1. Introduction
With the growing demand for optical signal detection and optical communication systems across a wide range of ultraviolet (250–400 nm), visible (450–800 nm), and infrared (900–1700 nm) wavelengths, photodetectors (PDs) have become essential elements in various optoelectronic applications [
1,
2,
3]. Currently, most commercial PDs are fabricated using silicon (Si) or crystalline III–V semiconductors such as indium gallium arsenide (InGaAs) due to their broad spectral coverage, high mechanical stability, cost-effective scalability, and excellent spatial resolution [
4]. However, their mechanical rigidity and vacuum-based complex fabrication processes hinder their integration into flexible, wearable, and deformable substrates and large-area devices [
5,
6,
7,
8,
9]. Furthermore, state-of-the-art inorganic PDs often fail to meet the stringent performance requirements of modern applications, prompting extensive research into alternative photoactive semiconductors with enhanced photoresponse speed, sensitivity, and detectivity under low-light conditions [
10].
Recently, organolead trihalide perovskites have garnered significant attention as promising semiconducting materials for a wide range of optoelectronic applications. Their exceptional properties, including high light absorption coefficients, long carrier diffusion lengths, and superior charge carrier mobilities, render them highly suitable for integration into solar cells and display technologies [
11,
12,
13,
14,
15,
16]. The perovskite semiconductors are also applied in PD applications as effective photoactive materials [
17,
18,
19,
20]. However, a major limitation of organic cation-based perovskites is their instability in humid environments and high thermal/electrical stress, which poses a significant challenge for their practical implementation in real-world uses [
21]. To address the stability issue, all-inorganic lead halide perovskites, particularly cesium-based compositions (CsPbX
3, where X = I, Br, Cl), have been extensively investigated as potential alternatives due to their superior environmental resilience [
22]. These perovskites exhibit remarkable optoelectronic properties, including a high light absorption coefficient (>10
5 cm
−1) and impressive charge carrier mobility (~1000 cm
2 V
−1 s
−1), comparable to silicon. As a result, they have garnered significant attention for application in photodetectors (PDs) [
23]. For instance, CsPbBr
3 microwire-based PDs have been demonstrated to achieve an obvious responsivity improvement of 145% [
24]. Shoaib et al. synthesized ultralong CsPbBr
3 nanowire-based PDs with a responsivity of 4.4 × 10
3 A W
−1 and a response speed of 0.252 ms, highlighting their potential for advanced optoelectronic applications [
25].
Despite these promising advancements, the operational stability and reliability of all-inorganic perovskite-based PDs remain hindered, particularly when organic interface materials are incorporated into device architectures. Typically, organic
p-type hole-transporting materials such as poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS) and poly[bis(4-phenyl)(2,4,6-trimethylphenyl)amine] (PTAA) are utilized in perovskite PDs with a
p-i-n structure [
26]. Conversely,
n-i-p architectures often integrate inorganic electron-transporting layers (ETLs), including zinc oxide (ZnO) or tin(IV) oxide (SnO
2), which provide high mechanical stability and long-term reliability of the interface in devices [
27]. These inorganic
n-type oxide semiconductors support a stable foundation for perovskite layers, thereby enhancing device performance and stability. However, a key challenge in
n-i-p perovskite PDs with
n-type oxide semiconductors is balancing strong photoresponse with low dark-state current, because the inherently high conductivity of the oxide semiconductors can lead to excessive dark current, ultimately degrading the on/off ratio of PDs [
28]. To compensate for those restrictions, a bilayer structure ETL has been designed to improve the electrical properties of single-layer
n-type oxide semiconductor-based devices. Sun et al. have incorporated ZnO into the TiO
2 ETL to regulate the grain size and improve the surface uniformity of the perovskite absorber layer in photovoltaic devices [
29]. A bilayer ETL structure comprising SnO
2 and ZnO has also been developed for high-performance perovskite solar cells, enabling a tunable electronic structure for efficient electron extraction [
30]. However, bilayer ETLs composed of different
n-type metal oxides have not been extensively investigated in perovskite photodetectors, despite their potential to provide favorable energy level alignment with perovskite absorbers for self-powered operation.
In this study, we designed a bilayer inorganic ETL comprising SnO2 and ZnO for integration into high-performance all-inorganic perovskite PDs, aiming to enhance stability and operational reliability. This bilayer ETL design leverages the high electrical conductivity of ZnO to promote efficient electron transport, thereby improving charge extraction and transfer efficiency. As a result, the bilayer ETL-based PDs exhibited notable performance enhancements, achieving a responsivity that optimized charge extraction while suppressing dark-state current. The resulting PDs achieve a responsivity of 0.45 A W−1 at 630 nm, a specific detectivity of 9 × 1013 Jones, and a fast response time of less than 50 μs. These improvements surpass those observed in devices utilizing a single ETL layer (ZnO or SnO2) in all-inorganic perovskite PDs. Furthermore, the optimized bilayer ETL-based PDs maintained excellent stability for a 14-day period, demonstrating their potential as a viable pathway toward highly stable and reliable PDs with superior optoelectronic performance.
2. Results and Discussion
To construct a bilayer ETL, SnO
2 and ZnO layers were sequentially deposited onto an ITO substrate.
Figure 1a illustrates the fabrication process of the bilayer ETL, where SnO
2 and ZnO thin films were formed via spin-coating technique. The ZnO and SnO
2 precursor solutions were applied successively, followed by a thermal annealing process. The detailed procedure for preparing the bilayer ETL is provided in the Experimental section. To analyze the energy level alignment of each ETL layer, ultraviolet photoelectron spectroscopy (UPS) measurements were conducted.
Figure 1b presents the secondary electron cut-off and valence band (VB) region, extracted from the UPS binding energy spectra of the samples. The VB maximum (VBM) value of each ETL was determined from the onset of the binding energy spectrum where the intensity rises above the noise baseline, as indicated in each spectrum in
Figure 1b. It was observed that the work function (WF) of each ETL deposited on ITO was approximately 2.6 eV. However, the VBM values exhibited significant modulation: SnO
2 had a VBM of 3.70 eV, whereas ZnO showed a VBM of 2.46 eV, indicating a substantial vacuum level (VL) shift. This pronounced VL modulation can be attributed to the differing electron transfer capabilities of the materials, with SnO
2 demonstrating a stronger electron transfer to ITO compared to the ZnO layer. When constructing the bilayer ETL with ZnO/SnO
2, the VBM of the ITO/ETL structure was determined to be 2.50 eV, closely resembling that of a single ETL of ZnO film. This configuration enables stepwise energy level alignment at the interface between the ITO and CsPbI
2Br perovskite layers; therefore, the inverted bilayer ETL structure (SnO
2/ZnO) is not favorable for efficient electron transport and extraction from the perovskite photo-harvesting layer. The resulting stepwise band alignment effectively enhances the energy cascade matching with CsPbI
2Br films, thereby reducing electron recombination, thermionic losses, and current leakage at the perovskite/ETL interface. Furthermore, considering the wide optical bandgap of oxide semiconductors (SnO
2~3.7 eV and ZnO~3.3 eV), the conduction band (CB) of the bilayer ETL provides a favorable electron extraction channel from the CsPbI
2Br perovskite to the ITO cathode [
27]. This configuration minimizes energetic disorder across the interfaces between the ITO and perovskite layers, as depicted in
Figure 1c.
The surface topographical characteristics of the bilayer ETL in comparison to the single ETL were investigated using atomic force microscopy (AFM). As shown in
Figure 1d, the SnO
2 layer deposited on ITO exhibited a rough film with large grain structures, reflecting the surface morphology of the underlying ITO substrate. Similarly, the ITO/ZnO sample also displayed a rough surface with prominent grain formations due to the influence of the bottom ITO layer. In contrast, the bilayer ETL film demonstrated a more uniform and smoother surface, indicating that the bilayer structure effectively covers the roughness of the underlying substrate. The highly uniform and compact morphology is beneficial for reducing leakage current at the device interfaces, which in turn lowers the dark current, a critical factor in enhancing the responsivity of photodetectors. The root-mean-square (RMS) roughness values of the surfaces were measured to be 5.34 nm and 4.63 nm for the ITO/SnO
2 and ITO/ZnO samples, respectively. Notably, the ITO/SnO
2/ZnO bilayer film exhibited a significantly reduced RMS roughness of 2.37 nm. This substantial decrease in surface roughness suggests that the bilayer ETL promotes the formation of a highly compact and defect-minimized film, which can enhance electron transport efficiency and suppress non-radiative recombination. As a result, improved electrical properties are anticipated for the bilayer ETL-based devices. Since the largely aggregated nanoparticles could act as a barrier for homogeneous growth of the perovskite absorber, well-passivated surfaces of the bilayer ETLs could facilitate uniform nucleation and growth of the perovskite absorber. In addition, minimized shunt pathways and corresponding suppressed trap-assisted recombination at the interfaces could be expected in the bilayer ETL-based device [
31].
Figure 2a displays the absorption spectra of perovskite films grown on different ETL substrates. All films exhibited a characteristic absorption peak at 627 nm, typical of CsPbI
2Br, with variations in the absorption tail near 700 nm. The Tauc plots of the four samples reveal that the absorption edge shape is strongly associated with electronic transitions in the band-edge region. As shown in
Figure 2b, a reduced optical bandgap observed in the Tauc plots suggests the presence of excitonic tail states, likely due to a significant density of defects in the CsPbI
2Br films. The bandgap of the CsPbI
2Br film on bare ITO was 1.59 eV, whereas films grown on ETLs exhibited slightly higher bandgaps of approximately 1.78 eV and 1.82 eV for ITO/SnO
2 and ITO/ZnO, respectively. Notably, the bilayer ETL resulted in the widest bandgap, indicating a more stable electronic structure with lower defect density compared to films grown on single ETL-based substrates.
The crystalline structure of the perovskite films was further analyzed using X-ray diffraction (XRD). Intense diffraction peaks were observed at 2θ ~12.6° and 2θ ~29.7°, corresponding to the (1 0 0) and (2 0 0) facets of CsPbI
2Br lattices. The average crystallite size was determined using the following Debye–Scherrer equation:
where
D(hkl) is the average crystal size for the (
h k l) facet,
λ is the wavelength of X-ray (0.154 nm),
θ is the Bragg diffraction angle, and
β is the full width at half maximum (FWHM) of the peak. While the peak positions of the (1 0 0) and (2 0 0) reflections remained nearly identical across all samples, the FWHM of the (2 0 0) peak was narrowest in the CsPbI
2Br films grown on bilayer ETL substrates, suggesting improved crystallinity. Additionally, the XRD peaks were most intense in films grown on bilayer ETL-based substrates, further confirming the highest degree of crystallinity. These findings align well with the Tauc plot analysis, as shown in
Figure 2c.
To assess the impact of different ETL substrates on the topographical characteristics of perovskite films, AFM and Kelvin probe force microscopy (KPFM) were performed on CsPbI
2Br films, as shown in
Figure 2d,e. Notably, films deposited on rough substrates, including bare ITO, ITO/SnO
2, and ITO/ZnO, exhibited significant surface roughness. In contrast, perovskite films grown on the bilayer ETL demonstrated a relatively smoother morphology, attributed to the uniform surface properties of the bilayer ETL film. Additionally, KPFM measurements revealed variations in the contact potential difference across the perovskite films. For samples prepared on ITO, ITO/SnO
2, and ITO/ZnO, substantial potential barriers were observed at the grain boundaries, indicating non-uniform charge distribution. However, the bilayer ETL substrate (ITO/SnO
2/ZnO) facilitated a more homogeneous and smooth potential distribution across the perovskite film. These results further confirm that rougher bottom layers do not provide an ideal platform for the growth of highly crystalline and uniform perovskite films. Conversely, the bilayer ETL, with its compact and pinhole-free surface, enables the formation of high-quality CsPbI
2Br films, which are beneficial for efficient charge transport and reduced charge carrier recombination in perovskite PD applications.
To characterize device performance of the perovskite PDs,
n-i-p planar heterojunction devices (ITO/ETL/CsPbI
2Br/TFB/P3HT/Au) were fabricated, as illustrated in
Figure 3a. Details of the device fabrication condition are described in the Experimental section. A cross-sectional SEM image of a representative device incorporating a bilayer ETL (SnO
2/ZnO) is shown in
Figure 3b, demonstrating a well-defined layered structure. The thickness of each ETL was measured to be approximately 20–30 nm, and the total thickness of the bilayer ETL in the optimized condition is estimated to be around 50 nm, using a surface profilometer. A bilayer ETL thicker than 50 nm led to reduced device performance, characterized by an increase in dark current and a decrease in EQE, compared to the optimized condition (
Figure S1). The current density–voltage (
J–V) characteristics of the ETLs with optimized conditions, recorded in both dark conditions and under illumination (100 mW cm
−2, AM 1.5G) across a bias range of −0.2 V to 1.2 V, are presented in
Figure 3c. The photocurrent density generated from the CsPbI
2Br-based PDs was comparable regardless of the ETL structure, and the open-circuit voltage (V
OC) was approximately 1.15 V, attributed to the large band gap of the CsPbI
2Br perovskite absorber layer. For single ETL devices, the dark current density at zero bias ranged between 10
−5 and 10
−6 mA cm
−2, which further dropped to 10
−8 mA cm
−2 with the incorporation of the bilayer ETL. This reduction is primarily attributed to minimized interfacial leakage current and a decrease in perovskite film defects, both of which stem from the favorable stepwise energy level alignment and the compact surface morphology, as discussed above. The suppressed leakage current in bilayer ETL-based devices was further verified through space charge limited current (SCLC) measurements in electron-only devices, as outlined in the
Supplementary Materials (Figure S2). The trap-filled limit voltage (
VTFL) was measured at 0.70 V and 0.39 V for SnO
2- or ZnO-based devices, respectively, whereas the bilayer ETL device exhibited a substantially lower
VTFL of 0.30 V. Correspondingly, the trap density (
NT) of the single ETL-based device was calculated to be 5.37 × 10
15 cm
−3 and 2.99 × 10
15 cm
−3 for SnO
2 and ZnO, respectively, and the bilayer ETL mitigates the trap density to 2.30 × 10
15 cm
−3, confirming that the bilayer ETL effectively suppresses trap-assisted recombination and stabilizes the current flow in CsPbI
2Br-based PDs (
Table S1).
To assess the influence of different ETLs on the photodetection performance of perovskite PDs, the
EQE spectra were analyzed. As illustrated in
Figure 3d, all three devices demonstrated strong photoresponse within the 300–700 nm wavelength range. Notably, the
EQE exceeded 90% in the 450–650 nm region for three devices, suggesting that highly crystalline CsPbI
2Br perovskite films were successfully deposited on each ETL. However, the bilayer ETL exhibited the highest photoresponse characteristics among the three devices. Based on the
EQE spectra, the spectral responsivity, a key parameter reflecting the efficiency of photodetection, was calculated using the following Equation (2):
where
Jph,
I, and
EQE denote the photocurrent density, the light intensity, and the
EQE value at a given wavelength (
λ), respectively [
32]. As shown in
Figure 3e, all devices operated at zero bias and exhibited a broad UV–vis photoresponse (300–700 nm). The bilayer ETL-based PDs achieved a peak spectral responsivity of 0.45 A W
−1 at 630 nm, surpassing the 0.41 A W
−1 recorded for single ETL-based devices. Additionally, the specific detectivity (
D*), a critical metric for evaluating photodetector sensitivity, particularly under low-light conditions, was determined under the assumption that shot noise dominates. The detectivity was calculated using the following Equation (3):
where
R,
A,
e, and
Jdark represent the spectral responsivity, the effective area of the device, the elementary charge, and the dark current density [
33]. The SnO
2-based PDs achieved a detectivity of 2.5 × 10
13 Jones, while the ZnO-based devices reached 6× 10
12 Jones. Notably, the bilayer ETL-based PDs exhibited an enhanced detectivity of 9 × 10
13 Jones, significantly outperforming their single ETL counterparts at zero bias. This improvement is attributed to the substantially lower dark current in the bilayer ETL-based device.
Given the excellent photoresponse properties of the perovskite PDs with an ETL of SnO
2/ZnO, we further explored the temporal response behavior of the optimized devices employing a bilayer ETL. The time-resolved photoresponse of the passivated PD was measured under white light illumination at an intensity of 3 mW cm
−2, with no applied electrical bias and varying light on-off frequencies, as illustrated in
Figure 4a–c. As the modulation frequency increased from 5 Hz to 10 Hz and 50 Hz, the device maintained a consistent photoresponse, which presents stable and reproducible photoswitching behavior with rapid on/off response. To assess the response speed, the time-resolved photocurrent at 50 Hz was analyzed by magnifying the “
on” and “
off” states, as shown in
Figure 4d. The photocurrent displayed a sharp transition between the two states, indicating a swift reaction to light illumination. The response and recovery times, which define the time required for the dark current to reach 90% of the maximum photocurrent and to drop to 10%, were estimated to be 14 μs and 32 μs, respectively. These values demonstrate the exceptional capability of perovskite PDs to detect high-speed optical signals, making them ideal for high-performance optoelectronic applications. The rapid photoresponse performance of the bilayer ETL-based perovskite PDs can be attributed to the facilitated electron transport properties with minimized charge recombination at the ETL interfaces, as well as high-quality CsPbI
2Br perovskite films. This improvement facilitates efficient charge transport while minimizing recombination losses, thereby enhancing overall device performance.
Finally, we evaluated the stability of the bilayer ETL-based perovskite PDs, as long-term durability is crucial for practical applications. Perovskite semiconductors often face stability challenges when used as photoharvesting layers in PDs. To assess stability, the photoresponse characteristics of freshly prepared devices were measured after storage for 7 and 14 days. As shown in
Figure 5a, the responsivity of the perovskite PDs remained stable over 14 days across all ETLs; however, the specific detectivity gradually decreased during the same period. Notably, PDs employing a single ETL exhibited a continuous decline in detectivity, with almost no detectable performance after 7 days (
Figure S3 for dark J-V curves and EQE spectra for 2 weeks). In contrast, the bilayer ETL-based device retained over 80% of its original detectivity after 1 week and more than 50% after 2 weeks, indicating significantly improved long-term stability. Furthermore, when subjected to continuous heating at 80 °C, the devices preserved under ambient conditions for 2 h maintained more than 50% of their initial photoresponse (
Figures S4 and S5). This suggests that CsPbI
2Br perovskite PDs with a bilayer ETL are robust against thermal stress. Enhanced stability is primarily attributed to the bilayer ETL structure, which effectively protects the perovskite layer from moisture and oxygen during continuous heating or prolonged storage under ambient conditions. These results underscore the promising potential of bilayer ETL for improving the performance and durability of self-powered perovskite-based photodetectors, particularly for energy-efficient sensing applications such as environmental monitoring, biomedical diagnostics, and wearable technologies.