, Wang WXet al. Fabrication of highly fluorescent graphene quantum dots using L-glutamic acid for in vitro/in vivo imaging and sensing . J Mater Chem C 2013 ; 1 : 4676 – 84 . Google Scholar Crossref Search ADS WorldCat 82. Colvin VL , Schlamp MC, Alivisatos AP. Light-emitting-diodes made from cadmium selenide nanocrystals and a semiconducting polymer . Nature 1994 ; 370 : 354 – 7 . Google Scholar Crossref Search ADS WorldCat 83. Lim J , Jeong BG, Park Met al. Influence of shell thickness on the performance of light-emitting devices based on CdSe/Zn1-XCdXS core/shell heterostructured quantum dots . Adv Mater 2014 ; 26 : 8034–40 . Google Scholar PubMed OpenURL Placeholder Text WorldCat 84. Kwak J , Bae WK, Lee Det al. Bright and efficient full-color colloidal quantum dot light-emitting diodes using an inverted device structure . Nano Lett 2012 ; 12 : 2362 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 85. Shen HB , Cao WR, Shewmon NTet al. High-efficiency, low turn-on voltage blue-violet quantum-dot-based light-emitting diodes . Nano Lett 2015 ; 15 : 1211 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 86. Park JS , Kyhm J, Kim HHet al. Alternative patterning process for realization of large-area, full-color, active quantum dot display . Nano Lett 2016 ; 16 : 6946 – 53 . Google Scholar Crossref Search ADS PubMed WorldCat 87. Kim BH , Onses MS, Lim JBet al. High-resolution patterns of quantum dots formed by electrohydrodynamic jet printing for light-emitting diodes . Nano Lett 2015 ; 15 : 969 – 73 . Google Scholar Crossref Search ADS PubMed WorldCat 88. Matioli E , Weisbuch C. Direct measurement of internal quantum efficiency in light emitting diodes under electrical injection . J Appl Phys 2011 ; 109 : 3114 . Google Scholar Crossref Search ADS WorldCat 89. Forrest SR , Bradley DDC, Thompson ME. Measuring the efficiency of organic light-emitting devices . Adv Mater 2003 ; 15 : 1043 – 8 . Google Scholar Crossref Search ADS WorldCat 90. Kim S , Im SH, Kim SW. Performance of light-emitting-diode based on quantum dots . Nanoscale 2013 ; 5 : 5205 – 14 . Google Scholar Crossref Search ADS PubMed WorldCat 91. Hines MA , Guyot-Sionnest P. Synthesis and characterization of strongly luminescing ZnS-Capped CdSe nanocrystals . J Phys Chem-US 1996 ; 100 : 468 – 71 . Google Scholar Crossref Search ADS WorldCat 92. Bae WK , Padilha LA, Park YSet al. Controlled alloying of the core-shell interface in CdSe/CdS quantum dots for suppression of Auger recombination . ACS Nano 2013 ; 7 : 3411 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 93. Park YS , Bae WK, Baker Tet al. Effect of auger recombination on lasing in heterostructured quantum dots with engineered core/shell interfaces . Nano Lett 2015 ; 15 : 7319 – 28 . Google Scholar Crossref Search ADS PubMed WorldCat 94. Peng XG , Schlamp MC, Kadavanich AVet al. Epitaxial growth of highly luminescent CdSe/CdS core/shell nanocrystals with photostability and electronic accessibility . J Am Chem Soc 1997 ; 119 : 7019 – 29 . Google Scholar Crossref Search ADS WorldCat 95. Garcia-Santamaria F , Brovelli S, Viswanatha Ret al. Breakdown of volume scaling in Auger recombination in CdSe/CdS heteronanocrystals: the role of the core-shell interface . Nano Lett 2011 ; 11 : 687 – 93 . Google Scholar Crossref Search ADS PubMed WorldCat 96. Supran GJ , Song KW, Hwang GWet al. High-performance shortwave-infrared light-emitting devices using core-shell (PbS-CdS) colloidal quantum dots . Adv Mater 2015 ; 27 : 1437 – 42 . Google Scholar Crossref Search ADS PubMed WorldCat 97. Gong K , Kelley DF. A predictive model of shell morphology in CdSe/CdS core/shell quantum dots . J Chem Phys 2014 ; 141 : 194704 . Google Scholar Crossref Search ADS PubMed WorldCat 98. Talapin DV , Koeppe R, Gotzinger Set al. Highly emissive colloidal CdSe/CdS heterostructures of mixed dimensionality . Nano Lett 2003 ; 3 : 1677 – 81 . Google Scholar Crossref Search ADS WorldCat 99. Reiss P , Bleuse J, Pron A. Highly luminescent CdSe/ZnSe core/shell nanocrystals of low size dispersion . Nano Lett 2002 ; 2 : 781 – 4 . Google Scholar Crossref Search ADS WorldCat 100. Smith AM , Mohs AM, Nie S. Tuning the optical and electronic properties of colloidal nanocrystals by lattice strain . Nat Nanotechnol 2009 ; 4 : 56 – 63 . Google Scholar Crossref Search ADS PubMed WorldCat 101. Chen O , Zhao J, Chauhan VPet al. Compact high-quality CdSe-CdS core-shell nanocrystals with narrow emission linewidths and suppressed blinking . Nat Mater 2013 ; 12 : 445 – 51 . Google Scholar Crossref Search ADS PubMed WorldCat 102. Cui J , Beyler AP, Coropceanu Iet al. Evolution of the single-nanocrystal photoluminescence linewidth with size and shell: implications for exciton-phonon coupling and the optimization of spectral linewidths . Nano Lett 2016 ; 16 : 289 – 96 . Google Scholar Crossref Search ADS PubMed WorldCat 103. Bae WK , Char K, Hur Het al. Single-step synthesis of quantum dots with chemical composition gradients . Chem Mater 2008 ; 20 : 531 – 9 . Google Scholar Crossref Search ADS WorldCat 104. Boldt K , Kirkwood N, Beane GAet al. Synthesis of highly luminescent and photo-stable, graded shell CdSe/CdxZn1-xS nanoparticles by in situ alloying . Chem Mater 2013 ; 25 : 4731 – 8 . Google Scholar Crossref Search ADS WorldCat 105. Sonawane KG , Agarwal KS, Phadnis Cet al. Manifestations of varying grading level in CdSe/ZnSe core-shell nanocrystals . J Phys Chem C 2016 ; 120 : 5257 – 64 . Google Scholar Crossref Search ADS WorldCat 106. Roy D , Routh T, Asaithambi AVet al. Spectral and temporal optical behavior of blue-, green-, orange-, and red-emitting CdSe-based core/gradient alloy shell/shell quantum dots: ensemble and single-particle investigation results . J Phys Chem C 2016 ; 120 : 3483 – 91 . Google Scholar Crossref Search ADS WorldCat 107. Bae WK , Park YS, Lim Jet al. Controlling the influence of Auger recombination on the performance of quantum-dot light-emitting diodes . Nat Commun 2013 ; 4 : 2661 . Google Scholar Crossref Search ADS PubMed WorldCat 108. Boles MA , Ling D, Hyeon Tet al. The surface science of nanocrystals . Nat Mater 2016 ; 15 : 141 – 53 . Google Scholar Crossref Search ADS PubMed WorldCat 109. Liang XY , Yi Q, Bai Set al. Synthesis of unstable colloidal inorganic nanocrystals through the introduction of a protecting ligand . Nano Lett 2014 ; 14 : 3117 – 23 . Google Scholar Crossref Search ADS PubMed WorldCat 110. Dirin DN , Dreyfuss S, Bodnarchuk MIet al. Lead halide perovskites and other metal halide complexes as inorganic capping ligands for colloidal nanocrystals . J Am Chem Soc 2014 ; 136 : 6550 – 3 . Google Scholar Crossref Search ADS PubMed WorldCat 111. Tang J , Kemp KW, Hoogland Set al. Colloidal-quantum-dot photovoltaics using atomic-ligand passivation . Nat Mater 2011 ; 10 : 765 – 71 . Google Scholar Crossref Search ADS PubMed WorldCat 112. Owen J. Nanocrystal structure: the coordination chemistry of nanocrystal surfaces . Science 2015 ; 347 : 615 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 113. Pradhan N , Battaglia DM, Liu Yet al. Efficient, stable, small, and water-soluble doped ZnSe nanocrystal emitters as non-cadmium biomedical labels . Nano Lett 2007 ; 7 : 312 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 114. Tamang S , Beaune G, Texier Iet al. Aqueous phase transfer of InP/ZnS nanocrystals conserving fluorescence and high colloidal stability . ACS Nano 2011 ; 5 : 9392 – 402 . Google Scholar Crossref Search ADS PubMed WorldCat 115. Ding TX , Olshansky JH, Leone SRet al. Efficiency of hole transfer from photoexcited quantum dots to covalently linked molecular species . J Am Chem Soc 2015 ; 137 : 2021 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 116. Sun L , Choi JJ, Stachnik Det al. Bright infrared quantum-dot light-emitting diodes through inter-dot spacing control . Nat Nanotechnol 2012 ; 7 : 369 – 73 . Google Scholar Crossref Search ADS PubMed WorldCat 117. Yang Y , Qin HY, Peng XG. Intramolecular entropy and size-dependent solution properties of nanocrystal-ligands complexes . Nano Lett 2016 ; 16 : 2127 – 32 . Google Scholar Crossref Search ADS PubMed WorldCat 118. Lee KH , Lee JH, Song WSet al. Highly efficient, color-pure, color-stable blue quantum dot light-emitting devices . ACS Nano 2013 ; 7 : 7295 – 302 . Google Scholar Crossref Search ADS PubMed WorldCat 119. Anikeeva PO , Madigan CF, Halpert JEet al. Electronic and excitonic processes in light-emitting devices based on organic materials and colloidal quantum dots . Phys Rev B 2008 ; 78 : 5434 . Google Scholar Crossref Search ADS WorldCat 120. Cho KS , Lee EK, Joo WJet al. High-performance crosslinked colloidal quantum-dot light-emitting diodes . Nat Photonics 2009 ; 3 : 341 – 5 . Google Scholar Crossref Search ADS WorldCat 121. Jang I , Kim J, Ippen Cet al. Inverted InP quantum dot light-emitting diodes using low-temperature solution-processed metal-oxide as an electron transport . Jpn J Appl Phys 2015 ; 54 : BC01 . Google Scholar Crossref Search ADS WorldCat 122. Liu GH , Zhou X, Sun XWet al. Performance of inverted quantum dot light-emitting diodes enhanced by using phosphorescent molecules as exciton harvesters . J Phys Chem C 2016 ; 120 : 4667 – 72 . Google Scholar Crossref Search ADS WorldCat 123. Kim HM , Kim J, Lee Jet al. Inverted quantum-dot light emitting diode using solution processed p-type WOx doped PEDOT:PSS and Li doped ZnO charge generation layer . ACS Appl Mater Inter 2015 ; 7 : 24592 – 600 . Google Scholar Crossref Search ADS WorldCat 124. Liu P , Wang H, Chen Jet al. Rapid and high-efficiency laser-alloying formation of ZnMgO nanocrystals . Sci Rep 2016 ; 6 : 28131 . Google Scholar Crossref Search ADS PubMed WorldCat 125. Liang XY , Ren YP, Bai Set al. Colloidal indium-doped zinc oxide nanocrystals with tunable work function: rational synthesis and optoelectronic applications . Chem Mater 2014 ; 26 : 5169 – 78 . Google Scholar Crossref Search ADS WorldCat 126. Kim JH , Han CY, Lee KHet al. Performance improvement of quantum dot-light-emitting diodes enabled by an alloyed ZnMgO nanoparticle electron transport layer . Chem Mater 2015 ; 27 : 197 – 204 . Google Scholar Crossref Search ADS WorldCat 127. Kim HH , Park S, Yi Yet al. Inverted quantum dot light emitting diodes using polyethylenimine ethoxylated modified ZnO . Sci Rep 2015 ; 5 : 8968 . Google Scholar Crossref Search ADS PubMed WorldCat 128. Lim J , Park M, Bae WKet al. Highly efficient cadmium-free quantum dot light-emitting diodes enabled by the direct formation of excitons within InP@ZnSeS quantum dots . ACS Nano 2013 ; 7 : 9019 – 26 . Google Scholar Crossref Search ADS PubMed WorldCat 129. Peng HR , Wang WG, Chen SM. Efficient quantum-dot light-emitting diodes with 4,4,4-Tris(N-Carbazolyl)-triphenylamine (TcTa) electron-blocking layer . IEEE Electr Device L 2015 ; 36 : 369 – 71 . Google Scholar Crossref Search ADS WorldCat 130. Sun QJ , Subramanyam G, Dai LMet al. Highly efficient quantum-dot light-emitting diodes with DNA-CTMA as a combined hole-transporting and electron-blocking layer . ACS Nano 2009 ; 3 : 737 – 43 . Google Scholar Crossref Search ADS PubMed WorldCat 131. Ji WY , Lv Y, Jing PTet al. Highly efficient and low turn-on voltage quantum dot light-emitting diodes by using a stepwise hole-transport layer . ACS Appl Mater Inter 2015 ; 7 : 15955 – 60 . Google Scholar Crossref Search ADS WorldCat 132. Tekin E , Smith PJ, Hoeppener Set al. Inkjet printing of luminescent CdTe nanocrystal-polymer composites . Adv Funct Mater 2007 ; 17 : 23 – 8 . Google Scholar Crossref Search ADS WorldCat 133. Zorn M , Bae WK, Kwak Jet al. Quantum dot-block copolymer hybrids with improved properties and their application to quantum dot light-emitting devices . ACS Nano 2009 ; 3 : 1063 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 134. Gao J , Nguyen SC, Bronstein NDet al. Solution-processed, high-speed, and high-quantum-efficiency quantum dot infrared photodetectors . ACS Photonics 2016 ; 3 : 1217 – 22 . Google Scholar Crossref Search ADS WorldCat 135. Lee S , Lee B, Kim BJet al. Free-standing nanocomposite multilayers with various length scales, adjustable internal structures, and functionalities . J Am Chem Soc 2009 ; 131 : 2579 – 87 . Google Scholar Crossref Search ADS PubMed WorldCat 136. Bae WK , Lim J, Zorn Met al. Reduced efficiency roll-off in light-emitting diodes enabled by quantum dot-conducting polymer nanohybrids . J Mater Chem C 2014 ; 2 : 4974 – 9 . Google Scholar Crossref Search ADS WorldCat 137. Bansal AK , Sajjad MT, Antolini Fet al. In situ formation and photo patterning of emissive quantum dots in small organic molecules . Nanoscale 2015 ; 7 : 11163 – 72 . Google Scholar Crossref Search ADS PubMed WorldCat 138. zur Borg L , Lee D, Lim Jet al. The effect of band gap alignment on the hole transport from semiconducting block copolymers to quantum dots . J Mater Chem C 2013 ; 1 : 1722 – 6 . Google Scholar Crossref Search ADS WorldCat 139. Fokina A , Lee Y, Chang JHet al. Side-chain conjugated polymers for use in the active layers of hybrid semiconducting polymer/quantum dot light emitting diodes . Polym Chem-UK 2016 ; 7 : 101 – 12 . Google Scholar Crossref Search ADS WorldCat 140. Kinder E , Moroz P, Diederich Get al. Fabrication of all-inorganic nanocrystal solids through matrix encapsulation of nanocrystal arrays . J Am Chem Soc 2011 ; 133 : 20488 – 99 . Google Scholar Crossref Search ADS PubMed WorldCat 141. Brovelli S , Bae WK, Galland Cet al. Dual-color electroluminescence from dot-in-bulk nanocrystals . Nano Lett 2014 ; 14 : 486 – 94 . Google Scholar Crossref Search ADS PubMed WorldCat 142. Gong XW , Yang ZY, Walters Get al. Highly efficient quantum dot near-infrared light-emitting diodes . Nat Photonics 2016 ; 10 : 253–7 . Google Scholar Crossref Search ADS WorldCat 143. Park JS , Kyhm JH, Kim Het al. An alternative patterning process for realization of large-area, full-color, active quantum dot display . Nano Lett 2016 ; 16 : 6946 – 53 . Google Scholar Crossref Search ADS PubMed WorldCat © The Author(s) 2017. Published by Oxford University Press on behalf of China Science Publishing & Media Ltd. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/), which permits non-commercial reuse, distribution, and reproduction in any medium, provided the original work is properly cited. for commercial re-use, please contact journals.permissions@oup.com © The Author(s) 2017. Published by Oxford University Press on behalf of China Science Publishing & Media Ltd. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com TI - Colloidal quantum-dots surface and device structure engineering for high-performance light-emitting diodes JF - National Science Review DO - 10.1093/nsr/nww097 DA - 2017-03-01 UR - https://www.deepdyve.com/lp/oxford-university-press/colloidal-quantum-dots-surface-and-device-structure-engineering-for-kvzq22jpII SP - 170 EP - 183 VL - 4 IS - 2 DP - DeepDyve ER -