Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Interface engineering of co-evaporated Sb2Se3 solar cells using an ALD SnOx interlayer

Van-Quy Hoang ab, Jaebaek Leea, Geumha Limc, Amanat Alid, Bashiru Kadiri-Englishd, Dong-Hwan Jeona, Dae-Ho Sona, Hyo Jeong Joa, Dae-Kue Hwanga, Kee-Jeong Yanga, Eunkyung Choa, Jin-Kyu Kanga, William Jo*c, Shi-Joon Sung*ad and Dae-Hwan Kim*ad
aDivision of Energy & Environmental Technology, DGIST, Daegu, 42988, Republic of Korea. E-mail: sjsung@dgist.ac.kr; monolith@dgist.ac.kr
bCenter of Environmental Intelligence, College of Engineering and Computer Science, VinUniversity, Gia Lam District, Hanoi, 14000, Vietnam
cDepartment of Physics, Ewha Womans University, Seoul, 03760, Republic of Korea. E-mail: wmjo@ewha.ac.kr
dDepartment of Interdisciplinary Engineering, DGIST, Daegu, 42988, Republic of Korea

Received 13th March 2025 , Accepted 3rd September 2025

First published on 18th September 2025


Abstract

To compensate for the limited efficiency of co-evaporated Sb2Se3 solar cells, effective physical and chemical passivation of the interface between the Sb2Se3 absorber and the CdS buffer layer was achieved through the deposition of an ultrathin SnOx interlayer via atomic layer deposition (ALD). Due to the passivation effect of the ALD SnOx interlayer, carrier recombination at both the intra-grain and grain-boundary regions was suppressed, and Sb interdiffusion from the Sb2Se3 absorber to the cadmium sulfide (CdS) buffer was effectively blocked. Additionally, the rough surface of the co-evaporated Sb2Se3 absorber was mitigated by the conformal deposition of the ALD SnOx interlayer, reducing the statistical variation in the photovoltaic parameters of the co-evaporated Sb2Se3 solar cells. The ultrathin ALD SnOx interlayer was demonstrated to be a practical strategy for enhancing Sb2Se3 solar cell performance regardless of the absorber's morphology, achieving a substrate-type Sb2Se3 solar cell with an efficiency of 7.395% through the co-evaporation process.



Broader context

Low-carbon renewable energy technologies are essential to achieving carbon neutrality. Solar energy, an abundant resource, can replace traditional energy sources. While CuInGaSe2 (CIGSe) and kesterite Cu2ZnSn(S,Se)4 (CZTSSe) are promising thin-film solar materials, their applications are hindered by the use of rare elements and a large open-circuit voltage deficit. Antimony selenide (Sb2Se3) has emerged as an alternative due to its 1D crystal structure, high absorption coefficient, and benign grain boundaries. Unlike kesterite, Sb2Se3 offers strong anisotropic charge transport, reducing recombination losses. However, challenges such as interfacial recombination, suboptimal band alignment, and interfacial defects limit its efficiency. To enhance Sb2Se3 solar cell efficiency, researchers focus on interface passivation, doping engineering, and heterojunction optimization. Deposition techniques like close-space sublimation, hydrothermal deposition, and vapor transport deposition have shown promising results. Advancements in defect passivation and band alignment control are crucial for Sb2Se3 to become a commercially viable vacuum thin-film photovoltaic technology, ensuring device uniformity and facilitating large-area fabrication.

1. Introduction

Antimony chalcogenides, such as Sb2S3, Sb2Se3, and Sb2(S,Se)3, are well-known compound semiconductors widely used in optoelectronic applications, including photovoltaics, photodetectors, and photoelectrochemical water splitting.1–3 Due to its unique one-dimensional (1D) crystalline structure, Sb2(S,Se)3 exhibits excellent potential for high-performance photovoltaics, enabling efficient carrier transport along the 1D framework. Various deposition techniques for forming Sb2(S,Se)3 absorber layers have been reported, including hydrothermal deposition,4,5 closed-space sublimation,6–8 vapor transport deposition,9,10 thermal evaporation,11,12 rapid thermal evaporation,13 injection vapor deposition,14 sputtering with selenization,15 and chemical bath deposition (CBD).16 Although hydrothermal and CBD techniques have achieved Sb2Se3 solar cell efficiencies exceeding 10%,4,16 these solution-based processes pose challenges for large-scale production.

Among various deposition techniques, evaporation has been demonstrated to achieve high efficiency in thin-film solar cells, particularly in Cu(In,Ga)Se2 (CIGS) devices.17,18 The co-evaporation process offers several advantages, including a tailored multi-step deposition profile for enhanced efficiency, precise control over the crystalline structure of thin films, elimination of the need for an additional crystallization process, and suitability for large-area deposition. However, in Sb2Se3 solar cell research, the co-evaporation process has yielded relatively lower efficiencies compared to other techniques.11,12,19–23 Due to the unique 1D crystalline structure of Sb2Se3, controlling the morphology of Sb2Se3 absorbers remains challenging. Additionally, insufficient selenization during the co-evaporation process can degrade absorber quality and promote defect formation.24,25

To enhance the photovoltaic performance of solar cell devices, various passivation techniques are commonly employed during device fabrication.26–28 In particular, due to the relatively poor quality of co-evaporated Sb2Se3 absorber layers, physical and chemical passivation are essential for achieving high-performance Sb2Se3 solar cells. Effective passivation of the interface between the Sb2Se3 absorber and the CdS buffer layer is critical for improving device efficiency. For solar cell passivation, a metal oxide, TiO2, is widely used as a passivation layer due to its wide bandgap and specific electrical properties.6 Additionally, passivation layers play a crucial role in physically smoothing rough absorber layers.29 The conformal deposition of passivation layers at interfaces enables the formation of a more uniform cross-sectional structure, which is vital for ensuring device uniformity and facilitating large-area fabrication.

In this study, ultrathin SnOx films deposited via atomic layer deposition (ALD) were selected as the interlayer for Sb2Se3 solar cells. ALD is a well-established technique for fabricating high-quality, ultrathin films.30–32 Additionally, due to its conformal deposition capability, ALD enables the formation of uniform thin films on various surfaces, regardless of morphology. These unique characteristics make ALD highly suitable for integrating interlayers into Sb2Se3 solar cells. SnOx is a widely used metal oxide for solar cell passivation due to its wide bandgap, high optical transmittance, superior electron mobility, and chemical stability.33,34 In this work, the ALD SnOx interlayer was introduced between the Sb2Se3 absorber and the CdS buffer layer, and its role in device performance was systematically investigated.

An ultrathin ALD SnOx interlayer with a thickness of 2 nm was sufficient to inhibit carrier recombination at the interface between the Sb2Se3 absorber and the CdS buffer layer. Due to the reduced carrier recombination, the short-circuit current density (JSC) of Sb2Se3 solar cells incorporating the ALD SnOx interlayer significantly improved. Additionally, the ALD SnOx interlayer served as an effective barrier to elemental diffusion between the Sb2Se3 absorber and CdS buffer layer. Elemental interdiffusion between the absorber and buffer layer is a known factor contributing to the deterioration of the fill factor (FF) in Sb2Se3 solar cells.6,35 In this study, Sb diffusion into the CdS buffer layer was confirmed, and its suppression by the ALD SnOx interlayer was also observed. Consequently, Sb2Se3 solar cells with the ALD SnOx interlayer exhibited a pronounced enhancement in FF compared to devices without the interlayer.

In addition to its barrier effect, the ALD SnOx interlayer was found to compensate for the relatively rough surface of co-evaporated Sb2Se3 absorbers through uniform and conformal deposition. This morphological passivation effect improved the overall device structure, leading to a narrower distribution of photovoltaic parameters in Sb2Se3 solar cells. Furthermore, the impact of the ALD SnOx interlayer on charge distribution at grain boundaries in Sb2Se3 solar cells was observed. The introduction of the ALD SnOx interlayer weakened band bending at CdS grain boundaries and significantly reduced potential variations. This altered potential profile effectively inhibited carrier recombination at grain boundaries, further enhancing device performance.

The investigation of the ALD SnOx interlayer in Sb2Se3 solar cells demonstrated its effectiveness as a passivation strategy for co-evaporated Sb2Se3 solar cells, regardless of absorber morphology. Sb2Se3 solar cells incorporating the ALD SnOx interlayer achieved an efficiency of 7.395%, the highest reported for co-evaporated Sb2Se3 solar cells. Further optimization of the ALD SnOx interlayer has the potential to further enhance device performance.

2. Results and discussion

2.1 Characteristics of thin MoSe2-based devices with a passivation layer

Interfacial engineering at the absorber/buffer interface in Sb2Se3-based substrate-configuration solar cells has been extensively studied, particularly with the application of CdZnS or In2Se to modify energy band alignment and analyze their effects on surface chemistry.36,37 However, no studies have yet reported the use of an SnOx layer in substrate-configured cells for absorber surface passivation, though such advancements are expected as the technology continues to develop. In this study, the substrate configuration of Sb2Se3 thin-film solar cells was investigated, utilizing the co-evaporation method to enhance efficiency. The source-to-substrate distance was set at 50 cm to ensure a smooth Sb2Se3 film and prevent overheating. A detailed schematic of the co-evaporation process is presented in Fig. S1 (SI). The low efficiency of co-evaporated Sb2Se3 has primarily been attributed to challenges in fabricating high-quality absorbers in a high-vacuum chamber and the difficulty of controlling pyro-temperatures.

To evaluate the effect of the ALD SnOx interlayer on device performance, Sb2Se3 solar cells were fabricated with a substrate configuration of Mo/MoSe2/Sb2Se3/SnOx/CdS/i-ZnO/AZO/Al. The statistical distributions of power conversion efficiency (PCE), open-circuit voltage (VOC), short-circuit current density (JSC), and fill factor (FF) for devices with and without the ALD SnOx interlayer are presented in Fig. 1a–d. The results demonstrated that the ALD SnOx interlayer effectively enhanced the efficiency of Sb2Se3 solar cells. Additionally, co-evaporated Sb2Se3 solar cells with SnOx exhibited a narrower distribution of photovoltaic parameters, likely due to the reduced roughness of the Sb2Se3/CdS interface facilitated by the ALD SnOx interlayer. The current density–voltage (JV) curves of the best-performing devices are shown in Fig. 1e, with the corresponding photovoltaic parameters listed in Table 1. The Sb2Se3 solar cell without the ALD SnOx interlayer achieved a maximum PCE of 3.998%, with a VOC of 0.478 V, a JSC of 22.075 mA cm−2, and an FF of 37.836%. Notably, the Sb2Se3 device incorporating the ALD SnOx interlayer achieved an enhanced PCE of 6.250%, with a VOC of 0.434 V, JSC of 28.007 mA cm−2, and an FF of 51.421%. Among these parameters, JSC and FF exhibited a significant increase due to the non-ohmic space-charge-limited current, which contributed to the nonlinear shunt current in Sb2Se3 thin-film solar cells. The highly resistive transparent SnOx layer functioned as a buffer between the CdS window layer, mitigating the formation of shunt paths at the CdS/Sb2Se3 interface.38 This improvement can be attributed to the passivation and protection of the Sb2Se3 absorber layer, which reduced recombination losses and enhanced carrier transport within the device. These findings demonstrate the effectiveness of the ALD SnOx interlayer in improving the efficiency of co-evaporated Sb2Se3 solar cells.


image file: d5el00031a-f1.tif
Fig. 1 (a–d) Box plots of solar cell parameters for the tested devices. (e) Current density–voltage (JV) curves of the highest-efficiency device within each group. (f) EQE spectrum of the best-performing device based on the standard AM 1.5G solar spectrum. (g) The bandgap calculated from EQE spectra. (h) Urbach energy values derived from EQE spectra for the control-Sb2Se3 and SnOx–Sb2Se3 solar cells, respectively.
Table 1 Photovoltaic parameters of co-evaporated Sb2Se3 solar cells with and without the ALD SnOx interlayer, measured under AM 1.5G illumination
Samples VOC (V) JSC (mA cm−2) FF (%) PCE (%) Eg (eV) EA (eV) EA/Eg (%) G (mS cm−2) R (Ω cm2) A J0 (mA cm−2)
W/o SnOx Average 0.470 20.872 35.468 3.500              
StDev 0.008 2.157 2.254 0.529              
Champion 0.478 22.075 37.836 3.998 1.17 1.00 85.47 18.58 0.26 6.10 0.638
w/SnOx Average 0.422 27.553 50.911 5.926              
StDev 0.006 0.657 0.561 0.169              
Champion 0.434 28.007 51.421 6.250 1.17 1.04 88.89 7.30 2.43 2.08 0.027


The external quantum efficiency (EQE) data in Fig. 1f indicate that both devices exhibited a photoelectronic response from 300 to 1100 nm, with the bandgap of Sb2Se3 devices–with and without the ALD SnOx interlayer. Notably, the two EQE curves displayed identical onset wavelengths, suggesting that both cases of Sb2Se3 absorbers had equivalent bandgaps. Previous studies have demonstrated that introducing a high-resistivity SnO2 layer via pulsed laser deposition in the superstrate configuration reduces reflection at the front electrode and enhances carrier collection efficiency.39 Therefore, the Sb2Se3 device with the ALD SnOx interlayer exhibited increased EQE in the long-wavelength region (600–1100 nm). We observed that the integrated current densities derived from the EQE curves were 22.45 and 24.49 mA cm−2 for the control and modified devices, respectively. These values were lower than the JSC measured from the corresponding JV curves. This discrepancy in JSC between the EQE and JV curves has also been reported in previous studies on Sb2Se3 solar cells, which can be attributed to the deep levels in the junction region.16,40 The photogenerated carriers can recombine at these centers at low light levels (EQE measurement), whereas some of these centers are more likely to be occupied due to photoexcitation in high light intensity (under AM 1.5G illumination at 100 mW cm−2).

The indirect bandgaps of devices without and with the SnOx interlayer were determined to be 1.17 eV, by extrapolating the linear region of the (hv × ln(1 − EQE))1/2 versus hv plots to the horizontal photon energy axis (Fig. 1g).41 The observed bandgap variation was consistent with the chemical composition, as a higher Se content in the co-evaporated Sb2Se3 thin film resulted in a lower bandgap value, consistent with our density functional theory (DFT) calculations in the recent discovery.42 The ALD SnOx interlayer likely functioned as an efficient electron transport channel between the Sb2Se3 absorber and CdS buffer layers, helping electron movement across the Sb2Se3/CdS junction. Urbach energy (EU), a metric used to quantify the extent of the band tail effect, was derived from the ln(1 − EQE)) versus energy curve, as shown in Fig. 1h. Upon incorporating the ALD SnOx interlayer, EU significantly decreased from 24.90 meV to 21.05 meV, suggesting that recombination near the Sb2Se3/CdS junction was mitigated due to the passivation of detrimental defects, a topic further discussed in later sections.

The fabrication techniques for co-evaporated Sb2Se3 thin-film solar cells are illustrated in Fig. 2a and described in the “Experimental” section. The surface and cross-sectional morphologies of the as-prepared films were analyzed using scanning electron microscopy (SEM), revealing an estimated Sb2Se3 grain size of approximately 290 nm. Fig. 2b–g present top-view SEM images of the co-evaporated Sb2Se3 absorbers and CdS buffer layers deposited on Sb2Se3, as well as cross-sectional SEM images of the co-evaporated Sb2Se3 devices with and without the 2 nm ALD SnOx interlayer. The SEM images reveal a compact and uniform CdS grain coverage with no visible pinholes between the grains and strong adhesion to the Sb2Se3 absorber surface. However, the surface roughness is expected to be high, with wide grain boundaries clear in the surface SEM images. These deep, valley-like grain boundaries could hinder uniform CdS formation. The application of an ALD SnOx interlayer may mitigate this issue by improving coverage and promoting more uniform CdS growth in these regions. Notably, CdS buffer layer morphology undergoes significant changes with increasing ALD SnOx interlayer thickness from 0 to 5 nm, as evidenced by the SEM images in Fig. S2, SI. SnOx is expected to play a crucial role in bridging narrow and deep features that the CdS layer struggles to cover uniformly. However, our findings indicate that CdS exhibits inferior growth on the oxide surface compared to the chalcogenide surface. Increasing the thickness of the ALD SnOx interlayer altered the morphology of the CdS buffer layers, leading to the formation of discrete nanoparticles. Therefore, to ensure consistent CdS coverage, the SnOx layer was constrained to an optimal thickness of 2 nm. Furthermore, device performance was highly dependent on the ALD SnOx interlayer thickness. Notably, only ultrathin ALD SnOx interlayers enhanced performance, with an optimal thickness of 2 nm, while thicker interlayers (>2 nm) resulted in a decline in VOC (Fig. S3, SI and Table S1).


image file: d5el00031a-f2.tif
Fig. 2 (a) Schematic representation of the fabrication process for co-evaporated Sb2Se3 solar cells with glass/Mo/MoSe2/Sb2Se3/SnOx/CdS/ZnO/AZO/Al architecture. (b–d) Top-view SEM images of the Sb2Se3 layer, CdS/Sb2Se3 layer, and cross-sectional SEM images of devices without the ALD SnOx interlayer. (e–g) Corresponding SEM images of devices incorporating the ALD SnOx interlayer.

The crystal structure and phase purity of the co-evaporated Sb2Se3 thin film were analyzed using X-ray diffraction (XRD), as shown in Fig. 3a. The Sb2Se3 exhibited an orthorhombic crystal structure, classified under the space group Pbnm (JCPDS 00-015-0861), with no detectable impurity phases. Notably, only strong (hk1) diffraction peaks were observed in the XRD pattern, indicating a preferred orientation along the c-axis. The intensity ratios I101/I221 and I002/I221 were 0.13 and 0.34, respectively, confirming the (221)-preferred orientation of the co-evaporated Sb2Se3 thin film. To quantify differences in crystalline orientations, the texture coefficient (TC) of the diffraction peaks was calculated using the equation: image file: d5el00031a-t1.tif where I(hkl) represents the observed peak intensity of the (hkl) plane, I0(hkl) is the corresponding standard XRD intensity, and N is the total number of reflections considered for the calculation.43 A higher TC value for a given diffraction peak indicates a stronger preferred orientation along that direction. As shown in Fig. S4, SI, the co-evaporated Sb2Se3 thin film deposited at 315 °C exhibited a (hk1) preferred orientation, particularly along (221) and (211). The cross-sectional high-resolution transmission electron microscopy (HRTEM) image revealed a flat and uniform morphology of the co-evaporated Sb2Se3 thin film (Fig. 3b). Additionally, the interplanar d-spacing of 0.523 nm corresponded to the (210) planes of orthorhombic Sb2Se3, as shown in Fig. S5, SI. This value was consistent with the d-spacing observed in one-dimensional (1D) single-crystalline Sb2Se3 nanostructures synthesized by other methods.6,44,45


image file: d5el00031a-f3.tif
Fig. 3 (a) XRD pattern and (b) HRTEM image of the co-evaporated Sb2Se3 absorber on the MoSe2/Mo substrate. (c and d) Energy band structure of the Sb2Se3 absorber and CdS buffer without and with the ALD SnOx interlayer, respectively. (e–g) X-ray photoelectron spectroscopy (XPS) spectra of the co-evaporated Sb2Se3 absorber covered with the ALD SnOx interlayer at various etching depths (320–820 s).

The application of an ALD SnOx interlayer on the Sb2Se3 absorber via ALD was found to improve JSC, FF, and overall power conversion efficiency. To better understand this improvement, an energy band diagram was constructed based on ultraviolet photoelectron spectroscopy (UPS) analysis, comparing the Sb2Se3/CdS and Sb2Se3/SnOx/CdS heterojunctions. The results revealed that the Sb2Se3/CdS interface exhibited a weak spike-like conduction band offset (CBO) of ΔEC = 0.01 eV, whereas the introduction of the ALD SnOx interlayer led to a more pronounced spike CBO of ΔEC (Sb2Se3–SnOx) = 0.06 eV in the Sb2Se3/SnO/CdS structure. The increased spike CBO effectively suppresses the backflow of electrons, reducing interfacial carrier recombination. In the Sb2Se3/CdS heterojunction, the low CBO allowed electrons to transfer easily into CdS; however, it also increased the possibilities of electron backflow and recombination. In contrast, with the SnOx interlayer, the higher spike CBO restricts electron backflow while maintaining efficient electron transport toward CdS, leading to improved charge transport and increased JSC. However, an excessively large spike CBO can also increase transport resistance, impeding electron injection. This effect can result in charge accumulation, enhanced interfacial recombination, and a reduction in the VOC. As the CBO increases with the insertion of SnOx, an additional energy barrier is formed at the interface, potentially slowing electron transport. Therefore, in spike-type band alignment, it is crucial to balance electron backflow suppression and forward transport resistance minimization. Experimental results indicate that increasing the SnOx thickness correlates with a reduction in VOC (Fig. S3, SI). Thus, for optimal device performance, fine-tuning the SnOx thickness is expected to be an effective strategy for preventing VOC degradation while maintaining JSC improvement. Additionally, the insertion of SnOx significantly affects the valence band offset (VBO), enhancing the hole-blocking effect. The Sb2Se3/SnOx interface exhibited a substantial increase in VBO to −2.67 eV, effectively preventing hole backflow and reducing hole recombination at the interface (Fig. S6, SI). This strengthened VBO contributes to an increase in JSC and FF, further enhancing device efficiency. As a result, introducing an ALD SnOx interlayer improves device performance by forming a spike CBO that effectively blocks electron backflow and enhances JSC. However, an excessive CBO increase can lead to higher transport resistance and a subsequent decrease in VOC. To mitigate this, further optimization of the CBO and precise thickness control of SnOx are necessary. Furthermore, the increased VBO resulting from SnOx insertion strengthens hole blocking, reducing recombination and positively impacting JSC and FF. These findings demonstrate that SnOx incorporation is an effective strategy for improving Sb2Se3-based solar cell performance, highlighting the need for further design refinements to achieve optimal efficiency.

The XPS results at various etching depths (Fig. 3e–g) further support the penetration of SnOx into the Sb2Se3 absorber. The presence of Sn within the vacant regions of the Sb2Se3 layer, facilitated by the exceptional coverage of ALD SnOx, indicates that SnOx effectively passivated the entire Sb2Se3 layer, reducing the probability of shunt path formation. Fig. 3e presents the Sn content within the Sb2Se3 absorber at etching depths ranging from 320 s to 820 s. As the etching depth increased, the Sn ion content gradually decreased until reaching the Mo substrate, with the total Sb2Se3 thickness estimated at approximately 700 nm. Notably, Sn signals remained detectable even after prolonged etching, providing evidence of the deep penetration achieved by the ALD process, consistent with TEM and SEM analyses. Additionally, the presence of various surface nanorods may have contributed to Sn signal detection during depth profiling, further reinforcing the effectiveness of the coverage. The two peaks at 54.6 and 53.7 eV (Fig. 3f) correspond to Se 3d3/2 and Se 3d5/2 of Sb2Se3, respectively. Similarly, Fig. 3g shows peaks centered at high binding energies of 538.7 and 529.7 eV, attributed to Sb 3d3/2 and Sb 3d5/2 of Sb2Se3.20,46,47 Moreover, the XPS spectra of CdS buffers on Sb2Se3 and Sb2Se3/SnOx exhibited no significant peak shifts or the appearance of Cd and S after a 10 s etching time (Fig. S7, SI).

The ALD SnOx interlayer would likely affect charge transport between Sb2Se3 and CdS, potentially resulting in an alteration of the electrical potential distribution. The Kelvin probe force microscopy (KPFM) method was employed to investigate potential variations within the CdS layer by measuring surface potential distribution. The topography and local potential mapping results obtained on the CdS surface with and without SnOx are presented in Fig. 4a. The histogram of contact potential distribution (VCPD) was extracted from the mapping data (Fig. 4b). With the introduction of the ALD SnOx interlayer, the FWHM of the VCPD distribution decreased from 26 mV to 21.5 mV, showing the formation of a homogeneous potential distribution in the CdS buffer layer. The band bending arising from the irregular potential distribution could exert a force on electrons, the major charge carriers in the CdS layer. Therefore, without the SnOx layer, localized forces induced near the GBs could generate lateral electron flow, potentially disrupting carrier transport toward the TCO layer.48 Based on the results, the role of the ALD SnOx interlayer could be suggested as a capping layer, producing more uniform contact between the absorber and buffer layers by mitigating band bending at GBs.49 This consequently enables the formation of a more homogeneous and broader p–n junction area, enhancing electron transport across the interface.


image file: d5el00031a-f4.tif
Fig. 4 (a) KPFM images of the CdS buffer layer and (b) histogram of the potential distribution of the CdS surface with or without the ALD SnOx interlayer and (c) band structure diagram of CdS and SnOx from UPS measurements.

To further prove the enhanced charge transport facilitated by the ALD SnOx interlayer, the band structure of CdS was examined in relation to the presence of SnOx. We suggest an energy band alignment, as shown in Fig. 4c and Fig. S8, SI, constructed based on ultraviolet-visible (UV-vis) and ultraviolet photoelectron spectroscopy (UPS) measurements. The bandgap of CdS, calculated from UV-vis spectra, was determined to be 2.4 eV for both samples with and without SnOx. A reduction in the energy difference between ECBM and EF (|ECBMEF|) of 50 meV was observed in the CdS with SnOx compared to CdS without SnOx. The shift of the Fermi level toward the conduction band minimum (CBM) indicates an increase in electron concentration with the introduction of SnOx.50 This confirms improved charge separation at the junction between the Sb2Se3 and CdS layers, consistent with the KPFM results. Moreover, a decrease in the work function of 300 meV was observed in CdS with SnOx. The modified energy band alignment is energetically favorable, inducing a strong built-in potential at the CdS/TCO interface under short-circuit conditions.51 This configuration can facilitate carrier extraction and contribute to JSC enhancement.52

2.2 Interdiffusion at the absorber–buffer interface

To examine the effect of the ALD SnOx interlayer on the microstructure of co-evaporated Sb2Se3 solar cells, various analytical techniques were employed. High-resolution TEM images of the devices (Fig. S9, SI) provide detailed insights into the microstructural evolution and interfacial characteristics, particularly at the CdS/Sb2Se3 and Mo/MoSe2/Sb2Se3 interface regions. The introduction of the ALD SnOx interlayer improved the Sb2Se3 device structure, yielding a smoother interface and enhancing the adhesion between the Sb2Se3 and CdS layers. For a more precise microstructural analysis, high-angle annular dark-field (HAADF) imaging and energy-dispersive X-ray spectroscopy (EDS) elemental mapping were performed. Fig. S10 and S11, SI present the HAADF images and EDS elemental maps of the Sb2Se3/CdS interfacial region, both with and without the ALD SnOx interlayer. In both cases, Sb and Se elements exhibited uniform distribution throughout the Sb2Se3 absorber layer, confirming the homogeneous composition of the material. However, HAADF images of each Sb2Se3 device revealed differences in the distribution of layers. The Sb2Se3 device with the ALD SnOx interlayer exhibited improved uniformity of the CdS buffer and AZO transparent conducting oxide (TCO), likely due to morphological changes in the CdS buffer on the ALD SnOx interlayer. Additionally, EDS elemental composition maps showed no thermally induced interdiffusion, showing that the co-evaporated Sb2Se3 absorber and CdS buffer remained free of cross-contamination during the chemical bath deposition (CBD) process of the CdS buffer layer. The EDX line scan further confirmed the absence of thermal diffusion in the heterojunction.

To precisely examine the coverage of the CdS buffer layer on the Sb2Se3 absorber and the interdiffusion of elements at the CdS/Sb2Se3 interface, HRTEM-EDS analysis was performed. As shown in Fig. 5a–c, spatial elemental mapping of Sb revealed variations in color intensity, indicating notable interfacial interdiffusion of Sb into the CdS buffer layer, whereas no significant interdiffusion of Se was observed (Fig. S12, SI). The presence of Sb in the CdS buffer layer likely resulted from the dissolution of Sb2Se3 in the alkaline ammonia solution, leading to its reaction with NH4+ during the CBD process.6,53–55 However, Sb interdiffusion was mitigated by the ALD SnOx interlayer, which blocked direct contact between Sb2Se3 and NH4+ at the Sb2Se3/CdS interface (Fig. 5d–f). EDS line scans of Zn, Cd, O, and Sb across the interfaces are presented in Fig. 5g, h. The line scans show a small step increase in Sb over a distance of 30–35 nm in the CdS layer of the control sample, as confirmed by the slight accumulation of Sb at the highest intensity of the Cd signal before a significant increase in the Sb signal. In contrast, no step-function increase was observed for the SnOx-based device, convincingly suggesting that the SnOx layer is effective in reducing or preventing the diffusion of Sb into the CdS layer or near the Sb2Se3/CdS interlayer.


image file: d5el00031a-f5.tif
Fig. 5 HRTEM-EDS images of Sb2Se3 devices (a–c) without an ALD SnOx interlayer and (d–f) with an ALD SnOx interlayer. EDX line scans of Zn, Cd, O and Sb across (g) the TCO/CdS/Sb2Se3 interface and (h) the TCO/CdS/SnOx/Sb2Se3 interface. TOF-SIMS depth profile showing elemental mapping of devices (i) without an ALD SnOx interlayer and (j) with an ALD SnOx interlayer. TOF-SIMS 3D tomography of the Sn element: (k) front view and (l) 3D view focusing on the Sb2Se3/SnOx interlayer.

To verify the presence and interdiffusion of Sb at the Sb2Se3 device interface, time-of-flight secondary ion mass spectrometry (TOF-SIMS) depth profiling was performed on Sb2Se3 devices with and without the ALD SnOx interlayer (Fig. 5i and j). In the presence of the ALD SnOx interlayer, an intense Sn ion signal was detected at the interface between the Sb2Se3 absorber and CdS buffer. However, no significant difference in Sb ion concentration was observed due to the low interdiffusion level. These results align with the widespread diffusion pattern identified in the 3D tomography analysis of Sn ions, which were predominantly detected at the Sb2Se3/CdS junction along with the absorber thickness (Fig. 5k and l). The TOF-SIMS 3D tomography results for different elements in Sb2Se3 devices are shown in Fig. S13, SI. The 3D-rendered overlay of elements further clarifies the position of each component, confirming that Sn ions are incorporated into the Sb2Se3 absorber region.

2.3 Carrier transport dynamics and defect characterization

The JV curves of the solar cells (Fig. 6a) clearly demonstrate the significant impact of the ALD SnOx interlayer on the photovoltaic performance of co-evaporated Sb2Se3 solar cells. The Sb2Se3 device without the SnOx interlayer exhibited lower power conversion efficiency (PCE), whereas the device with the ALD SnOx interlayer achieved a higher PCE. The saturation current density and diode ideality factor were extracted using the single–diode equation: image file: d5el00031a-t2.tif where J0, G, V, q, A, k, T, and Rs represent the saturation current density, shunt conductance, voltage, electron charge, diode ideality factor, Boltzmann's constant, temperature, and series resistance, respectively.53 Notably, significant differences were primarily observed in the fill factor (FF) and short-circuit current density (JSC) by analyzing the values of A and J0 obtained from fitting the ln(J + JSCGV) versus (VRJ) curves. Specifically, dV/dJ versus (J + JSC)−1 and dJ/dV versus voltage (V) plots (Fig. 6b and c) were used to evaluate the series resistance and average shunt conductance, respectively. Notably, deep defects at grain boundaries (GBs) can create alternative paths for electron and hole transport, increasing shunt conductance while maintaining a relatively low series resistance. This phenomenon was observed in the Sb2Se3 device without the SnOx interlayer, resulting in a higher shunt conductance of 18.58 mS cm−2, as shown in Table 1. In contrast, the Sb2Se3 device with the ALD SnOx interlayer exhibited a lower shunt conductance of 7.30 mS cm−2 due to defect passivation.39 As shown in Fig. 6d, the loss mechanisms and variations between the ideal and actual performance of the SnOx-based devices were analyzed. The values of J0 and A decreased from 0.638 to 0.027 and from 6.10 to 2.08, respectively, due to reduced recombination at the Sb2Se3/CdS interface, attributed to the passivation effect of the ALD SnOx interlayer. The presence of the ALD SnOx interlayer resulted in a lower A, indicating reduced recombination losses from interface defects. Additionally, the reduced J0 in the Sb2Se3 device with SnOx suggests a lower defect density and diminished carrier recombination.
image file: d5el00031a-f6.tif
Fig. 6 Characteristic behavior of the best-performing devices: (a) JV curves, (b) derivative of dV/dJ under forward bias with fitting used to determine the series resistance and diode ideality factor, (c) derivative of dJ/dV for shunt characteristic analysis, and (d) ln(J + JSCGV) versus (VRJ)) curves. IVT measurements of Sb2Se3 devices under light and dark conditions (120 K to 300 K), showing the determination of series and shunt resistance (e and f) without the ALD SnOx interlayer and (g and h) with the ALD SnOx interlayer. (i) Temperature dependence of VOC. Capacitance–voltage (CV) and drive-level capacitance profiling (DLCP) of devices (j) without the ALD SnOx interlayer and (k) with the ALD SnOx interlayer.

To better understand the impact of the SnOx interlayer on device performance, we investigated the electrical properties of solar cells over a temperature range of 120 K to 300 K. The electrical characteristics of solar cells are intricately linked with photovoltaic performance and serve as an effective means of examining carrier transport and recombination behavior. Fig. 6e–h presents the temperature-dependent IVT curves under dark and illuminated conditions for devices with and without the ALD SnOx interlayer. IVT measurements were conducted for all samples at temperatures ranging from 120 K to 300 K. A more pronounced current-blocking effect on injection current was observed at higher temperatures in the Sb2Se3 device with the ALD SnOx interlayer compared to the device without SnOx.56 Fig. 6i presents the temperature dependence of VOC, which was analyzed to investigate the recombination characteristics at the Sb2Se3/CdS interface. The activation energy (EA) was figured out by linearly extrapolating the data within the measured VOC temperature range (120 K < T < 300 K), with EA corresponding to the value at T = 0 K. The EA/Eg ratios were 85.47% for the Sb2Se3 device without SnOx and 88.89% for the device with the ALD SnOx interlayer. An increase in EA/Eg correlates with improved photovoltaic performance due to reduced recombination losses from defects at the Sb2Se3/CdS interface (Table 1). The recombination mechanism follows the Shockley–Read–Hall (SRH) process in the space-charge region (SCR) when EA is close to Eg. In contrast, a lower EA/Eg suggests recombination at the absorber–buffer interface.57 Therefore, the ALD SnOx interlayer functioned as a surface passivation layer, mitigating recombination losses at the Sb2Se3/CdS interface.58

Fig. 6j and k present carrier density as a function of depletion width, obtained from capacitance–voltage (CV) profiling and drive-level capacitance profiling (DLCP) measurements, to examine the electrical properties near the Sb2Se3/CdS interface. The depletion width (Wd) was determined from DLCP measurements by evaluating the capacitance at zero bias and applying the formula Wd = ε0εA/C, where C is the measured capacitance for each DC bias, A is the device area (0.185 cm−2), and ε is the dielectric constant of the absorber (fixed at 14.3 in this study based on prior assumptions).41,59 Carrier density N and Wd were extracted from CV profiling and DLCP measurements. The calculated carrier density (NCV), depletion width (Wd), interface trap density (NIT), and bulk density (NDLCP) are summarized in Table 2. The NDLCP and NCV values for the Sb2Se3 device without the ALD SnOx interlayer were determined to be approximately 7.02 × 1016 cm−3 and 14.90 × 1016 cm−3, respectively. With the ALD SnOx interlayer, these values decreased to approximately 3.59 × 1016 cm−3 and 7.41 × 1016 cm−3, respectively. The high density of traps at the interface indicates an increased recombination rate at the Sb2Se3/CdS interface, which negatively affects device performance due to significant bulk defects in the Sb2Se3 absorber. Both Sb2Se3 devices, with and without the ALD SnOx interlayer, exhibited a relatively wide depletion region. The interface trap density (NIT), defined as the difference between NCV and NDLCP at zero bias, was 7.88 × 1016 for the device without the interlayer and 3.76 × 1016 cm−3 for the device with the interlayer. This reduction is consistent with the mitigation of interface recombination observed in IVT measurements. The lower NIT in the ALD SnOx device confirms the effective passivation of interface traps at the Sb2Se3/CdS junction. The interface defect density was reduced by a factor of 2.09 in the SnOx-treated device, further corroborating that SnOx effectively suppresses recombination centers, leading to improved device performance.

Table 2 Carrier density parameters of Sb2Se3 solar cells with and without the ALD SnOx interlayer
Samples NCV (cm−3) Wd (μm) NIT (cm−3) NDLCP (cm−3)
W/o SnOx 14.90 × 1016 0.167 7.88 × 1016 7.02 × 1016
w/SnOx 7.41 × 1016 0.141 3.76 × 1016 3.59 × 1016


2.4 Function of SnOx for the nanorod array Sb2Se3/CdS interface

To investigate the coverage of CdS buffer layers on the rough surface of Sb2Se3 absorbers and the interdiffusion of elements at the CdS/Sb2Se3 interface, we intentionally fabricated nanostructured Sb2Se3 absorbers by co-evaporating Sb2Se3 onto a 30 nm-thick MoSe2 layer, which is known to promote the formation of an Sb2Se3 nanorod array12 (Fig. S14, SI). As shown in Fig. 7a and b, the introduction of the 30 nm-thick MoSe2 layer facilitated the formation of an Sb2Se3 nanorod array on a flat Sb2Se3 film. Fig. 7c illustrates the schematic of a nanostructured Sb2Se3 solar cell incorporating the thick MoSe2 layer. Unlike the Sb2Se3 device with a 5 nm MoSe2 layer, depositing a CdS buffer layer with good coverage on the nanorod array of Sb2Se3 on 30 nm MoSe2 was challenging. Incomplete coverage could lead to shunt leakage due to local discontinuities or pinholes at the Sb2Se3/CdS junction. Therefore, the nanostructured Sb2Se3 absorber required an effective passivation and protection layer to ensure strong adhesion and well-defined junction formation between the Sb2Se3 absorber and the CdS buffer layer. Fig. S15, SI presents TEM-EDS images of co-evaporated Sb2Se3 solar cells fabricated with thick MoSe2 layers. To address this issue, the ultrathin ALD SnOx interlayer improved contact between the CdS buffer layer and the rough Sb2Se3 nanorod array absorber, thereby reducing shunt leakage and potentially enhancing JSC. The HRTEM images (Fig. 7d and e) clearly showed a 30 nm-thick MoSe2 layer at the interface between Mo and Sb2Se3. Further analysis of Sb2Se3 films grown on 30 nm-thick MoSe2 revealed that the CdS buffer layer not only covered the top of the absorber but also penetrated the spaces, conformally coating the sidewalls and valleys (Fig. 7f). Although a high density of nanorods was observed in the Sb2Se3 absorber grown on thick MoSe2 (as seen in the top SEM image), a region with a flat and compact bottom Sb2Se3 absorber was selected for cross-sectional HRTEM analysis to emphasize the role of the MoSe2 layer in facilitating bulk Sb2Se3 formation. The location and size of voids within the absorber varied depending on the MoSe2 thickness. As previously observed, small voids formed at the bottom of the absorber in the thin MoSe2-based structure, whereas voids and larger spaces were distributed throughout the absorber in the thick MoSe2-based structure. Thus, the ALD passivation layer played a crucial role in protecting the absorber in deep and inaccessible regions during the CdS deposition process, thereby preventing shunt effects.
image file: d5el00031a-f7.tif
Fig. 7 (a) Top-view and (b) cross-sectional SEM images of co-evaporated Sb2Se3 nanorod arrays on a 30 nm-thick MoSe2/Mo substrate. (c) Schematic illustration of the co-evaporated Sb2Se3 nanorod array solar cell device. High-resolution cross-sectional TEM (HRTEM) images of (d) the best-performing device, (e) an enlarged HRTEM image of the MoSe2/Sb2Se3 interface (highlighted rectangular region), and (f) the Sb2Se3/CdS interface. (g) EDX line scans of the device and (h) EDX line scans of the Se/Sb ratio in co-evaporated Sb2Se3 devices with varying MoSe2 thicknesses. (i–l) Box plots of the solar cell parameters for co-evaporated Sb2Se3 nanorod array devices utilizing the thick MoSe2/Mo substrate.

Fig. 7g and h presents an EDX line scan of elemental distribution and the Se/Sb ratio (mol%) in Sb2Se3 solar cell devices incorporating the ALD SnOx interlayer with a thick MoSe2 layer. The introduction of a 30 nm-thick MoSe2 layer induced morphological changes in the Sb2Se3 absorber while maintaining its chemical composition, as compared to a 5 nm-thick MoSe2 layer. This finding indicates that the optical properties and energy band structure of the Sb2Se3 absorber remained unchanged. To assess the charge extraction and passivation effects of the ALD SnOx interlayer on device performance with a 30 nm-thick MoSe2 layer, Sb2Se3 solar cell devices were fabricated with the structure Mo/MoSe2 (30 nm)/Sb2Se3/SnOx/ZnO/AZO/Al electrode. As shown in Fig. 7i–l, the best-performing control device exhibited a PCE of 1.971%, with a VOC of 0.454 V, a JSC of 15.381 mA cm−2, and an FF of 28.200%. In comparison, Sb2Se3 devices incorporating the ALD SnOx interlayer achieved a maximum PCE of 5.627%, with a VOC of 0.414 V, a JSC of 27.214 mA cm−2, and an FF of 49.880% (corresponding values are provided in Table S2). The diode properties of Sb2Se3 solar cells based on the thick MoSe2 layer are shown in Fig. S16 (SI), showing an improvement in diode behavior due to the passivation layer. The incorporation of SnOx at the interface significantly enhanced device performance consistency, even when using inherently rough Sb2Se3 films, leading to a narrower distribution of PV parameters. Additionally, Sb2Se3 solar cells with a 30 nm-thick MoSe2 layer showed a reduced variation in photovoltaic parameters, aligning with the previously observed results for Sb2Se3 devices using a 5 nm-thick MoSe2 layer.

The Sb2Se3 solar cells fabricated on a MoSe2/Mo substrate at 410 °C for 10 minutes were analyzed by an external certified laboratory to confirm our in-house PCE measurements. The JV and EQE curves of the best-performing devices are shown in Fig. 8, with their photovoltaic parameters summarized in Fig. 8a. The device showed an active area efficiency of 7.395% using an anti-reflective magnesium fluoride (MgF2) coating, with an FF of 51.097%, JSC of 32.457 mA cm−2, and VOC of 0.445 V. The highest-efficiency Sb2Se3 device provided a relatively low J0 value, showing that the passivation layer effectively reduced photo-generated charge carrier losses (Fig. S17, SI). For further comparison, Fig. 8c and Table 3 summarize the PCEs of this work alongside previously reported Sb2Se3 solar cells fabricated using evaporation-based methods. Notably, this study achieved the highest efficiency among (co)evaporation methods using a substrate configuration. The ultra-high vacuum environment used in the (co)evaporation method, often reaching pressures below 10−7 torr, ensures superior interface integrity and significantly mitigates defect formation. This results in higher material quality compared to other vacuum-based techniques such as close-spaced sublimation (CSS) or atomic layer deposition (ALD), which typically run at lower vacuum levels (10−2 to 10−5 torr).


image file: d5el00031a-f8.tif
Fig. 8 (a) Current density–voltage (JV) curve and (b) external quantum efficiency (EQE) spectra of the highest-performing Sb2Se3 solar cells fabricated with a passivation layer. (c) Efficiency distribution of Sb2Se3 heterojunction solar cells produced using evaporation and co-evaporation methods. An anti-reflective (AR) coating was applied to the device.
Table 3 Detailed photovoltaic performance parameters of high efficiency (co)evaporated Sb2Se3 devices
Configuration PCE (%) VOC (mV) JSC (mA cm−2) FF (%) Eg (eV) Date Institute
Substrate 1.47 0.407 12.11 30.00   2018 CNU60
Superstrate 1.90 0.300 13.20 48.00   2014 HUST19
Superstrate 2.10 0.354 17.80 33.50 1.20 2014 HUST20
Substrate 3.38 0.362 18.54 50.39   2023 SRMIST61
Superstrate 3.47 0.364 23.14 41.26   2016 HB22
Superstrate 3.50 0.339 20.70 49.00   2019 LAPS62
Superstrate 3.60 0.352 23.50 44.20   2021 LAPS23
Superstrate 3.70 0.335 24.40 46.80 1.18 2014 HUST63
Substrate 4.25 0.420 17.11 58.15   2016 HB21
Substrate 4.51 0.370 25.39 47.24   2019 DGIST11
Superstrate 5.52 0.367 26.44 56.95 1.25 2023 USTC64
Substrate 5.63 0.430 27.43 47.35   2021 DGIST12
Superstrate 6.24 0.380 28.10 59.10 1.13 2020 CUAS65
Substrate 7.395 0.445 32.457 51.097 1.17 2024 This work


3. Conclusions

We reported an effective passivation method for fabricating high-efficiency Sb2Se3 solar cells using the co-evaporation process. An ultrathin SnOx interlayer, deposited by ALD, was introduced between the Sb2Se3 absorber and the CdS buffer layer, resulting in improved JSC, FF, and overall device efficiency. The ALD SnOx interlayer effectively suppressed carrier recombination at the grain interiors and grain boundaries of the Sb2Se3 absorber while serving as a barrier against Sb element interdiffusion. This passivation of the Sb2Se3/CdS interface led to a significant enhancement in the performance of Sb2Se3 solar cells, regardless of absorber morphology. Additionally, the conformal deposition of the ALD SnOx interlayer resulted in a uniform device structure, enhancing the reproducibility of co-evaporated Sb2Se3 solar cells. As a result, Sb2Se3 solar cells with the ALD SnOx interlayer achieved an efficiency of 7.395%, the highest recorded for co-evaporated Sb2Se3 devices. Further optimization of the ALD SnOx interlayer could lead to performance improvements.

4. Experimental

4.1 Preparation of the ALD SnOx layer

The SnOx deposition was conducted using a custom-built ALD system at 100 °C, with TDMASn as the precursor (supported at 45 °C) and O3 as the oxidizer. The base pressure was kept at 3 × 10−2 torr, with an added Ar flow of 20 sccm under working conditions. During deposition, ALD-SnOx films on Sb2Se3 were heated using a hot trap at 350 °C. The growth rate of SnOx was determined to be 0.14 nm per cycle, as confirmed by TEM analysis, which measured a 20 nm thickness on a Si wafer after 140 cycles. The average O/Sn ratio was calculated to be 2.781 based on TEM-EDS measurements (Fig. S18, SI).

4.2 Fabrication of the Sb2Se3 device

Solar cells were fabricated as a multilayer stack consisting of soda lime glass (SLG)/Mo/Sb2Se3/SnOx/CdS/i-ZnO/ZnO:Al. The Mo back electrode was deposited onto the SLG substrate via DC sputtering. A thin MoSe2 layer (∼5 nm) was grown at Tsub = 600 °C using co-evaporation equipment, while a thicker MoSe2 layer (30 nm) was deposited at Tsub = 430 °C for 10 min using a two-zone furnace system. The morphology of the three different Mo substrates is illustrated in Fig. S19 (SI). The Sb2Se3 absorber was deposited via co-evaporation, using a lower-temperature growth step with Tsource = 630 °C and Tsub = 315 °C. The CdS buffer layer was later deposited via chemical bath deposition (CBD) at 65 °C for 11.5 min. The device was completed by sputtering a 50 nm i-ZnO layer followed by a 350 nm ZnO:Al layer (sheet resistance ≈ 30 Ω sq−1). An anti-reflective coating using a magnesium fluoride (MgF2) anti-reflective film formed by thermal evaporation deposition was only applied to the best cells in this study. Finally, each sample was sectioned into 16 solar cells (active area = 0.185 cm2) by scribing.

4.3 Characterization

Surface and cross-sectional images were obtained using a field-emission scanning electron microscope (FESEM; Hitachi S-4800). Elemental mapping of the prepared materials was conducted using scanning transmission electron microscopy with energy-dispersive spectrometry (STEM-EDS). Optical absorption spectra were recorded using a UV–visible spectrophotometer (NEOSYS-2000, SINCO Co., Ltd, Seoul, Korea). X-ray diffraction (XRD) analysis was performed using an Empyrean diffractometer (PANalytical Co.) with CuKα radiation (λ = 0.15406 nm) to determine the crystal structure of the films. Raman spectroscopy measurements were conducted using a Raman system equipped with an Mmac 750 spectrometer and a 532 nm excitation laser (irradiation power < 1 mW, spot size 0.7–1 μm). A 532 nm green laser was used, with its power reduced to 1% using an Nd filter. The laser was focused on the sample surface through a 100× magnification lens, and measurements were taken at three locations: left, center, and right of the sample surface. Ultraviolet photoelectron spectroscopy (UPS) spectra were acquired using an ESCALAB 250Xi system (Thermo Scientific Co.) to determine the work function and valence band maximum of the absorber layer surfaces. The elemental distribution of the thin films was analyzed using secondary ion mass spectrometry (SIMS) (TOF-SIMS 5, ION-TOF GmbH). The sputtering source parameters were O2+ ions with an energy of 1 keV, a current of 220 nA, an area of 300 × 300 μm, and a sputtered ion dose density (SPIDD) of 3.29 × 1018 ions per cm2. Amplitude modulation (AM)-mode Kelvin probe force microscopy (KPFM) measurements were performed using a commercial scanning probe microscope (n-Tracer, NanoFocus Inc.) under ambient conditions to investigate local electronic properties. The noncontact mode was used to scan the surface with a Pt/Ir-coated tip (Nanocensors Inc.), maintained at a 50 nm lift height. The scan rate was kept below 0.25 Hz to minimize tip damage. A lock-in amplifier (SR830, Stanford Research Systems Co.) provided KPFM feedback, with an applied AC amplitude of 1.0 V. Current–voltage (IV) curves were recorded under simulated air mass 1.5 global (AM 1.5G) illumination at 100 mW cm−2 (1 sun) and 25 °C using a 94022A solar simulator (Newport Co.). External quantum efficiency (EQE) was measured using an SR830 DSP lock-in amplifier system. Temperature-dependent voltage characteristics were analyzed under white light illumination using a source meter (2400, Keithley Co.) with an AM 1.5G spectrum Xe lamp (100 mW cm−2, Abet Technology Co.) over a temperature range of 120–300 K. Capacitance–voltage (CV) and drive-level capacitance profiling (DLCP) measurements were conducted using an LCR meter (E4980A, Agilent) to estimate space charge width and carrier density.

Author contributions

S. J. S. and D. H. K. conceived the idea and directed the overall project. V. Q. H. and J. L. fabricated the devices and conducted the characterization. B. K. E. conducted the CdS CBD process. G. L. and W. J. conducted the KPFM measurements and performed data analysis. A. A. and D. H. S. conducted the CV and DLCP measurements and performed data analysis. D. H. J., H. J. J., D. K. H., J. K. K., E. C. and K. J. Y. participated in discussing the experimental results. V. Q. H. and S. J. S. wrote the manuscript. All authors discussed the results and commented on the paper.

Data availability

The data that support the findings of this study are available from the corresponding author (D.-H. Kim, monolithic@dgist.ac.kr) upon reasonable request.

Supplementary information: Extensive material characterization, including structural and morphological analysis (XRD, TEM, SEM), electronic properties (UPS), and the performance metrics (JV curves, EQE spectra) of the highest-efficiency devices. See DOI: https://doi.org/10.1039/d5el00031a.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This research was supported by grants from the Korea Institute for Advancement of Technology (KIAT), funded by the Korean government (MOTIE) (No. P0024567), and the National Research Foundation of Korea (NRF), funded by the Korean government (MSIT) (No. RS-2025-02315803, RS-2023-NR076874). This work was also supported by the DGIST R&D programs, funded by the Korean government (MSIT) (No. 25-ET-01, 25-CoE-ET-01). The research at EWU was supported by the Basic Science Research Program through the NRF, funded by the Ministry of Education (No. NRF-2018R1A6A1A03025340) and the Ministry of Science and ICT (No. NRF-2022M3J1A1064229, RS-2024-00355905).

References

  1. S. Barthwal, R. Kumar and S. Pathak, ACS Appl. Energy Mater., 2022, 5, 6545–6585 CrossRef CAS.
  2. I. Gharibshahian, A. A. Orouji and S. Sharbati, Sol. Energy, 2021, 227, 606–615 CrossRef CAS.
  3. W. Yang and J. Moon, J. Mater. Chem. A, 2019, 7, 20467–20477 RSC.
  4. Y. Zhao, S. Wang, C. Jiang, C. Li, P. Xiao, R. Tang, J. Gong, G. Chen, T. Chen and J. Li, Adv. Energy Mater., 2022, 12, 2103015 CrossRef CAS.
  5. X. Wang, R. Tang, C. Jiang, W. Lian, H. Ju, G. Jiang, Z. Li, C. Zhu and T. Chen, Adv. Energy Mater., 2020, 10, 2002341 CrossRef CAS.
  6. Z. Li, X. Liang, G. Li, H. Liu, H. Zhang, J. Guo, J. Chen, K. Shen, X. San and W. Yu, Nat. Commun., 2019, 10, 125 CrossRef PubMed.
  7. X. Liang, C. Guo, T. Liu, Y. Liu, L. Yang, D. Song, K. Shen, R. E. Schropp, Z. Li and Y. Mai, Sol. RRL, 2020, 4, 2000294 CrossRef CAS.
  8. S. Rijal, D. B. Li, R. A. Awni, C. Xiao, S. S. Bista, M. K. Jamarkattel, M. J. Heben, C. S. Jiang, M. Al-Jassim and Z. Song, Adv. Funct. Mater., 2022, 32, 2110032 CrossRef CAS.
  9. G. Liang, M. Chen, M. Ishaq, X. Li, R. Tang, Z. Zheng, Z. Su, P. Fan, X. Zhang and S. Chen, Adv. Sci., 2022, 9, 2105142 CrossRef CAS PubMed.
  10. G. Lim, H. K. Park, Y. Wang, S. H. Ji, B. Shin and W. Jo, J. Phys. Chem. Lett., 2024, 15, 2825–2833 CrossRef CAS PubMed.
  11. S.-N. Park, S.-Y. Kim, S.-J. Lee, S.-J. Sung, K.-J. Yang, J.-K. Kang and D.-H. Kim, J. Mater. Chem. A, 2019, 7, 25900–25907 RSC.
  12. S.-N. Park, S.-Y. Kim, S.-J. Lee, S.-J. Sung, K.-J. Yang, J.-K. Kang and D.-H. Kim, Mater. Adv., 2022, 3, 978–985 RSC.
  13. K. Li, C. Chen, S. Lu, C. Wang, S. Wang, Y. Lu and J. Tang, Adv. Mater., 2019, 31, 1903914 CrossRef CAS PubMed.
  14. Z. Duan, X. Liang, Y. Feng, H. Ma, B. Liang, Y. Wang, S. Luo, S. Wang, R. E. Schropp and Y. Mai, Adv. Mater., 2022, 34, 2202969 CrossRef CAS.
  15. X. Wen, Z. Lu, X. Yang, C. Chen, M. A. Washington, G.-C. Wang, J. Tang, Q. Zhao and T.-M. Lu, ACS Appl. Mater. Interfaces, 2023, 15, 22251–22262 CrossRef CAS PubMed.
  16. Y. Zhao, S. Wang, C. Li, B. Che, X. Chen, H. Chen, R. Tang, X. Wang, G. Chen and T. Wang, Energy Environ. Sci., 2022, 15, 5118–5128 RSC.
  17. V.-Q. Hoang, D.-H. Jeon, S.-Y. Kim, J. Lee, D.-H. Son, K.-J. Yang, J.-K. Kang, S.-J. Sung, D.-K. Hwang and D.-H. Kim, J. Sci.: Adv. Mater. Devices, 2024, 9, 100665 CAS.
  18. V.-Q. Hoang, D.-H. Jeon, H. K. Park, S.-Y. Kim, W.-H. Kim, D.-K. Hwang, J. Lee, D.-H. Son, K.-J. Yang and J.-K. Kang, ACS Appl. Energy Mater., 2023, 6, 12180–12189 CrossRef CAS.
  19. M. Luo, M. Leng, X. Liu, J. Chen, C. Chen, S. Qin and J. Tang, Appl. Phys. Lett., 2014, 104, 173904 CrossRef.
  20. X. Liu, J. Chen, M. Luo, M. Leng, Z. Xia, Y. Zhou, S. Qin, D.-J. Xue, L. Lv and H. Huang, ACS Appl. Mater. Interfaces, 2014, 6, 10687–10695 CrossRef CAS PubMed.
  21. Z. Li, X. Chen, H. Zhu, J. Chen, Y. Guo, C. Zhang, W. Zhang, X. Niu and Y. Mai, Sol. Energy Mater. Sol. Cells, 2017, 161, 190–196 CrossRef CAS.
  22. Z. Li, H. Zhu, Y. Guo, X. Niu, X. Chen, C. Zhang, W. Zhang, X. Liang, D. Zhou and J. Chen, Appl. Phys. Express, 2016, 9, 052302 CrossRef.
  23. V. Kumar, E. Artegiani, P. Punathil, M. Bertoncello, M. Meneghini, F. Piccinelli and A. Romeo, ACS Appl. Energy Mater., 2021, 4, 12479–12486 CrossRef CAS.
  24. C. Chen and J. Tang, ACS Energy Lett., 2020, 5, 2294–2304 CrossRef CAS.
  25. J. Dong, Y. Liu, Z. Wang and Y. Zhang, Nano Sel., 2021, 2, 1818–1848 CrossRef CAS.
  26. Y. Zhang, T. Shi, L. Duan, B. Hoex and Z. Tang, Nano Energy, 2024, 131, 110282 CrossRef CAS.
  27. J. Schmidt, R. Peibst and R. Brendel, Sol. Energy Mater. Sol. Cells, 2018, 187, 39–54 CrossRef CAS.
  28. P. M. Salomé, B. Vermang, R. Ribeiro-Andrade, J. P. Teixeira, J. M. Cunha, M. J. Mendes, S. Haque, J. Borme, H. Aguas and E. Fortunato, Adv. Mater. Interfaces, 2018, 5, 1701101 CrossRef.
  29. K. Gao, Q. Bi, X. Wang, W. Liu, C. Xing, K. Li, D. Xu, Z. Su, C. Zhang and J. Yu, Adv. Mater., 2022, 34, 2200344 CrossRef CAS PubMed.
  30. S. M. George, Chem. Rev., 2010, 110, 111–131 CrossRef CAS PubMed.
  31. R. L. Puurunen, J. Appl. Phys., 2005, 97, 121301 CrossRef.
  32. M. Knez, K. Nielsch and L. Niinistö, Adv. Mater., 2007, 19, 3425–3438 CrossRef CAS.
  33. C. Altinkaya, E. Aydin, E. Ugur, F. H. Isikgor, A. S. Subbiah, M. De Bastiani, J. Liu, A. Babayigit, T. G. Allen and F. Laquai, Adv. Mater., 2021, 33, 2005504 CrossRef CAS PubMed.
  34. S. Lan, W. Zheng, S. Yoon, H. U. Hwang, J. W. Kim, D.-W. Kang, J.-W. Lee and H.-K. Kim, ACS Appl. Energy Mater., 2022, 5, 14901–14912 CrossRef CAS.
  35. Y. Zhou, Y. Li, J. Luo, D. Li, X. Liu, C. Chen, H. Song, J. Ma, D.-J. Xue and B. Yang, Appl. Phys. Lett., 2017, 111, 013901 CrossRef.
  36. C. Wang, S. Lu, S. Li, S. Wang, X. Lin, J. Zhang, R. Kondrotas, K. Li, C. Chen and J. Tang, Nano Energy, 2020, 71, 104577 CrossRef CAS.
  37. G. Li, Z. Li, X. Liang, C. Guo, K. Shen and Y. Mai, ACS Appl. Mater. Interfaces, 2018, 11, 828–834 CrossRef PubMed.
  38. K. Shen, C. Ou, T. Huang, H. Zhu, J. Li, Z. Li and Y. Mai, Sol. Energy Mater. Sol. Cells, 2018, 186, 58–65 CrossRef CAS.
  39. K. Yang, B. Li and G. Zeng, Sol. Energy Mater. Sol. Cells, 2020, 208, 110381 CrossRef CAS.
  40. D. Ren, B. Fu, J. Xiong, Y. Wang, B. Zhu, S. Chen, Z. Li, H. Ma, X. Zhang and D. Pan, Adv. Mater., 2025, 37, 2416885 CrossRef CAS PubMed.
  41. C. Chen, W. Li, Y. Zhou, C. Chen, M. Luo, X. Liu, K. Zeng, B. Yang, C. Zhang and J. Han, Appl. Phys. Lett., 2015, 107, 043905 CrossRef.
  42. E. Cho, S.-J. Sung, K.-J. Yang, J. Lee, V.-Q. Hoang, B. Kadiri-English, D.-K. Hwang, J.-K. Kang and D.-H. Kim, J. Mater. Chem. A, 2025, 13, 8507–8517 RSC.
  43. S. Dias, B. Murali and S. Krupanidhi, Sol. Energy Mater. Sol. Cells, 2015, 143, 152–158 CrossRef CAS.
  44. Q. Xie, Z. Liu, M. Shao, L. Kong, W. Yu and Y. Qian, J. Cryst. Growth, 2003, 252, 570–574 CrossRef CAS.
  45. T. Zhai, M. Ye, L. Li, X. Fang, M. Liao, Y. Li, Y. Koide, Y. Bando and D. Golberg, Adv. Mater., 2010, 22, 4530–4533 CrossRef CAS PubMed.
  46. Z. Cai, B. Che, Y. Gu, P. Xiao, L. Wu, W. Liang, C. Zhu and T. Chen, Adv. Mater., 2024, 36, 2404826 CrossRef CAS PubMed.
  47. S. K. Kim, H. K. You, K. R. Yun, J. H. Kim and T. Y. Seong, Adv. Opt. Mater., 2023, 11, 2202625 CrossRef CAS.
  48. N. Ahmad, Y. Zhao, F. Ye, J. Zhao, S. Chen, Z. Zheng, P. Fan, C. Yan, Y. Li and Z. Su, Adv. Sci., 2023, 10, 2302869 CrossRef CAS PubMed.
  49. W. Yang, J. H. Kim, O. S. Hutter, L. J. Phillips, J. Tan, J. Park, H. Lee, J. D. Major, J. S. Lee and J. Moon, Nat. Commun., 2020, 11, 861 CrossRef CAS PubMed.
  50. Y. M. Lee, B. K. Jung, J. Ahn, T. Park, C. Shin, T. N. Ng, I. S. Kim, J. H. Choi and S. J. Oh, Adv. Electron. Mater., 2022, 8, 2200297 CrossRef CAS.
  51. A. Chen and K. Zhu, Sol. Energy, 2014, 107, 195–201 CrossRef CAS.
  52. C. Otalora, M. Botero and G. Ordonez, J. Mater. Sci., 2021, 56, 15538–15571 CrossRef CAS.
  53. N. Maticiuc, A. Katerski, M. Danilson, M. Krunks and J. Hiie, Sol. Energy Mater. Sol. Cells, 2017, 160, 211–216 CrossRef CAS.
  54. R. Ortega-Borges and D. Lincot, J. Electrochem. Soc., 1993, 140, 3464 CrossRef CAS.
  55. L. Protesescu, M. Nachtegaal, O. Voznyy, O. Borovinskaya, A. J. Rossini, L. Emsley, C. Copéret, D. Günther, E. H. Sargent and M. V. Kovalenko, J. Am. Chem. Soc., 2015, 137, 1862–1874 CrossRef CAS PubMed.
  56. T. P. Weiss, S. Nishiwaki, B. Bissig, R. Carron, E. Avancini, J. Löckinger, S. Buecheler and A. N. Tiwari, Adv. Mater. Interfaces, 2018, 5, 1701007 CrossRef.
  57. S. Kim, J. Lee, D.-H. Son, W. H. Kim, S.-J. Sung, D.-K. Hwang, T. E. Hong, N. Otgontamir, E. Enkhbayar and T.-H. Lee, Energy Environ. Sci., 2024, 17, 8609–8620 RSC.
  58. T. K. Todorov, J. Tang, S. Bag, O. Gunawan, T. Gokmen, Y. Zhu and D. B. Mitzi, Adv. Energy Mater., 2013, 3, 34–38 CrossRef CAS.
  59. K. Zeng, D.-J. Xue and J. Tang, Semicond. Sci. Technol., 2016, 31, 063001 CrossRef.
  60. D. Lee, J. Y. Cho and J. Heo, Sol. Energy, 2018, 173, 1073–1079 CrossRef CAS.
  61. A. K. Jain, R. Anandan and P. Malar, Mater. Res. Express, 2023, 10, 105502 CrossRef CAS.
  62. V. Kumar, E. Artegiani, A. Kumar, G. Mariotto, F. Piccinelli and A. Romeo, Sol. Energy, 2019, 193, 452–457 CrossRef CAS.
  63. M. Leng, M. Luo, C. Chen, S. Qin, J. Chen, J. Zhong and J. Tang, Appl. Phys. Lett., 2014, 105, 083905 CrossRef.
  64. L. Huang, J. Yang, Y. Xia, P. Xiao, H. Cai, A. Liu, Y. Wang, X. Liu, R. Tang and C. Zhu, J. Mater. Chem. A, 2023, 11, 16963–16972 RSC.
  65. S. Yao, J. Wang, J. Cheng, L. Fu, F. Xie, Y. Zhang and L. Li, ACS Appl. Mater. Interfaces, 2020, 12, 24112–24124 CrossRef CAS PubMed.

Footnote

These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.