Yan
Yan
*ab,
Hong-Yu
Liu
c,
Le
Bian
ab,
Yan-Yan
Dai
c,
Bo-Ning
Zhang
c,
Shuang-Mei
Xue
ab,
Ye
Zhou
a,
Jian-Long
Xu
*c and
Sui-Dong
Wang
c
aState Key Laboratory of Radio Frequency Heterogeneous Integration (Shenzhen University), Shenzhen, Guangdong 518060, P. R. China. E-mail: yanyan@szu.edu.cn
bCollege of Electronics and Information Engineering, Shenzhen University, Shenzhen, Guangdong 518060, P. R. China
cInstitute of Functional Nano & Soft Materials, Jiangsu Key Laboratory for Carbon-Based Functional Materials & Devices, Soochow University, Suzhou, Jiangsu 215123, P. R. China. E-mail: xujianlong@suda.edu.cn
First published on 1st November 2024
The development of cost-effective and highly sensitive short-wave infrared (SWIR) photodetectors is crucial for the expanding applications of SWIR imaging in civilian applications such as machine vision, autonomous driving, and augmented reality. Colloidal quantum dots (CQDs) have emerged as promising candidates for this purpose, offering distinct advantages over traditional III–V binary and ternary semiconductors. These advantages include the ability to precisely tune the bandgap through size modulation of CQDs and the ease of monolithic integration with Si readout integrated circuits (ROICs) via solution processing. Achieving a minimal reverse bias dark current density (Jd) while maintaining high external quantum efficiency is essential for enhancing the light detection sensitivity of CQDs-based SWIR photodiodes to a level competitive with III–V semiconductors. This challenge has garnered increasing research attention in recent years. Herein, the latest advancements in understanding and mitigating Jd in CQDs SWIR photodiodes are summarized. Starting with a brief overview of the material fundamentals of CQDs, the origins of Jd in CQDs photodiodes, including reverse injection from electrode, diffusion/drift currents, Shockley–Read–Hall generation/recombination currents, trap-assisted tunneling, and shunt/leakage currents, are discussed together with their latest research progresses about strategies adopted to suppress Jd. Finally, a brief conclusion and outlook on future research directions aimed at minimizing Jd and retaining high photoresponse of CQDs SWIR photodiodes are provided.
Wider impactSolution-processed thin-film photodetectors have garnered significant attention due to their capability of monolithical integration with silicon-based complementary metal–oxide–semiconductor readout circuits. This integration offers the potential to overcome the cost and resolution constraints inherent in compound semiconductor infrared focal plane arrays. Notably, colloidal quantum dots (CQDs) have emerged as promising candidates for this purpose, offering distinct advantages over traditional III–V binary and ternary semiconductors. Consequently, significant advancements have been achieved in the development of CQDs-based SWIR photodiodes and state-of-the-art SWIR imagers, paving the way for future advancements in this field. Dark current of a photodiode is defined as any current generated under an applied reverse bias voltage in the absence of light. For most applications, minimizing the CQDs SWIR dark current density (Jd) is crucial to improve key figures of merits. The topic of Jd of CQDs SWIR photodiodes have gained increasing attention in recent years, but the reports and results are scattered in the literature. It is the purpose of this paper to provide an overview to summarize the existing articles on this topic, and to discuss its current status and future research directions about how to minimize Jd and retaining high photoresponse of CQDs SWIR photodiodes. |
Fig. 1 (a) Images acquired in different scenes/applications for comparison with SWIR and visible imaging. Reproduced from ref. 16 with permission from [Institute of Electrical and Electronics Engineers], copyright [2022]. (b) Hybrid infrared imagers fabricated by a flip-bonding-based process. Reproduced from ref. 17 with permission from [Versita], copyright [2012]. (c) Monolithic integration fabrication process of CQDs with CMOS ROICs. Reproduced from ref. 18 with permission from [American Chemical Society], copyright [2023]. (d) Transmission electron microscopy (TEM) image of a fabricated CQDs SWIR photodiode stack on CMOS ROICs and schematic diagram of the CQDs photodiode stack integration on a ROIC. Reproduced from ref. 19 with permission from [Institute of Electrical and Electronics Engineers], copyright [2021]. |
To extend the array-format of Si CMOS imager, the most efficient strategy is to monolithically integrate SWIR sensing materials with Si ROICs. The first strategy is to epitaxially grow SWIR sensitive materials onto Si, which mainly includes germanium silicon (GeSi) and germanium tin (GeSn) alloy.27,28 By fine-tuning the GeSn alloy composition, the detection range can be well modulated and even extend to 3.65 μm, which can fully cover the entire 1–3 μm range of SWIR region.29 Moreover, the GeSn growth process on Si substrates is highly compatible with standard CMOS technology, offering prospects for large-scale and cost-effective manufacturing. This is especially important for constructing low-cost and high-performance SWIR images toward wide applications in civilian fields. However, though progress has been made, GeSn photodetectors still suffer from several critical challenges, including large lattice mismatch (exceeding 4.2%) between Si and GeSn, phase separation and material inhomogeneity at high Sn concentrations, and relatively poor stability of Sn at high temperatures.30 The heterogeneous epitaxial growth of SWIR sensing materials on the complex surface conditions of ROICs, including both oxides and metals, still hinders the achievement of high-performance SWIR photo-detectors and imagers. Another feasible and effective strategy is the monolithic integration of SWIR sensing materials with Si ROICs to construct thin-film photodetectors and image sensors. In thin-film photodetectors, all functional layers including the active layer, carrier transport layers, carrier blocking layers and electrodes can be directly fabricated into a pixel stack on top of ROICs with a high or even-unity fill factor. In this process, no epitaxial growth or flip-chip bonding procedure is required, making it possible to achieve low-cost, high-resolution and high-performance SWIR photodetectors. They are getting significant attention as a promising SWIR imaging platform. Among the promising candidates for converting low-energy photons to electric charge carriers, colloidal quantum dots (CQDs) are particularly well suited and have been considered a promising and low-cost SWIR sensitive semiconductor material owing to high potential for pixel size scaling down, high manufacturing throughput and cost-effectiveness.
Typical SWIR sensitive CQDs include PbX (X = S, Se and Te) CQDs, Hg-based CQDs (e.g., HgTe), and environmetal-friendly CQDs (e.g., Ag2Te).13,31–33 Owing to small Bohr radius and strong quantum confinement effects, the bandgap of PbX CQDs could be precisely modulated by varying the size of CQDs, and the absorption spectrum can fully cover the entire SWIR region, which provides a fundamental foundation for implementation of CQDs SWIR photodetectors.34,35 Similar to doping effects in bulk semiconductors, the semiconducting characteristics including n-/p-type doping, doping levels, carrier mobility, carrier concentration and energy band structure can also be precisely modulated by surface chemistry, making it possible to construct various photodetector structures with optimized photodetection performances. Indeed, PbX CQDs can be mass-produced via solution-phase synthesis procedures and then dispersed into highly stable inks, followed by a solution process on top of ROICs (Fig. 1c).18 After the thickness reached the desired value, the wafer was diced into individual imager chips, which required no hybrid integration process at the chip level by flip-chip bonding. The monolithically integrated pixel, including the photodiode and readout circuit, is shown in Fig. 1d, where the PbS CQDs photodiode is directly fabricated onto the metal pad of the ROICs.19 Monolithic processing of the CQDs pixels allows us to significantly scale down the pixel size of CQDs imagers compared with flip-chip integrated image sensors, as exemplified by the recent demonstration of a pixel pitch down to 1.62 μm for a 12-inch SWIR imager wafer.36
With extensive efforts in the last decade, significant progress has been made in the photodetection performance of CQDs SWIR photodetectors, benefiting from innovations in CQDs synthesis, defect passivation, device structure design, and pixel optimization. The CQDs SWIR photodetectors stand out among various emerging thin-film SWIR photodetection technologies owing to their superior SWIR photodetection performances and ease of wafer-scale uniform monolithical integration with Si-based ROICs. Fig. 2 shows the scatter plots of photoresponsivity (R) versus dark current density (Jd) and specific detectivity (D*) versus wavelength including CQDs SWIR photodiodes13,37–57 and other typical material-based SWIR photodiodes such as traditional bulk III–V binary and ternary semiconductors, two-dimensional transition metal dichalcogenides (2D TMDs),58–62 conjugated polymers63–68 and nanowires.69–71 From the R versus Jd scatter plot shown in Fig. 2a, it can be clearly seen that the lowest Jd levels of CQDs SWIR photodiodes reported to date are comparable to commercial InGaAs SWIR photodiodes and outperform other emerging technologies. Moreover, the D* of CQDs SWIR photodiodes reported to data are comparable or superior to other typical SWIR photodetectors (Fig. 2b), offering great potential for further practical applications. Herein, it should be especially noted that D* values in Fig. 2b are assessed by either using dark current as only source of noise current or adopting measured noise current at different frequencies, and thus such comparison is only for a preliminary reference.
Fig. 2 (a) R versus Jd and D* versus wavelength scatter plots of CQDs SWIR photodiodes and other typical material-based SWIR photodiodes including traditional bulk III–V binary and ternary semiconductors, 2D TMDs, conjugated polymers and nanowires. Here, the photodetection parameters of CQDs, 2D TMDs, conjugated polymers and nanowires based SWIR photodiodes are extracted from previous literatures,13,37–71 while those of III–V ternary InGaAs photodiodes are from the commercial product of Hamamatsu Photonics including G12180-003A, G12181-003A, G12182-003A and G12183-003A. |
For PbS CQDs photodiodes, two designs have been implemented based on PbS CQDs/metal oxide heterojunctions and PbS CQDs homojunctions. The built-in field at both the heterojunction and homojunction interfaces serves to separate photo-generated carriers and transport them to the electrodes, as illustrated in the band diagram in Fig. 3a.19 The heterojunction device usually shows relatively low dark current with a typical range of 10−8–10−6 A cm−2 at a −1 V bias, however, is limited by low photoresponse in the SWIR region.41,49,72–81 Inspired by CQDs solar cells, photodetectors based on CQDs homojunctions show much higher photoresponse due to their more suitable energy band alignment and interfacial characteristics. However, the homojunction device typically exhibits high dark current with a typical range of 10−6–10−4 A cm−2 at −1 V bias, hindering its further practical applications.19,37,39,50,52–54,56,82,83 Therefore, the main challenge for CQDs SWIR photodiodes is how to simultaneously achieve low dark current and high photoresponse, which is mainly limited by the presence of numerous defects in CQDs and the limited carrier mobility arising from carrier hopping between CQDs. Consequently, optimizing the photodetection characteristics of CQDs SWIR detectors is a current research priority.54
Fig. 3 (a) Illustrated band diagrams of heterojunction and homojunction PbS CQDs photodiodes under a reverse bias and SWIR light illumination. Reproduced from ref. 19 with permission from [Institute of Electrical and Electronics Engineers], copyright [2021]. (b) Typical CQDs photodiode current density versus voltage (J–V) characteristics in a semilogarithmic plot with and without illumination, showing Jd under a given bias voltage. |
The topic of Jd of CQDs SWIR photodiodes has gained increasing attention in recent years, but the reports and results are scattered in the literature. It is the purpose of this paper to provide an overview to summarize the existing articles on this topic, and to discuss its current status and future research directions. Exploration of the origin of Jd and development of specific strategies to reduce Jd form the bulk of this review paper.
(1) |
Fig. 4 Size-tunable PbS CQDs. (a) Quantum size effects in PbS CQDs allowing the absorption spectra extending 2.0 μm, outperforming existing Si and InGaAs technologies. Reproduced from ref. 86 with permission from [Institute of Electrical and Electronics Engineers], copyright [2021]. (b) Relationship between PbS CQDs bandgap and size including theoretical calculation and experimental results from different groups. Reproduced from ref. 87 with permission from [American Chemical Society], copyright [2009]. (c) TEM images of ultra-large PbS CQDs with an absorption peak at 2460 nm. Reproduced from ref. 88 with permission from [American Chemical Society], copyright [2019]. (d) Absorption spectra of PbS CQDs with an absorption peak ranging from 1100 to 2600 nm. Reproduced from ref. 88 with permission from [American Chemical Society], copyright [2019]. |
It can be clearly seen that Eg decreases with the increase of d, and thus large-size PbS CQDs should be synthesized toward SWIR detection applications. This also allows us to determine the diameter directly from Eg, avoiding a lengthly TEM analysis for each synthesized sample. The absorption peak of PbS CQDs has been extended to be over 2500 nm, which can almost cover the entire SWIR region. Dong et al. reported ultra-large PbS CQDs with an average diameter of 15.3 nm and a size dispersity lower than 10%, extending the first exciton absorption peak to 2500 nm, as shown in Fig. 4c and d.88 The resulted broadband heterojunction photodetectors with an external quantum efficiency (EQE) of 25% at 2100 nm were obtained.
Except for PbX CQDs, mercury telluride (HgTe) CQDs have also emerged as a compelling low-cost candidate for SWIR photodetectors due to its narrow energy transition capabilities and versatile spectral tunability.89 Different from HgS and HgSe CQDs showing intraband transition, HgTe CQDs usually exhibit interband electronic transitions upon optical excitation, since their Fermi levels are within the bandgap. Their first band edge optical features can span an incredibly wide range from the visible to the THz range, which directly stems from their semimetallic bulk nature.90–92 Moreover, surface chemistry modification for HgTe CQDs has been well developed, and this maturity in surface chemistry allows for precise control and manipulation of electrical properties of CQDs including CQDs coupling, doping level and charge transport behaviors.93,94 Significantly, similar to PbX CQDs, the energy levels of semiconducting HgTe CQDs can be well modulated by ligand engineering or doping, which also provides a platform for designing different device structures of photodiodes and phototransistors. Furthermore, the high air stability of HgTe CQDs enables material processing in air ambient while retaining their photoconductive properties.95 Benefiting from extensive studies on synthesis methodology and energy level alignments optimization in recent years,96,97 dramatic improvements in HgTe-based SWIR and MWIR photodetectors have been achieved.
However, heavy metals of Pb and Hg are Restriction of Hazardous Substance (ROHS) regulated elements, which impose major regulatory concerns and impede their employment in consumer electronics. In 2024, a novel synthesis protocol of heavy-metal free silver telluride (Ag2Te) CQDs was developed and successfully applied in high-performance SWIR photodiodes, as shown in Fig. 5.98 In the synthesis protocol, silver-oleylamine and tellurium-thiol complexes were used as Ag and Te precursors for hot-injection. The absence of phosphine results in high stability of Ag2Te CQDs in crude solution for hours under heating without showing Ostwald ripening, enabling fine modulation of CQDs size by controlling the precursor supply. The resulted AgTe CQDs photodetector exhibits a spectral range of 350–1600 nm with a specific detectivity over 1012 Jones, −3 dB bandwidth over 0.1 MHz and LDR over 118 dB.
Fig. 5 Size-tunable Ag2Te CQDs. Reproduced from ref. 98 with permission from [Springer Nature], copyright [2024]. (a) Absorption spectra of Ag2Te CQDs with various sizes. (b)–(k) TEM images and their size distribution histograms of Ag2Te CQDs with an absorption peak at 1310, 1430, 1520, 1750 and 1940 nm. Scale bar: 20 nm. |
Fig. 6 Synthesis protocol of PbS CQDs using a hot-injection strategy. (a) Scheme of a hot injection reaction apparatus setup for synthesis of PbS CQDs. Reproduced from ref. 103 with permission from [American Institute of Physics], copyright [2023]. (b) Change of PbS precursor with the reaction time during a hot injection reaction, showing the nucleation and growth stage of monodisperse PbS CQDs according to the LaMer model. Reproduced from ref. 103 with permission from [American Institute of Physics], copyright [2023]. (c) Relationship between first exciton peak wavelength of PbS CQDs and the reaction time of period under various reaction temperatures. Reproduced from ref. 104 with permission from [Royal Society of Chemistry], copyright [2017]. (d) Illustration of the two-step synthesis process for ultra-large PbS CQDs. Reproduced from ref. 88 with permission from [American Chemical Society], copyright [2019]. (e) First exciton peak wavelength and added sulfur precursor versus times of injection in the above two-step synthesis procedure of ultra-large PbS CQDs. Reproduced from ref. 88 with permission from [American Chemical Society], copyright [2019]. |
How to obtain high-quality large-size PbX CQDs is crucial for SWIR photodetectors. Commonly, the CQDs diameter can be tunable by adjusting the injection temperature and reaction time in the CQDs synthesis process. The first exciton peaks for PbS CQDs synthesized at different temperatures and growth time are depicted in Fig. 6c.104 Increasing the injection temperature would decrease the supersaturation, and thus create fewer nuclei and result in larger size of CQDs. Extending the growth time can also enlarge the size of CQDs. Even altering the ligand type is also effective in tuning nucleation rates and thus size of PbS CQDs.103 Besides, Dong et al. reported a feasible strategy to grow ultra-large PbS CQDs with an absorption peak at 2500 nm by multiple injections, using PbS seed crystals with an absorption peak at 1960 nm (Fig. 6d).88 The peak wavelength of PbS CQDs increases with the time of injection, where the required sulfur source for each injection increases with the number of injections to maintain the growth of CQDs (Fig. 6e). The authors pointed out that the size of initial seed crystals, time of injection and amount of sulfur source in each injection were of great importance to determine whether uniform large-size PbS CQDs could be successfully synthesized with excellent charge transport and optoelectronic properties.
Fig. 7 Ligand exchange in PbS CQDs. (a) Schematic diagram of conventional two-step ligand exchange process including solid-state ligand exchange and solution-phase ligand exchange processes. Reproduced from ref. 111 with permission from [Springer Nature], copyright [2019]. (b) Schematic diagram of newly developed one-step direct synthesis of iodide capped PbS CQDs inks. Reproduced from ref. 111 with permission from [Springer Nature], copyright [2019]. (c) LaMer model for the one-step direct synthesized PbS CQDs inks, showing the effect of precursor conversion kinetics on the size of final CQDs. Reproduced from ref. 113 with permission from [John Wiley and Sons], copyright [2023]. (d) Absorption of one-step direct synthesized PbS CQDs inks under various reaction concentrations and temperatures. Reproduced from ref. 113 with permission from [John Wiley and Sons], copyright [2023]. (e) Chemical structures of the short-chain ligands commonly adopted for exchange with OA in PbS CQDs. Reproduced from ref. 114 with permission from [American Chemical Society], copyright [2014]. (f) Energy level diagrams of PbS CQDs exchanged with different short-chain ligands shown in (e). Reproduced from ref. 114 with permission from [American Chemical Society], copyright [2014]. (g) Effects of short-chain ligand and treatment atmosphere on the doping levels of PbS CQDs. Reproduced from ref. 115 with permission from [American Chemical Society], copyright [2012]. |
To overcome limitations of the SSLE process, the solution-phase ligand exchange process was developed and have been widely applied in PbX CQDs SWIR optoelectronic devices. In such a process, the ligand was exchanged in a solution phase to obtain high-concentration CQDs ink, followed by spin-coating one time to achieve the desired thickness film (Fig. 7a).111 Compared with SSLE method, the homogeneous reaction in the solution-phase results in more thorough and uniform ligand exchange on the surface of CQDs, and also the ink solvent component and concentration is easier to manipulate. Nowadays, PbX CQDs photodetectors with highest photoresponsivity values are almost based on solution-phase ligand exchange process. However, the complex ligand exchange process usually results in the defect regeneration problems, which significantly hinders the photodetection performance enhancement. At the same time, such ligand exchange process greatly limits the synthesis yield in one batch (usually in the milligram range), making it difficult for scalable industrial production. More importantly, the CQDs ink can remain stable within several minutes, which is difficult to meet requirements of practical applications. Therefore, it is necessary to develop stable semiconducting PbX CQDs inks with desired ligand type and concentration on CQDs surface and also low defect concentration.
In 2019, Wang et al. developed a one-step synthesis strategy to directly prepare I-passivated PbS CQDs by combining CQDs synthesis and ligand exchange (Fig. 7b),111 which can significantly simplify the preparation of CQDs films. In the synthesis process, metal halide (PbI2) and N,N-Diphenylthiourea (DPhTA) were used as lead and sulfide source, respectively. PbI2 and DPhTA were dissolved in dimethyl formamide (DMF) to obtain the precursor, and butyl amine (BA) was injected into the solution to directly synthesize I-passivated PbS CQDs. This method enables the in situ iodide passivation of PbS CQDs and the fabrication of thick films without ligand exchange, which significantly reduces the processing steps and the costs from more than 16 dollars g−1 to lower than 6 dollars g−1. Based on these one-step synthesized halide capped PbS (PbS-I) CQDs, photoconductor-type devices show a high detectivity of 1.4 × 1011 Jones in the near-infrared region. Later in 2023, they revealed the detailed reaction mechanisms of one-step direct synthesis PbS CQDs (Fig. 7c) and realized larger size tunability of PbS CQDs inks with absorption region extending to the SWIR region (Fig. 7d).113 The photodiode-type photodetector utilizing the SWIR PbS inks combine a low Jd of 2 × 10−6 A cm−2 at −0.8 V bias and a high external quantum efficiency (EQE) of 70% at 1300 nm, compared to or superior to previously reported state-of-the-art PbS CQDs SWIR photodetectors using SSLE or solution-phase ligand exchange process.
Except for enhancing carrier transport, another important role of ligand exchange is to modulate the band structure and doping levels of CQDs films since the electronic properties of CQDs are critically dependent on the surface chemistry. This allows us to modulate the energy levels of CQDs by introducing different ligands on CQDs surface. Fig. 7e shows the chemical structures of 12 most common adopted ligands employed for replacing OA for higher electrical conductivity, including thiols [benzenethiol (BT), 1,2-, 1,3-, and 1,4-benzenedithiol (1,2-BDT, 1,3-BDT, and 1,4-BDT), 1,2-ethanedithiol (EDT), and 3-mercaptopropionic acid (MPA)], a primary amine [1,2-ethylenediamine (EDA)], ammonium thiocyanate (SCN), and halides [tetrabutylammonium iodide (TBAI), bromide (TBABr), chloride (TBACl), and fluoride (TBAF)].114 The energy levels of PbS CQDs (with the absorption peak at 963 nm) after exchanged with these different ligands are measured by photoemission and absorption spectra and shown in Fig. 7f.114 A shift of energy levels by up to 0.9 eV between different chemical ligand treatments can be obtained, matching well with the atomistic density functional theory calculation results. Even between chemically similar ligands, the shift of energy levels can also reach 0.2 eV. This arises from different surface dipoles induced between CQDs and different ligands. Furthermore, different ligand treatments of CQDs can also results in n-/p-type behaviors and different doping levels of CQDs, as shown in Fig. 7g.115 Usually, halide ligands including TBAI, CdI2 and CdBr2 induce n-type doping in CQDs, while thoils including MPA, EDT and 1,2-BDT induce p-type doping in CQDs. It should also be noted that ligand exchange atmosphere (air or inert) has significant effects on doping levels of CQDs. The surface-chemistry-mediated band structures and doping levels allow us to predictably modulated electronic properties of CQDs films and optimize the energy level structure of CQDs photodetectors, which is consistent with bulk materials (in bulk materials such as Si, Ge, InGaAs, etc., the band structure and doping levels can be modulated by adjusting the doping element and concentration).
Recently, both theoretical and experimental studies have revealed the origin of trap states of CQDs, as shown in Fig. 8a.117 In analogy to their bulk counterparts, trap states in CQDs are typically associated with deep-level electronic states located within the bandgap of individual CQDs (i.e. in-gap trap state model). The electronic mid-gap states in individual CQDs mainly arise from physical surface defects including surface vacancies, additional atoms, uncoordinated surface atoms, Pb–Pb formation, etc.105,122–124 However, recent theoretical computational studies show that even nonstoichiometric CQDs generally tend to exhibit defect-free bandgaps.125 Until now, this model has received the most attention, and various surface defect passivation strategies have been proposed to reduce density of electronic mid-gap states to increase EQE and reduce Jd of CQDs SWIR photodiodes. Two other origins of trap state formation have also been identified and experimentally demonstrated, which can occur simultaneously in the same CQDs film. Firstly, trap states emerge as a result of two CQDs fusing together (i.e. CQDs dimers), which is greatly dependent on the synthesis and thin film fabrication steps of solution-processed CQDs. Recently, surface –OH groups have been demonstrated to induce epitaxial CQDs fusion/dimer.126,127 The fused CQDs exhibit less pronounced electronic confinement and thus show a smaller bandgap, which can introduce intermediate states within the bandgap. In detail, an electron/hole trap arises from the energy mismatch of conductive/valence band between a CQDs dimer and surrounding non-fused CQDs. Therefore, CQDs fusion or dimerization has also been identified as a trap state formation model (i.e. dimer state model). In recent years, the significance of dimer state model (i.e. the effects of CQDs fusion/dimer on photodetection performances) has also received increased research focus. Secondly, trap states can also originate from doped CQDs (i.e. charge state model).128 Reduction (oxidation) of p-type (n-type) doped CQDs cause the shift of entire electronic band structure with respect to its surrounding non-doped CQDs, also resulting in a hole (or electron) trap. However, the effects of oxidation or reduction of doped CQDs on photodetection performances of CQDs SWIR photodiodes remains relatively unexplored and warrants further investigation.
Fig. 8 Traps in CQDs. (a) Schematic diagram of the models of electronic trap state formation in the bandgap including (i) in-gap stap state model, (ii) dimer state model and (iii) charge state model. Reproduced from ref. 117 with permission from [American Chemical Society], copyright [2020]. (b) Experimental techniques to probe the electronic structures and thus trap state information in CQDs including photoemission spectroscopy, electrochemical voltammetry and scanning tunneling spectroscopy. Reproduced from ref. 118 with permission from [American Chemical Society], copyright [2020]. |
While the physical origins of trap states in the above three models differ, each provides a plausible explanation for the defect state behaviors observed in device performance, and thus the insights into all three models could enable further improvement of photodetection performances to theoretical values. Owing to strong correlation between trap states and electronic structures, various optical and electronic techniques have been developed to determine the electronic structure in CQDs, which consist of photoemission (or photoelectron) spectroscopy, electrochemical voltammetry and scanning tunneling spectroscopy (Fig. 8b).118 Full potential of CQDs photodetectors will be fulfilled after full understanding and control of the electronic structure of individual CQDs and also assembled CQDs films.
(2) |
(3) |
(4) |
Here, it should be pointed that R and EQE are usually spectrally invariant and should be as high as possible across the entire operational wavelength range.
In thin-film CQDs photodetectors, shot noise, thermal noise, and 1/f noise coexist, necessitating their careful consideration during design and performance evaluation. In specific applications, photodetectors may exhibit greater sensitivity to noise at particular frequencies. Therefore, in addition to assessing noise levels, it is also necessary to evaluate the spectral characteristics of the noise.
Noise equivalent power (NEP), a pivotal metric for quantifying the sensitivity, reflects the minimum detectable optical power. NEP represents the incident light power yielding a signal-to-noise ratio (SNR) equal to 1, with the unit of watt (W). Its definition is expressed as:
(5) |
Since in scales with device area and R is area-independent in photodiodes, it is unfair to directly compare NEP values of photodiodes with varying device areas. To facilitate the comparison between devices with varying device areas, NEP is normalized to the square root of the device area Ad, and defined as D*, which is formulated as:
(6) |
For noise and D* measurement, two points should be considered. (i) in depends on the applied voltage bias and operating frequency, while R is dependent on wavelength, applied light intensity, and applied bias voltage. Therefore, photodetection performance comparisons between different devices should take these conditions into consideration, and only the comparison between parameters measured and calculated under the same experimental conditions are meaningful. (ii) The contribution of dark current (shot noise, directly calculated from ) is usually adopted as the only source of noise, leading to an overestimation of specific detectivity. It should be noted that in this case, thermal and 1/f noise are ignored, meaning that the calculated D* represents the maximum theoretical value the device can achieve theoretically. Because it is not always possible to compare D* values measured under the same experimental conditions, the calculation and comparison of maximum D* values are preferred. However, accurate evaluation of NEP and D* should consider all noise sources, not just the contribution of dark current. Therefore, the realistic and accurate noise power spectrum, comprising components such as shot noise, thermal Johnson noise, and 1/f noise spectral power density (i1/f2(ω)), can be represented as:134
(7) |
Furthermore, higher Jd increases the noise current and thus decreases D*. It should be noted that, unlike R, Jd can span multiple orders of magnitude depending on device structure and material properties. During the design and optimization of detector performance, rigorous control of Jd must be taken into account to ensure effective operation under low-light scenarios while maintaining superior imaging quality and signal processing accuracy. Moreover, a high level of Jd usually results in significant challenge in designing ROICs and further imaging accuracy in CQD SWIR imagers. Therefore, from the aspects of photodetection performance optimization, ROIC design and final imaging applications, suppression of Jd is crucial for CQD photodetectors to be competitive with traditional III–V binary and ternary semiconductor SWIR photodetectors and imagers.
Jd = Jinj + JTAT + JSRH + JDIFF + JSH | (8) |
This model describes a two-step process: the initial carrier injection into the transport layer followed by a certain probability of the carrier escaping to the opposite electrode. It considers key factors such as the image charge effect at the electrode, the energetic site disorder (e.g., interface trap states inherent to the HBL/EBL), and the hopping charge transport mechanism in the CQDs layer. Charge carriers are injected from the Fermi level of electrodes into tail states of DOS of the transport layer. As these carriers move away from the electrode interface, they drift under the electric field towards the opposite electrode, undergoing a diffusive random walk through the disordered energetic landscape. Notably, this injection process is not spatially homogeneous but rather filamentary, driven by spatially fixed states like trap states near the interface between electrodes and tranport layers, which create localized pathways for carrier movement.
Based on the assumptions of the Arkhipov et al. model, the reverse injection current, Jinj, which depends on the charge carrier density at the metal–semiconductor interface (nint) and the carrier mobility (μ0) within the semiconductor bulk, can be defined as follows:
Jinj = qnintμ0F | (9) |
(10) |
In this equation, A refers to a dimensionless prefactor considered equal to unity, and Nh represents the volume density of molecular sites between which the hopping takes place. Besides, V and L denotes the bias voltage applied and the thickness of active layer, respectively.
(11) |
In this equation, mt stands for the effective tunneling masses, NT the activated trap density, Eg the bandgap energy, Et the trap state energy, and Emax the maximum electric field, respectively. Moreover, the trap potential, commonly adopted in infrared detectors, is M2 = 10−23 eV2 cm3.
To mitigate TAT, improving the material quality to reduce trap density is crucial. In CQDs photodetectors, deep-level defects typically arise from incomplete passivation during ligand exchange or local CQDs concentration non-uniformities after spin-coating. The influence of band-to-band tunneling (BTBT) is generally not considered because the CQDs layer is relatively thick, resulting in a lower electric field strength.
Under reverse bias conditions, the carrier concentration in the depletion region is significantly lower than that at equilibrium, resulting in enhanced generation of electron–hole pairs via defect states within the bandgap. The thermal excitation of an electron from the valence band to a trap state leaves a free hole in the valence band, which subsequently moves to the electrode by drift or diffusion. Moreover, the trap state energy landscape can be altered by an external field, known as the Poole–Frenkel effect,137 which reduces the effective energy required for electrons to escape the trap. Consequently, the SRH generation current increases with the increase of reverse bias due to the greater depletion of carriers and reduced activation energy. Conversely, under forward bias, the carrier concentration is significantly higher than at equilibrium, and recombination predominates over generation. Defect states within the bandgap serve as efficient recombination centers, increasing the recombination current in proportion to the density of these states. The equation of the SRH current can be considered as:
(12) |
The SRH current is characterized by an activation energy typically around half of Eg, indicating the significant role of mid-gap traps. In CQDs photodetectors, the density and distribution of these traps are influenced by surface chemistry and the quality of CQDs passivation. Reducing the density of mid-gap states through improved surface passivation techniques can effectively lower the SRH generation–recombination current.
(13) |
(14) |
Understanding the various mechanisms contributing to the dark current in CQDs SWIR photodetectors is crucial for enhancing their performance. Current research indicates that the primary contributors to reverse dark current in devices are predominantly JSH and JTAT. Particularly under large reverse bias conditions, JTAT can be more dominant due to typically high trap states in CQDs.53,55,56 This makes trap-assisted mechanisms somewhat more unique in CQDs devices compared to their bulk counterparts. Addressing these mechanisms through material optimization, defect reduction, and advanced fabrication techniques can lead to significant improvements in device sensitivity and reliability. Future research should focus on developing strategies to mitigate these dark current sources, thereby paving the way for high-performance CQDs photodetectors suitable for a wide range of infrared sensing applications.
(15) |
The function ϕ(E) can be derived from the blackbody spectrum.141 Multiplying this by a factor of q and converting it to units of mA cm−2 eV−1 provides the spectrum over a range of temperatures (Fig. 11a). The simulated EQE spectra of the optical absorption material can be modeled by a function of the form:
(16) |
Fig. 11 Intrinsic limit of dark current calculations. (a) Blackbody thermal photon flux ϕ(E) with a factor q and simulated EQE spectra with EQEmax, Eg and σ. (b) The temperature-dependent dark current density calculated through eqn (15). The red solid line represents the result of experimental EQE measurements of CQDs. Additional data points represent the dark current density of different materials, including HgTe,46,57 PbSe,43,45 PbS,13,39,49,53 and Ag2Te,98 extracted from previously published literature. |
In 2011, Sarasqueta et al. experimentally verified the role of hole/electron blocking layers and compared the blocking effects of various HBLs and EBLs in the indium tin oxide (ITO)/MoO3/PbSe CQDs/Al photodiodes, as shown in Fig. 12a-I.43 By introducing HBLs and EBLs to suppress the minority carrier injection from ITO and Al electrodes, Jd was obviously suppressed accompanying with the improvement of photocurrent and thus specific detectivity. Through comparative studies, p-type Poly-TPD and n-type ZnO nanoparticles exhibited higher hole and electron blocking effects (Fig. 12a-II), while ensuring the transportation of photo-generated electrons and holes. Ka et al. introduced low-cost solution-processed CuSCN as EBL/HTL of PbS CQDs photodiodes.44 Regardless of the electrode material, the topper minimum conduction band contributes to a high energy barrier, preventing the injection of minority electrons from Au electrodes (Fig. 12b-I). The dark current drops about two orders in the resulting devices, reaching 6 nA cm−2 under −1 V bias (Fig. 12b-II). The simplified method and attractive performances render CuSCN as a potential candidate of EBLs. Until now, various HBLs and EBLs have been proposed and adopted in PbS photodiodes to simultaneously block the minorities and transport the majorities. The commonly used EBLs include metal oxides (e.g., NiOx), thiols-capped PbX CQDs and organic materials (e.g., TFB, Poly-TPD, TAPC, CuSCN), while HBLs mainly include metal oxides (e.g., ZnO, TiO2, WO3, SnO2), metal sulfides (e.g., Bi2S3, CdS, SnS2), halide-capped PbX CQDs, and organic materials (e.g., C60, PCBM), as shown in Fig. 12(c-I).38,43,44,142,143
Fig. 12 Energy band alignment optimization of ETLs and HTLs to suppress reverse carrier injection and thus Jd. (a-I) Energy band alignment of photodiodes utilizing EBLs and HBLs. (a-II) Jd and specific detectivity values of photodiodes utilizing different EBLs and HBLs. (b-I) Energy band alignment of photodiodes with CuSCN layer. (b-II) J–V curves of Au and CuSCN/Au photodiodes in dark and under illumination. (c-I) Energy band diagrams of photodiodes utilizing different inorganic and organic ETLs. (c-II) Schematic illustration of Si integrated front-side illuminated PbS CQDs photodiodes utilizing ZnO and SnO2 ETLs. (c-III) Comparison of voltage and Jd of photodiodes with SnO2 and ZnO ETLs. (d-I) and (d-II) Energy band diagram of Au/Ag:HgTe/HgTe/ITO photodiodes w/o and w/Bi2S3 ETLs. (d-III) Cross-sectional SEM image of glass/ITO/Bi2S3/HgTe CQDs/Ag:HgTe CQDs/Au photodiodes. (d-IV) J–V curves of HgTe CQDs photodiodes w/o and w/Bi2S3 ETLs in dark and under 1550 nm illumination. Figure reproduced with permission from: (a) ref. 43, [John Wiley and Sons], copyright [2011]; (b) ref. 44, [American Chemical Society], copyright [2020]; (c) ref. 38, [American Chemical Society], copyright [2023]; (d) ref. 142, [American Chemical Society], copyright [2023]. |
The EBLs and HBLs, especially when deposited on top of the CQDs layer, were mainly fabricated using a solution process due to its fabrication simplicity and negligible damage to the underlying CQDs layer. However, it is challenging for solution-process EBL and HBL layers to ensure excellent crystallization and low defect density in such a low-temperature process (Usually, the PbS CQDs films cannot tolerate long-time annealing process above 100 °C). Starting from this point and also considering higher compatibility with standard CMOS process, Zhang et al. demonstrated atomic layer deposition (ALD) as a suitable way for depositing oxide functional layers on top of CQDs layer at low temperature (Fig. 12c-II).38 During the sputtering process, it is possible to import impurity ions in the functional layer and introduce damage to the below CQDs layer, while ALD is a mild strategy to avoid such issues. Zhang et al. utilized ALD prepared SnO2 as the blocking layer here instead of sputtered ZnO. Owing to better aligned energy level and more qualified heterojunction interface prepared from ALD, the SnO2 device obtained an extremely low Jd of 3.5 nA cm−2 at −10 mV (Fig. 12c-III). Furthermore, Bi2S3 is reported as a new type of ETL/HBL suitable for HgTe CQDs SWIR photodiodes.142 As shown in Fig. 12d-I, the suitable energy levels of Bi2S3, along with their efficient electron extraction efficiency, could simultaneously suppress the reverse minority carrier injection from ITO to HgTe CQDs and enable the transpothe transportation of electrons from HgTe CQDs to ITO, which is expected to suppress Jd and improve Jph. The high surface flatness of Bi2S3 films also endows the multilayer structure of Au/Ag:HgTe CQDs/HgTe CQDs/Bi2S3/ITO photodiodes with qualified heterojunction interface, as shown in Fig. 12d-II. As expected, the integration of Bi2S3 ETL/HBL has led to a substantial decrease in Jd, down to as low as 1.6 × 10−5 A cm−2 at −400 mV (Fig. 12d-III), and a remarkable increase in D* to approximately 1011 Jones at room temperature.
Fig. 13 Interfacial and bulk defect passivation of ETLs and HTLs to suppress defects assisted Jd. (a-I) J–V curves in dark and under illumination of PbS CQDs photodiodes utilizing sol–gel ZnO and sputtered ZnO ETLs. (a-II) Schematics for the mechanism of trap-assisted tunneling current. (b-I) Stability of photocurrent density under constant IR illumination (5 mW cm−2) before and after 37 h UV activiation. (b-II) J–V characteristics of ALD ZnO devices in dark and under various illumination intensities. (c-I) Energy band diagram of ITO/HTL/PbS CQDs/ETL/Ag photodiode with PbS-TBAI or PbS-TFCA ETLs. (c-II) Schematics for ion migration at Ag/TBAI and Ag/TFCA interfaces under large and long-time external electric field bias. (c-III) Jdversus time curves of photodiodes utilizing PbS-TBAI and PbS-TFCA ETLs at −0.3 V bias. (d-I) Schematics for the mechanisms of defect assisted tunneling and generation–recombination current. (d-II) J–V curves in dark and under illumination of PbS CQDs photodiodes before and after oxidation. (e-I) Schematic diagram and cross-sectional SEM image of Si integrated front-side illuminated ITO/PbS-EDT/PbS-I/Si/Al photodiode w/o and w/ZnO ETLs. (e-II) Jd–V curves of Si integrated photodiodes w/o and w/ZnO ETLs. Figure reproduced with permission from: (a) ref. 40, [American Chemical Society], copyright [2023]; (b) ref. 39, [John Wiley and Sons], copyright [2022]; (c) ref. 37, [John Wiley and Sons], copyright [2022]; (d) ref. 42, [American Chemical Society], copyright [2023]; (e) ref. 145, [American Institute of Physics], copyright [2020]. |
Besides, the critical role of defects in CQDs HTL/EBL and ETL/HBL on Jd is also verified. Zhang et al. introduced n-type PbS CQDs armed with a strongly bound ligand, trans-4-(trifluoromethyl) cinnamic acid (TFCA), instead of a weakly bound inorganic ligand, TBAI, as a new class of ETL/HBL in homojunction photodiodes (Fig. 13c-I), which combined the advantages of more flexible energy band engineering and surface passivation.37 As a result, the SWIR photodetector realized 66% EQE and 1 × 10−3 mA cm−2 dark current under −1 V bias. Moreover, TFCA suppressed the bias-induced ion migration inside ETL (Fig. 13c-II), and thus improved the operation stability of photodiodes by 50 times compared to TBAI, resulting in a lower Jd at reverse bias stress (Fig. 13c-III). Wang et al. introduced a controlled oxidation process to passivate the defects of PbS-EDT HTL/EBL layer and thus suppress Jd.42 This process involves annealing the PbS CQDs by exposing them to air at 90 °C with a relative humidity of 35% and an ambient temperature of 25 °C. The controlled oxidation process led to the surface passivation of PbS-EDT films and suppression of the nonradiative recombination. As a result, the TAT current is significantly suppressed, and the oxides also act as gap fillers for cracks in PbS-EDT films fabricated from an SSLE process, which helps in suppressing ohmic leakage, as shown in Fig. 13d-I. Benefiting from this physical process, Jd greatly decreases from 8159 nA cm−2 to 218 nA cm−2, as shown in Fig. 13d-II. Nevertheless, it was also demonstrated that excessive oxidation procedures would lead to a substantial decrease in photocurrent. Thus, it is significantly important to find an appropriate level of oxidation for PbS-EDT films to achieve a balance between dark current and photocurrent.
The construction of the heterojunction structure between PbS CQDs films and bulk Si paves another way to achieve CMOS-compatible SWIR photodiodes. However, the fabrication of high quality CQDs:Si van der Waals heterojunctions is always hindered by the large energy offset between Si and small bandgap PbS CQDs, as well as a high density of surface defects on the Si surface. Xiao et al. inserted a layer of solution-processed ZnO nanoparticles between Si and PbS CQDs to simultaneously introduce energy barriers between Si and PbS CQDs and passivate the surface dangling bond of Si (Fig. 13e-I), resulting in suppressed reverse minority injection from Si to PbS CQDs.145 As a result, the Si:CQDs heterojunction photodetector shows a significantly reduced Jd at reverse bias (Fig. 13e-II), which results in a high D* of 4.08 × 1011 Jones at −0.25 V bias at room temperature.
Fig. 14 Improving the electrode/carrier transport layer interface to suppress Jd. (a-I) Schematic illustration of the interfacial SIL LiF layer. (a-II) Schematic diagram of carrier transport in SIL LiF layer with different LiF thicknesses. (a-III) Voc and Jd of initial and aging photodiodes w/o and w/interfacial SIL LiF layer. (b-I) Device structure of an ITO/NiOx/PbS/ZnO/AI photodiode. (b-II) Schematic illustration of self-formed interfacial AlOx layer at the ZnO/Al interface and mechanism of Jd suppression. (b-III) J–V curves of photodiodes before and after post-exposure to dry air for different periods of time. (c-I) Simplified diagram of two adjacent pixels in the top view. (c-II) Energy band diagram of the PbS CQDs SWIR photodiodes. (c-III) Jd–V curves of organic and CQDs photodiodes developed in this imager chip. (d-I) Device structure of the top-illuminated PbS CQDs photodiode integrated on Si utilizing solution-processed MXene TCEs instead of sputtered ITO. (d-II) Energy band diagrams of the corresponding photodiode. (d-III) Experimental and simulated Jd–V curves. Figure reproduced with permission from: (a) ref. 48, [American Chemical Society], copyright [2024]; (b) ref. 49, [American Institute of Physics], copyright [2023]; (c) ref. 83, [Springer Nature], copyright [2023]; (d) ref. 147, [John Wiley and Sons], copyright [2024]. |
Except for introducing an additional interfacial layer, Yang et al. demonstrated how the self-formed AlOx at the ZnO nanoparticles/Al interface dramatically reduced Jd of the PbS CQDs SWIR photodiodes with an ITO/NiOx/PbS CQDs/ZnO/Al structure (Fig. 14b-I).49 In pristine PbS CQDs photodiodes, there exists a large number of defect-induced states at the ZnO/Al interface,148,149 which provide hopping sites for hole injection from Al electrode under a reverse bias, leading to a relatively high Jd. However, after post-exposure to dry air, these interfacial defects were thoroughly passivated by the formation of an interfacial AlOx layer, which reduced the defect-induced hopping sites for minority hole injection from the electrode, consequently lowering Jd (Fig. 14b-II). The resulting device showed an impressively low Jd of 1.58 × 10−7 A cm−2, a high R of 0.62 A W−1 at 1413 nm, and a remarkably high D* of 2.05 × 1012 Jones at −0.5 V bias. Moreover, the suppression of Jd is highly dependent on the duration of dry air post-exposure (e.g., coverage of AlOx layer at ZnO/Al interface), as illustrated in Fig. 14b-III.
For front-side illuminated photodiodes, a transparent conducting electrode (TCE) is always required, with ITO being the most commonly adopted SWIR TCE. However, during the ITO sputtering process, the high-energy particles and plasma emission can irreversibly damage the bottom functional layer,150 potentially creating alternative leakage pathways and consequently increasing Jd. To reduce contact resistance and minimize damage effects of ITO sputtering, Lee et al. inserted a 10 nm N1,N1′-(biphenyl-4,4′-diyl)bis(N1-phenyl-N4,N4-di-m-tolylbenzene-1,4-diamine) (DNTPD) layer with 5% F6TCNNQ doping between DNTPD HTL and sputtered ITO TCE (Fig. 14c-I).83 Along with optimized energy band alignment (Fig. 14c-II), Jd of CQDs photodiodes with a pixel size of 10 × 10 μm2 was minimized to be as low as 1.1 × 10−7 A cm−2 at −3 V bias (Fig. 14c-III). Futhermore, instead of adding an extra layer, solution-processed TCEs were fabricated on top of photodiodes instead of the sputtered ITO TCE, thereby entirely avoiding the damage effects during the ITO sputtering process. Owing to their high electrical conductivity, suitable energy levels and high chemical stability, two-dimensional (2D) MXenes are often used as an alternative TCE to replace brittle, expensive and diffusive ITO.151 Di et al. proposed a solution-processed interface design of 2D Titanium Carbide (Ti3C2) MXene and PbS CQDs layers for photodiodes with an MXene/PbS-EDT/PbS-I/ZnO/Au structure (Fig. 14d-I).147 The solution-processed interface engineering of MXene/PbS CQDs remarkably decreases the interface defect density by forming a low-defect interface, which can be explained by avoiding the bombardment effects of sputtering and covering the surface cracks (Fig. 14d-II). A low Jd of 0.2 μA cm−2 at −0.5 V bias voltage is thus achieved (Fig. 14d-III). Meanwhile, the device exhibits a LDR of 140 dB, a −3 dB bandwidth of 0.76 MHz, and a D* of 5.51 × 1012 cm W−1 Hz1/2 in the self-powered mode.
Fig. 15 Thickness, morphology and composition optimization of PbS CQDs films to suppress Jd. (a-I) Dependence of Jd and (a-II) Dependence of serial and shunt resistance on PbS CQDs film thickness. (a-III) J–V curves of the PbS-EMII/PbS-EDT 136/44 nm and 136/66 nm devices in dark and under illumination. (b-I) The surface profile curves and (b-II) Jd–V curves of photodiodes utilizing different PbS CQDs film thicknesses. (b-III) Dependance of Jd and EQE on PbS CQDs film thickness. (c-I) Schematics of the PbS CQDs/PI composite films. (c-II) SEM images of the PbS CQDs films w/o and w/PI. (c-III) J–V curves of PbS CQDs photodiodes w/o and w/PI in dark and under illumination. (d-I) Cross-sectional SEM images of PbS CQDs photodiode. (d-II) SEM images of PbS and PbS/CuInSeS CQDs films. (d-III) Jd–V curves of PbS CQDs photodiodes w/o and w/CuInSeS layer. (e) Energy band diagram of electron transport in the PbSe-perovskite blending layer. (f) Schematics of local band bending and hole transport hindering caused by Ag nanocrystals. (g) Schematic of majority carrier reduction by Ag NPs doping. Figure reproduced with permission from: (a) ref. 50, [American Chemical Society], copyright [2022]; (b) ref. 51, [American Institute of Physics], copyright [2024]; (c) ref. 52, [Elsevier], copyright [2023]; (d) ref. 13, [Royal Society of Chemistry], copyright [2024]; (e) ref. 153, [John Wiley and Sons], copyright [2022]; (f) ref. 154, [American Chemical Society], copyright [2014]; (g) ref. 155, [John Wiley and Sons], copyright [2015]. |
However, though such method could enhance film quality, it is time-consuming and incompatible with industrial-scale production, where it is generally preferable to fabricate the CQDs film layer by one-step spin-coating process directly from a high-concentration CQD ink. Liang et al. introduced a Lewis base polyimide (PI) into the CQDs ink, and the schematic of PbS CQDs/PI composite is shown in Fig. 15c-I.52 After introducing PI into the PbS CQDs ink, the monodispersity is significantly improved, resulting in improved defect passivation and film morphology of PbS CQDs (Fig. 15c-II). The resulting CQDs photodiodes showed a tenfold reduction in defect density and a two-order magnitude reduction in Jd at −0.4 V compared to control devices (Fig. 15c-III). In addition to improving the quality of the CQDs ink, the interface between the CQDs layer and the carrier transport layer is crucial. A recently reported low dark current (about 70 nA cm−2) is attributed to a high-quality n-type CQDs solid film created via a novel solution-phase ligand exchange method. This approach, combined with mild MA ligand passivation, results in an optimal surface chemical environment, effectively passivating top interface defects, and reducing recombination at the CQDs ink/HTL heterojunction interface.47
Additionally, traditional SSLE techniques often cause inherent cracks, leading to significant leakage currents and limiting the sensitivity of detectors. To address this, Chen et al. introduced a CuInSeS CQDs interfacial layer on top of the PbS-EDT CQDs layer, and the device schematic is shown in Fig. 15d-I.13 After the modification of interfacial CuInSeS layer, the CQDs active layer results in a smoother and crack-free layer (Fig. 15d-II), which is expected to improve the interfacial contact between PbS CQDs and ZnO films. Benefitting from the improved film morphology and band alignment, Jd is significantly suppressed from 2 × 10−7 A cm−2 to 1.15 × 10−9 A cm−2 under −1 V bias (Fig. 15d-III).
Furthermore, under high-temperature and strong electric field conditions, halide ion migration through vacancies can increase defect density in the CQDs layer and block device interfaces, resulting in an increased Jd and delayed saturation voltage. Chen et al. introduced polyimide (PI) into CQDs films to effectively block halide ion migration,146 thereby reducing Jd under high bias and enhancing the stability of CQDs photodiodes under high-temperature and strong electric fields.
The dark current characteristics of photodetectors can also be optimized through plasmonic structures. He et al. demonstrated that co-depositing Ag nanoparticles (NPs) into the CQDs solution reduces device dark current. In darkness, Ag NPs induce local band bending and hinder hole transport, thus reducing dark current (Fig. 15f).154 A significant reduction in dark currents, up to two orders of magnitude, has been reported in ligand-exchanged PbS-based photodetector devices when PbS NPs are blended with 10% Ag nanocrystals (NCs).155 The Ag NCs, acting as electron donors due to their low work function, can effectively passivate the shallow electron traps present in the bandgap, thereby reducing the concentration of majority holes in the PbS films (Fig. 15g).
Fig. 16 Trap state management in PbS CQDs films to suppress Jd. (a-I) Differential charge densities of 4-AMPY adsorption on the (001) facet of PbS CQDs. (a-II) Adsorption energies of Namino, Npyridine, I− and Br− on PbS(100) facet. (a-III) J–V curves of PbS CQDs photodiodes w/o and w/4-AMPY passivation in dark and under 1300 nm illumination. (b-I) Schematic diagram of low level of lead bromide (LLB) and high level of lead bromide (HLB) treatment methods. (b-II) Jd–V curves of devices with different exchange routes versus control ones. (c-I) Diagrams of PbS CQDs (100) facets passivated by 1Br−, 1I−, 8Br−, and 8I−. (c-II) Adsorption energy versus Br/I coverage curves. (c-III) Jd of devices passivated by increasing concentrations of PbBr2 at −0.1 V. (d-I) Schematic diagram and (d-II) TEM images of the fused and perovskite-bridged CQDs. (d-III) J–V curves of control and perovskite-bridged PbS CQDs photodiodes in dark and under 1300 nm illumination. (e-I) Schematic of Se:HgTe and HgTe CQDs and their abilities against oxidation. (e-II) TEM images of HgTe and Se:HgTe CQDs. (e-III) J–V curves of HgTe and Se:HgTe CQDs photodiodes in dark and under 1550 nm illumination. Figure reproduced with permission from: (a) ref. 53, [John Wiley and Sons], copyright [2024]; (b) ref. 54, [Elsevier], copyright [2021]; (c) ref. 55, [American Physical Society], copyright [2023]; (d) ref. 56, [John Wiley and Sons], copyright [2023]; (e) ref. 57, [John Wiley and Sons], copyright [2024]. |
Notably, the amount of surface ligands in the solution has also been reported to induce atomic rearrangement, thereby altering the shape of CQDs and contributing to the passivation of the defect site. Vafaie et al. pursued a novel approach to obtain high-quality films for large-size PbS CQDs (>5 nm).54 Compared to low-level bromine (LLB) treatment, high-level bromine (HLB) treatment (2:5 PbBr2/PbI2) effectively passivated the (100) and (111) facets and introduced dual-passivation to remove unwanted residual organics and enhance CQDs colloidal stability (Fig. 16b-I), resulting in a tenfold reduction in Jd compared to controlled photodidoes at −1 V (Fig. 16b-II). Similarly, Yang et al. developed an excess concentration PbBr2 ligand strategy for large-size PbS CQDs to effectively suppress the defects and increase the carrier lifetime,55 resulting in a minimal Jd of 156 nA cm−2 at −0.1 V (Fig. 16c).
In recent years, the approach of passivating surface defects in PbS CQDs by introducing perovskite materials has garnered significant attention due to its potential to markedly enhance their optoelectronic performance. Chen et al. proposed a novel passivation strategy for large-size (d > 4 nm) PbS CQDs using halide perovskite CsPbI3-xBrx heteroepitaxial bridging of charge-neutral (100) facets by adding a small amount of CsI in the PbX2 ligand solution.56 This approach led to reduced defect states and improved dispersion stability (Fig. 16d-I&II), yielding perofilms with enhanced EQE, decreased Jd from 1 × 10−6 A cm−2 to 2 × 10−7 A cm−2 at −0.4 V (Fig. 16d-III), and increased operational stability compared to control CQDs photodiodes. Similarly, a high-performance PbS CQDs based broadband photodetector incorporates small molecular-scaled (1 nm) CH3NH3PbI3 perovskite hybrids that function both as trap state passivators and visible light sensitizers.157 This integration reduces Jd from 9.4 × 10−8 A cm−2 to 2.1 × 10−9 A cm−2 at 0 bias, effectively minimizing the recombination losses and enhancing the broadband photoresponse. Furthermore, the surface morphology and chemical state of PbS CQDs significantly influence perovskite crystallization. Therefore, selecting appropriate halide ligands is crucial for achieving PbS CQDs with high conductivity, good morphology, suitable surface chemical states, and compatible energy levels with perovskites. The impact of different ligands on dark current can also vary by up to two orders of magnitude.158 However, the effect of different ligands on photocurrent should also be considered to achieve high detectivity.
For other photomaterials like HgTe CQDs, Ackerman et al. optimized HgTe CQDs/Ag2Te CQDs photodiodes in the SWIR range by introducing HgCl2 in the EDT/hydrochloric acid (HCl) exchange process,159 significantly enhancing Rsh and reducing Jd by tenfold at −500 mV reverse bias from approximately 4 × 10−3 A cm−2 to 4 × 10−4 A cm−2. Through their discussions, it was determined that the majority of the improvements stemmed from a reduction in non-radiative recombination rather than solely from a decrease in mobility. Additionally, Yu et al. employed precursor reactivity engineering, using an Se stabilization strategy based on HgTe CQDs surface Se coating (Fig. 16e-I).57 This can significantly improve the colloidal dispersion of CQDs ink (Fig. 16e-II), along with improved defect passivation and doping modulation. The resulting Se:HgTe CQDs SWIR photodetector exhibits an ultra-low Jd of 3.26 × 10−6 A cm−2 at −0.4 V, which is significantly reduced by an order of magnitude compared to the control HgTe CQDs devices (Fig. 16e-III). Yang et al. also fabricated a PIN photodiode using thermally evaporated Bi2Se3 film/HgTe CQDs/Ag2Te CQDs stacking technology.160 This device employed a ligand engineering approach to produce well-separated HgTe CQDs with a sharp absorption edge, achieving a rectification ratio of two thousand and a Jd of 2.3 × 10−6 A cm−2 under −0.4 V reverse bias.
For CuInSe2 CQDs devices, Guo et al. demonstrated that Mn2+ doping improves CuInSe2 CQDs crystal quality.161 The Mn2+ doping acts as a hole capturer forming charge compensating pairs with Cu2+ defects, leading to long-lived Cu2+ radiative recombination. Furthermore, Mn2+ doping alters the conduction band minimum and valence band maximum levels, enhancing carrier transport properties. The optimal Mn2+ doping level (0.01 Mn feed ratio) balanced these factors, achieving a suppressed Jd of 1.6 × 10−10 A cm−2.
Finally, Table 1 contains a summary of representative strategies to suppress Jd of CQDs photodiodes published recently. In each work, the main Jd mechanisms, experimental approach to suppress Jd and resulting photodetection performance including detection wavelength, Jd, R and D* are provided.
Device structure | λ peak (nm) | J d mechanism | J d reduction strategies | J d (A cm−2) | R (A W−1) | D* (Jones)1 | Ref. |
---|---|---|---|---|---|---|---|
a The values not annotated with @Hz were calculated considering only shot noise, with other noise sources disregarded. | |||||||
ITO/PbS-EDT/PbS-mixed halide/PbS-TFCA/Ag | 1450 | Reverse injection | Strong ligand passivation of HBL | 1 × 10−6@−1 V | 0.766@−1 V | 3 × 1010@0 V@10 kHz | 37 |
ITO/SnO2/C60/PbS/NiO/Au | 1550 | Reverse injection | Heterojunction interface modification of HBL | 1 × 10−7@−10 mV | 0.82@−10 mV | 6.67 × 1011@−10 mV@100 kHz | 38 |
ITO/ZnO/PbS/PbS-EDT/Au | 950 | Reverse injection | ALD ZnO reducing oxygen adsorption | 2.9 × 10−8@−1 V | 0.536@−1 V | 1.66 × 1011@−1 V@10 kHz | 39 |
ITO/ZnO/PbS-BTA/PbS-EDT/Au | 1300 | Reverse injection | [002]-oriented ZnO reducing H2O absorption | 8.49 × 10−8@−0.1 V | 0.552@−0.1 V | 2.15 × 1012@0 V@94.6 kHz | 40 |
ITO/NiO/PbS-BDT/ZnO/Al | 1135 | Reverse injection | Energy level blocking | 3.4 × 10−8@−1 V | 0.2@−1 V | 1.1 × 1012@−1 V@100 Hz | 41 |
ITO/ZnO/PbS-I&Br/PbS-EDT/Au | 1300 | Reverse injection | Passivation of interface defects by oxidation control | 2.18 × 10−7@−0.6 V | 0.33@0 V | 6.8 × 1012@0 V@45 kHz | 42 |
ITO/MoO3/TFB/PbSe/ZnO/Al | 1100 | Reverse injection | Energy level blocking | 2 × 10−7@−1 V | 0.05@−1 V | 3 × 1011@−0.5 Va | 43 |
FTO/TiO2/PbS-QDs/CuSCN/Au | 1258 | Reverse injection | Energy level blocking | 6 × 10−9@−1 V | 0.2@−1 V | 7 × 1010@−1 Va | 44 |
ITO/PEDOT:PSS/PbSe/ZnO/Poly-TPD/PbSe/ZnO/Al | 1100 | Reverse injection | Coupling of double active layer | 4 × 10−7@−0.5 V | 0.14@−0.5 V | 1 × 1012@−0.5 Va | 45 |
FTO/SnO2/HgTe/Ag2Te/Au | 1750 | Reverse injection | Energy level blocking | 2 × 10−5@−0.2 V | 0.3@0 V | 5 × 1010@0 V@1 kHz | 46 |
ITO/LiF/SnO2/PbS-I/PbS-MA/MoOx/Au | 1050 | Reverse injection | Carrier-selective layer at HTL/electrode interface | 7 × 10−8@−0.5 V | 0.61@−0.5 V | 1.42 × 1012@−0.5 V@100 Hz | 47 |
ITO/ZnO/PbS-I&Br/PbS-EDT/Au | 1290 | Reverse injection (trap-assisted) | SIL passivation layer at HTL/electrode interface | 5.94 × 10−7@−0.05 V | 0.48@0 V | 4.8 × 1012@0 V@43 kHz | 48 |
ITO/NiOx/PbS/ZnO/Al | 1390 | Reverse injection (trap-assisted) | Surface Passivation of HTL at HTL/electrode interface | 1.58 × 10−7@−0.5 V | 0.62@−0.5 V | 2.05 × 1012@−0.5 V@1 kHz | 49 |
Au/PbS-EDT/PbS-EMII/SnO2/ITO | 1100 | TAT & shunt & reverse injection | Optimization of CQDs thickness | 3.15 × 10−6@−0.5 V | 0.24@−0.5 V | 1.24 × 1010@−0.5 V@100 Hz | 50 |
Al/TiO2/PbS-ZnI2:MPA/NiOx/ITO | 1420 | TAT & shunt | Optimization of CQDs thickness | 4.1 × 10−6@−3 V | 0.65@−3 V | 1.78 × 1012@−3 Va | 51 |
Au/PbS-EDT/PbS-PbX2/ZnO/ITO | 1300 | TAT & shunt | PI passivating PbS-PbX2 interface defects | 1.23 × 10−7@−0.1 V | 0.36@0 V | 6.36 × 1012@0 V@100 kHz | 52 |
ITO/NiOx/PbS-EDT/CuInSeS/ZnO/Al | 940 | TAT & shunt | Introducing CuInSeS CQDs interfacial layer | 1 × 10−9@−1 V | 0.23@−1 V | 1.87 × 1012@−1 Va | 13 |
Au/PbS-EDT/PbS-4-AMPY/ZnO/ITO | 1300 | SRH & TAT | Supplemental double-ended ligands exchange | 7 × 10−8@−0.5 V | 0.532@−0.5 V | 2.5 × 1012@−0.5 V@100 kHz | 53 |
Au/PbS-EDT/PbS-PbX2/ZnO/ITO | 1550 | SRH & TAT | Introducing an added exchange step | 1.6 × 10−6@−0.1 V | 1@0 V | 8 × 1011@0 V@500 kHz | 54 |
Au/PbS-EDT/PbS-I&Br/ZnO/ITO | 1300 | TAT | Excessive PbBr2 concentration ligand exchange strategy | 1.8 × 10−7@−0.1 V | 0.239@−0.1 V | 5.22 × 1012@0 V@38 kHz | 55 |
Au/PbS-EDT/PbS-Pero/ZnO/ITO/Au | 1300 | SRH & TAT | (100) facets epitaxially bridged with perovskite | 9 × 10−8@−0.1 V | 0.54@0 V | 6.2 × 1012@0 V@42 kHz | 56 |
Au/Se:HgTe/Bi2S3/ITO | 2000 | SRH & TAT | Surface coating of Se on HgTe CQDs | 2.7 × 10−6@−0.1 V | 0.69@0 V | 5.17 × 1011@0 V@500 Hz | 57 |
While enormous progress has been made in origins and suppression of Jd in CQDs SWIR photodiodes, more attention needs to be paid to further minimize Jd while retaining EQE, as shown in the following aspects.
Firstly, a quantitatively physical model of Jd should be established for understanding CQDs device characteristics both in dark and under illumination. The current experimental research predominantly employs the popular Shockley–Quiesser (S–Q) diode model, based on which the main components for Jd were obtained by curve fitting of the measured Jdversus V curves. However, the S–Q diode model does not capture the underlying physics of CQDs junctions. In 2023, Arya et al. developed a compact and easy-to-use physical model for CQDs junctions assuming that the transport of carriers across the CQDs layer obeys the thermionic emission and tunneling processes.162 However, the surface state and size non-uniformity among CQDs, assembly and energetic disorder in CQDs films and high density of trap states is negligible in this model, which are main factors affecting the charge transport behaviors as discussed above and should be taken into account in the modeling of CQDs photodiodes. Moreover, owing to structural and energetic disorder in the CQDs assembly, hopping model is usually adopted for characterization of charge transport behaviors of solution-processed CQDs, and thus we should consider the establishment of device model based on hopping mechanisms in future studies.
Secondly, as discussed above, trap states both in bulk and at interfaces contribute significantly to carrier trapping/de-trapping process and thus high Jd in CQDs SWIR photodiodes, and thus various defect passivation strategies have been developed to effectively suppress Jd. However, it still remains unclear which kinds of trap states are dominant factors responsible for Jd, and thus advanced defect characterization techniques including temperature-dependent low-frequency noise spectra, temperature-dependent charge transport behaviors, photoluminescence spectroscopy, etc. should be employed to determine the main defect categories dominating Jd. Herein, the size dependence should be considered. Moreover, except for deep-level mid-gap states coming from physical defects at the CQDs surface, surface –OH groups have been experimentally confirmed to trigger the epitaxial CQDs fusion/dimerization. Several experimental reports have attributed carrier trapping to the –OH groups and CQDs fusion in CQDs solar cells, whereas the effects of such trap states on Jd is so far lacking. Furthermore, the shift of entire electronic band structure induced by reduction (oxidation) of p-type (n-type) doped CQDs usually results in a hole (or electron) trap, hindering the construction of ideal PIN structure photodiodes. Therefore, passivation of physical defects at the CQDs surface, suppression of dimerization or fusion and challenges for efficient doping of CQDs, while maintaining trap-free conduction, were identified as key strategies to minimize Jd. Future device research will aim at employing this insight from aspects of defect physics and passivation strategies to further reduce Jd.
Thirdly, the mainstream device structure for reported CQDs photodiodes is PIN diode structure, where CQDs active layer is employed as intrinsic layer to combine with p-/n-type carrier transport/blocking layer. The reported PIN structured CQDs photodiodes are mainly concentrated on the 1–2 μm detection range. However, when the detection range exceeds 2 μm, the bandgap of CQDs decreases with gradual change of conduction and valence band levels. It is thus challenging for the energy level alignment between small-bandgap CQDs and carrier transport layer, posing significant difficulty in obtaining a rather low Jd level in PIN photodiodes. Future research could focus on modulating the energy level and carrier concentration of carrier transport layers by doping and CQDs by surface state modification. Moreover, novel device structures such as barrier-type, quantum-tuned cascade multijunction, etc. could be an alternative to obtain low dark current and high photoresponse in CQDs photodiodes with a detection range higher than 2 μm.
Finally, more attention should be paid to device performances after long-term bias and illumination stress. In practical applications, long-term reverse bias and light illumination stress is usually applied to the device, and how the device performances especially Jd and Jph evolute after long-time large reverse bias and high light illumination requires to be systematically investigated. Moreover, after fabrication of CQDs SWIR photodiodes, the wafer should further undergo the back end of line (BEOL) process, which is mainly composed of interconnection between top electrodes of photodiodes and ROICs, deposition of passivation layer and final wafer level packaging. In the BEOL process, the effects of annealing at an evaluated temperature, temperature cycling, exposure to various gases and chemicals and deposition of light management layer and transparent passivation layer onto the device stack remains unclear, and need to be further explored in detail. This is critical for fabricating high-performance CQDs SWIR image sensor wafers and their practical applications.
This journal is © The Royal Society of Chemistry 2024 |