1. Wuhan National Laboratory for Optoelectronics (WNLO) and School of Optical and Electronic Information (SOEI), Huazhong University of Science and Technology, Wuhan 430074, China
2. Optics Valley Laboratory, Wuhan 430074, China
jtang@mail.hust.edu.cn
cchen@hust.edu.cn
Show less
History+
Received
Accepted
Published Online
2025-09-03
2025-12-21
2026-01-07
PDF
(2183KB)
Abstract
All-perovskite tandem solar cells are a promising photovoltaic technology, but their efficiency is strongly limited by the tunnel junction. The tunnel junction enables carrier tunneling and recombination, which depend on the interfacial band alignment. Through quantitative simulations using Silvaco Technology Computer Aided Design (TCAD), we find that hole tunneling is intrinsically more difficult than electron tunneling in the tunnel junction. Efficient tunnel junctions require minimizing the barrier for holes while maintaining a moderate barrier for electrons to balance tunneling. For the SnO2/metal/PEDOT:PSS tunnel junction in all-perovskite tandem solar cells, tuning the metal work function achieves balanced electron and hole tunneling, reduces junction resistance, and directly enhances performance of tandem solar cells. This work provides quantitative design rules for tunnel junction optimization, offering a clear pathway toward high-performance all-perovskite tandem solar cells.
All-perovskite tandem solar cells (TSCs) have reached efficiencies above 30% and still show room for improvement toward the theoretical limit (~45%) [1–3]. Advancing toward this limit requires not only high-quality absorbers but also efficient electrical connection between sub-cells [4,5]. The tunnel junction is essential, as it connects the sub-cells in series and controls carrier recombination at the interface [6–8].
In most cases, the tunnel junction contains a thin metal layer between SnO2 electron transport layer (ETL) and PEDOT:PSS hole transport layer (HTL) [1,2,9]. Charge transport involves tunneling across SnO2/metal and PEDOT:PSS/metal interfaces [7]. Small changes in interfacial band alignment can strongly influence tunneling because the probability depends exponentially on barrier height [10]. Conventional III–V TSCs use heavily doped p+/n+ tunnel junction, enabling efficient band-to-band tunneling with weak sensitivity to interfacial band alignment [11]. In contrast, tunnel junctions of all-perovskite TSCs exhibit lower tunneling probabilities and are more prone to interfacial mismatch [7]. Progress on tunnel junctions of all-perovskite TSCs so far has mostly relied on empirical choices, such as using Au or Cu/Au composite as the interlayer metal [12–14], but understanding on the tunneling mechanism remains unclear.
In this work, we conducted fundamental analysis and simulations of the tunnel junction in all-perovskite TSCs. First, we examined the characteristics of the tunnel junction and found that hole tunneling is intrinsically more difficult than electron tunneling. Both electron and hole tunneling are strongly affected by the barrier height, which is determined by the metal work function (ΦM) in the tunnel junction. We then systematically tuned the ΦM to modify interface band alignment, which affects tunneling current and equivalent series resistance. We find that achieving an efficient tunnel junction requires minimizing the barrier at the PEDOT:PSS/metal interface to enable hole tunneling, while keeping a moderate barrier at the SnO2/metal interface to allow electron tunneling. Finally, we evaluated the effect of the tunnel junction on performance of TSCs and found that device-level trends follow analysis of the tunnel junction. The highest efficiency appears when hole and electron tunneling are balanced and efficient. These results provide quantitative design rules for tunnel junction engineering, offering clear guidance for advancing high-efficiency all-perovskite TSCs.
2 Results and discussion
2.1 Characteristics of tunneling junction in all-perovskite TSCs
A two-terminal tandem solar cell (TSC) consists of a wide-bandgap (~1.78 eV) top sub-cell, a tunnel junction, and a narrow-bandgap (~1.2 eV) bottom sub-cell (Fig. 1a). The tunnel junction conducts opposite to the sub-cells, so it operates under reverse bias when the TSCs work at maximum power point (Fig. 1b). Tunnel junctions typically use highly doped semiconductors to enable rapid carrier recombination and a low equivalent resistance. In III–V TSCs (Fig. 1c), n++/p++ tunnel junctions with doping ~1020 cm−3 allow efficient interband tunneling [15]. The resulting J–V curve exhibits a “S-shape” with valley current density (Jvalley) exceeding the sub-cell photocurrent by more than an order of magnitude, ensuring unimpeded current flow (Fig. 1d) [11].
In all-perovskite TSCs, the SnO2 ETL and PEDOT:PSS HTL are lightly doped, typically below 1018 cm−3 [16,17]. Such low doping is insufficient to achieve interband tunneling. To achieve efficient tunneling in all-perovskite TSCs, a composite tunnel junction uses an ultrathin metal layer between the ETL and HTL to assist tunneling [8,12,18]. The metal forms a near-Ohmic contact with the HTL while creating a Schottky junction with the ETL (Fig. 1e and 1f). In this configuration, charge transport across the tunnel junction is governed primarily by electron tunneling, thereby maintaining current conduction. The tunneling current density through the Schottky junction can be described by Eq. (1) [19],
where J is the saturation current density, q is the elementary charge, ℏ is the reduced Planck constant, m* is the electron or hole effective mass, εr is the relative permittivity of the semiconductor, ε0 is the vacuum permittivity, ND is the doping concentration, VD is barrier height, and V is the applied bias voltage. From Eq. (1), tunneling probability depends strongly on both barrier height and carrier effective mass. In SnO2, electrons have an effective mass of ~0.2 m0 [20], while in PEDOT:PSS, holes reach ~4.8 m0 [21]. Thus, the tunneling probability of holes is four orders of magnitude lower than electrons. Therefore, minimizing the HTL/metal barrier while keeping electron tunneling at the ETL/metal interface efficient is crucial. From Eq. (1), tunneling barriers for electrons and holes can be tuned by adjusting the ΦM. The following section will focus on the role of the ΦM in tailoring the barrier properties of the tunnel junction and its impact on performance of TSCs.
2.2 Simulation of tunnel junction in all-perovskite TSCs
Simulations of tunnel junction were performed using the Atlas module of Silvaco Technology Computer Aided Design (TCAD). Silvaco TCAD is a widely used commercial tool for semiconductor device modeling. It uses finite-element methods to solve the full set of semiconductor equations, including the Poisson equation, continuity equations, and drift-diffusion or carrier-transport models. Atlas is a physics-based semiconductor simulator that solves the Poisson equation along with the electron and hole continuity equations under steady-state conditions [22]. To accurately model tunneling transport across the Schottky barrier, the universal Schottky tunneling model was implemented [23]. Carrier transport at the tunnel-junction interface is modeled using the Schottky tunneling model. The tunneling probability is calculated via the WKB approximation. Image-force corrections are included in the calculation of the effective barrier height. We use the standard image-potential formulation as follows [24,25]:
where Eeff is the effective barrier height, VD is the ideal barrier height, q is the elementary charge, E is the interfacial electric field, and ϵs is the semiconductor dielectric constant. This treatment ensures that the computed tunneling current reflects realistic interfacial physics.
The work functions of ETL and HTL are approximately 4.6 eV and 5.3 eV, respectively, estimated from their doping levels and effective density of states (Table S1). To evaluate the impact of band alignment, ΦM was scanned from 4.2 eV to 5.6 eV, covering common electrode materials such as Ag (4.3 eV) and Au (5.1 eV) [26]. Based on the position of ΦM relative to the ETL and HTL work functions, the tunnel junction can be divided into three situations, as shown in Fig. 2a–c.
When ΦM ranges from 4.2 eV to 4.6 eV, the tunnel junction enters the regime shown in Fig. 2a. An Ohmic contact forms with the ETL, while the HTL/metal interface has a large barrier exceeding 0.7 eV. This high barrier can strongly suppress hole tunneling. The simulated J–V characteristics indicate that the forward tunneling current is limited (Fig. 2d), resulting in an extracted equivalent series resistance exceeding 100 Ω cm2 (Fig. 2g). In typical tandem sub-cells, intrinsic series resistance ranges from 30 to 40 Ω cm2 [27]. The additional resistance from the tunnel junction can be more than four times the total, significantly lowering the fill factor (FF) and efficiency of TSCs.
When ΦM ranges from 4.6 eV to 5.1 eV, the tunnel junction transitions to the regime illustrated in Fig. 2b. The turn-on voltage of tunnel junction gradually decreases from above 0.4 V to nearly 0 V (Fig. 2e). Consequently, the equivalent series resistance drops sharply from over 100 Ω cm2 to below 10−2 Ω cm2 (Fig. 2g), due to the reduction of the HTL/metal barrier from ~0.7 eV to ~0.2 eV. According to Eq. (1), lowering the hole barrier height by ~0.5 eV greatly increases the hole tunneling probability by over ten orders of magnitude, overcoming the intrinsic disadvantage of the larger hole effective mass. As a result, holes can efficiently tunnel and recombine with electrons in the metal, sharply reducing the equivalent resistance of the tunnel junction. At a representative ΦM of 5.1 eV, the HTL/metal barrier is reduced to ~0.2 eV, while the ETL/metal barrier remains moderate at ~0.5 eV. Under these barrier conditions, holes and electrons tunnel with balanced efficiency, reducing the tunnel junction series resistance to about 10−2 Ω cm2. This balance is crucial because electrons and holes have different effective masses in SnO2 and PEDOT:PSS, and the tunneling current depends on both the barrier height and carrier effective mass. The configuration balances hole and electron tunneling for optimal tunnel junction.
When ΦM ranges from 5.1 eV to 5.6 eV, the HTL/metal interface forms an Ohmic contact (Fig. 2c), while the ETL/metal barrier rises from ~0.5 eV to ~1.0 eV. The increasing ETL barrier suppresses electron tunneling, making carrier recombination progressively limited by electron tunneling (Fig. 2f). As a result, the tunnel junction series resistance increases sharply from ~102 Ω cm2 to ~105 Ω cm2 (Fig. 2g), reflecting the reduced electron conduction.
We further assessed the robustness of the optimal intermediate-metal work-function range under realistic material and device variations. Sensitivity studies on SnO2 doping (1010–1017 cm−3) and the hole effective mass of PEDOT:PSS (4.0–5.5 m0) all show that the minimum equivalent series resistance consistently appears at ΦM = 5.0–5.2 eV (Fig. S1). Temperature-dependent simulations from 25°C to 125°C also preserve this optimal range (Fig. S2), despite possible changes in carrier effective masses, metal work functions, and barrier heights. These results confirm that the identified optimal ΦM range remains stable against realistic variations in material parameters and operating temperature, supporting the general applicability of our conclusions. In addition, we validated our simulation results experimentally using SnO2/metal/PEDOT:PSS tunnel junctions with Au (5.1 eV), Cu (4.7 eV), and Ag (4.3 eV) as intermediate metals. The measured J–V characteristics and series resistances follow the trends predicted by simulations, with metals near the optimal ΦM range (5.0–5.2 eV) showing the best conduction and lowest resistance (Fig. S3).
2.3 Effect of ΦM on device performance
Simulations show that achieving balanced and efficient tunneling of holes and electrons is essential for a high-performance tunnel junction. The next step involves investigating the impact of the optimized tunnel junction on the performance of all-perovskite TSCs.
The J–V characteristics of the wide-bandgap sub-cell, narrow-bandgap sub-cell, and the all-perovskite TSCs were simulated (Fig. 3a–c). The tandem simulation first uses Au as the interlayer metal to verify the validity of the tandem model [8]. The simulated device parameters of the wide-bandgap sub-cell, narrow-bandgap sub-cell, and the TSC are summarized in Table 1. The simulation results indicate that the TSC achieves an open-circuit voltage (VOC) of 2.29 V, which is nearly the sum of the VOC values of the wide-bandgap (1.39 V) and narrow-bandgap (0.94 V) sub-cells, demonstrating effective series connection. The short-circuit current density (JSC) is 17.38 mA cm−2, limited by the lower-generating wide-bandgap sub-cell, in line with current-matching requirements. The FF reaches 85.4%, slightly higher than the wide-bandgap (85.0%) and narrow-bandgap (80.1%) sub-cells. The overall power conversion efficiency (PCE) achieves 34.01%, indicating efficient carrier recombination across the optimized tunnel junction. These results demonstrate that the Silvaco TCAD model accurately captures behavior of TSCs, providing a solid basis to analyze the effect of ΦM on TSCs.
When ΦM is low (4.2–4.6 eV), the TSC shows poor performance. The J–V curve (Fig. 3d) shows a slow increase of current density near the VOC, indicating hindered carrier transport. The FF drops to ~68% and the VOC is limited to ~1.78 V (Fig. 3f and 3g). Based on the previous tunnel junction analysis, a shallow ΦM creates a large barrier (> 0.7 eV) at the HTL/metal interface. The high barrier suppresses hole tunneling and recombination, preventing effective series connection between the two sub-cells.
As ΦM increases from 4.6 eV to 5.1 eV, the TSC performance improves. The J–V curve (Fig. 3d) becomes steeper near the VOC, indicating more efficient carrier transport. FF rises from 68% to 85% and VOC increases from 1.78 V to 2.29 V (Fig. 3f and 3g). At ΦM between 4.6 eV and 5.1 eV, the HTL/metal barrier reduces to ~0.1–0.2 eV while the ETL/metal barrier remains moderate (< 0.5 eV). Both holes and electrons tunnel efficiently. The tunnel junction provides minimal additional resistance, enabling effective series connection between the two sub-cells. At ΦM = 5.1 eV, the TSC reaches a maximum FF of ~85%, consistent with tunnel junction simulations showing the lowest equivalent series resistance (10−2 Ω cm2).
When ΦM exceeds 5.1 eV, the TSC performance declines slightly (Fig. 3e), with the FF dropping from 85% to 72% (Fig. 3g). The HTL forms an Ohmic contact, but the ETL/metal barrier rises above 0.5 eV. The higher barrier reduces electron tunneling and limits recombination in the tunnel junction. The tunnel junction no longer provides a low-resistance pathway, preventing further performance gains.
The JSC of TSCs remains nearly constant across the ΦM range (Fig. 3h). Under short-circuit or reverse-bias conditions of the TSCs, the tunnel junction operates at a substantial forward bias, allowing efficient tunneling. This forward tunneling maintains carrier flow, keeping the JSC nearly unchanged. The PCE of TSCs exhibits a clear maximum around ΦM ≈ 5.1 eV (Fig. 3i), following the trends in FF. The FF of TSCs varies significantly with the ΦM, consistent with the change in equivalent series resistance of the tunnel junction. The open-circuit voltage remains nearly unchanged in the range of 4.8 eV to 5.4 eV. These results show that tunnel junction performance directly controls efficiency of TSCs through barrier engineering. We further studied the effect of metal thickness on tunnel junction performance. Varying thickness from 5 nm to 10 nm has minimal impact on series resistance (Fig. S4a), as tunneling is controlled by interface barriers. Thicker metals mainly increase optical losses, reducing photon flux to the bottom sub-cell and lowering short-circuit current, while open-circuit voltage and fill factor remain nearly unchanged (Fig. S4b−e).
Device simulations reveal that the interlayer ΦM affects performance parameters of all-perovskite TSCs by dictating the recombination efficiency and equivalent series resistance of the tunnel junction. Adjusting ΦM to 5.0−5.2 eV establishes a small barrier at the HTL/metal interface and a moderate barrier at the ETL/metal interface, enabling efficient tunneling of both holes and electrons and minimizing the equivalent series resistance of the tunnel junction to ~10−2 Ω cm2. Under these conditions, the TSCs achieve a FF of 85% and a peak efficiency of 34%, demonstrating that careful tuning of the tunnel junction is essential for maximizing device performance. The simulated efficiency of 34.01% is higher than typical experimental values due to the use of a near-ideal device model. Bulk recombination is included using representative minority-carrier lifetimes [12,13] (1 μs for wide-bandgap perovskite and 5 μs for narrow-bandgap perovskite), while interface defects and non-radiative recombination are excluded. This efficiency represents an upper theoretical limit and serves as a benchmark for determining the optimal interlayer metal work-function range of 5.0–5.2 eV.
Our method optimizes interface energy alignment via the intermediate-layer work function ΦM to achieve balanced tunneling and minimal series resistance. It is general and can analyze any metal-based tunnel junction. The framework predicts energy alignment with adjacent charge transport layers (e.g., SnO2 and PEDOT:PSS or alternatives) and identifies metals or alloys for optimal conduction.
3 Conclusion
Through Silvaco TCAD simulations, we systematically elucidated the tunneling processes within the tunnel junction of all-perovskite TSCs. The study highlights that carrier transport across the junction is predominantly dictated by the energy barriers at the SnO2/metal and PEDOT:PSS/metal interfaces, which are in turn controlled by the ΦM. Optimal tunneling occurs when ΦM ≈ 5.1 eV, yielding a small ~0.2 eV barrier at the HTL side and a moderate ~0.5 eV barrier at the ETL side. This balanced alignment enables efficient bidirectional tunneling of holes and electrons, reduces the equivalent junction resistance of tunnel junction to ~10−2 Ω cm2, and ensures effective carrier recombination within the interlayer. These results establish ΦM-driven band alignment as the key design principle for engineering high-performance tunnel junctions in all-perovskite TSCs.
Liu, Z., Lin, R., Wei, M., Yin, M., Wu, P., Li, M., Li, L., Wang, Y., Chen, G., Carnevali, V., Agosta, L., Slama, V., Lempesis, N., Wang, Z., Wang, M., Deng, Y., Luo, H., Gao, H., Rothlisberger, U., Zakeeruddin, S.M., Luo, X., Liu, Y., Gratzel, M., Tan, H.: All-perovskite tandem solar cells achieving >29% efficiency with improved (100) orientation in wide-bandgap perovskites. Nat. Mater24(2), 252–259(2025)
[2]
Yang, X., Ma, T., Hu, H., Ye, W., Li, X., Li, M., Zhang, A., Ge, C., Sun, X., Zhu, Y., Yan, S., Yan, J., Zhou, Y., Li, Z., Chen, C., Song, H., Tang, J.: Understanding and manipulating the crystallization of Sn–Pb perovskites for efficient all-perovskite tandem solar cells. Nat. Photonics19(4), 426–433(2025)
[3]
Zhu, Y., Ge, C., Xu, Q., Liu, D., Zhang, P., Zu, X., Dong, C., Yang, X., Ye, W., Hu, H., Zhu, Z., Zhang, Z., Du, J., Pang, S., Hu, L., Song, H., Xu, L., Zhou, Y., Chen, C., Tang, J.: Simplified surface defects of Sn-Pb perovskite for efficient all-perovskite tandem solar cells. Nano Energy139, 110927(2025)
[4]
Zhang, M., Lin, Z.: Efficient interconnecting layers in monolithic all-perovskite tandem solar cells. Energy Environ. Sci15(8), 3152–3170(2022)
[5]
Li, C., Wang, Y., Choy , W.C.H.: Efficient interconnection in perovskite tandem solar cells. Small Methods4(7), 2000093(2020)
[6]
Ye, W., Chen, A., Fu, P., Tang, J., Chen, C.: Simulation study of reducing reflection losses in all-perovskite tandem solar cells through dual serrated structure. Front Optoelectron18(1), 9(2025)
[7]
Dong, C., Yan, S., Liu, D., Zhu, Y., Chen, C., Tang, J.: Principle and progress of interconnection layers in monolithic perovskite‐based tandem photovoltaics. Adv. Energy Mater15(13), 2404628(2025)
[8]
Wen, J., Hu, H., Chen, C., McMeekin, D.P., Xiao, K., Lin, R., Liu, Y., Snaith, H.J., Tang, J., Paetzold, U.W., Tan, H.: Present status of and future opportunities for all-perovskite tandem photovoltaics. Nat. Energy10(6), 681–696(2025)
[9]
Ge, C., Xu, Q., Liu, D., Ye, W., Zhu, Y., Zhang, P., Yang, J., Liang, G., Xu, L., Zhou, Y., Song, H., Chen, C., Tang, J.: Light radiation annealing enables unidirectional crystallization of vacuum-assisted Sn–Pb perovskites for efficient tandem solar cells. Energy Environ. Sci18(1), 430–438(2025)
[10]
Werner, J.H., Güttler , H.H.: Barrier inhomogeneities at Schottky contacts. J. Appl. Phys69(3), 1522–1533(1991)
[11]
Dawidowski, W., Ściana , B. , Zborowska-Lindert, I., Mikolášek , M. , Kováč , J. , Tłaczała , M.: Tunnel junction limited performance of InGaAsN/GaAs tandem solar cell. Sol. Energy214, 632–641(2021)
Li, M., Yan, J., Zhang, A., Zhao, X., Yang, X., Yan, S., Ma, N., Ma, T., Luo, D., Chen, Z., Li, L., Li, X., Chen, C., Song, H., Tang, J.: Vacuum-driven precrystallization enables efficient all-perovskite tandem solar cells. Joule9(4), 101825(2025)
[14]
Li, C., Wang, Z.S., Zhu, H.L., Zhang, D., Cheng, J., Lin, H., Ouyang, D., Choy , W.C.H.: Thermionic emission–based interconnecting layer featuring solvent resistance for monolithic tandem solar cells with solution‐processed perovskites. Adv. Energy Mater8(36), 1801954(2018)
[15]
Salem, M.S., Saif, O.M., Shaker, A., Abouelatta, M., Alzahrani, A.J., Alanazi, A., Elsaid, M.K., Ramadan, R.A.: Performance optimization of the InGaP/GaAs dual-junction solar cell using Silvaco TCAD. Int. J. Photoenergy2021, 1–12(2021)
[16]
Biswas, S.K., Sumon, M.S., Sarker, K., Orthe, M.F., Ahmed, M.M.: A numerical approach to analysis of an environment-friendly Sn-based perovskite solar cell with SnO2 buffer layer using SCAPS-1D. Adv. Mater. Sci. Eng2023, 1–10(2023)
[17]
Hossain, M.K., Mohammed , M.K.A. , Pandey, R., Arnab, A.A., Rubel , M.H.K. , Hossain, K.M., Ali, M.H., Rahman, M.F., Bencherif, H., Madan, J., Islam, M.R., Samajdar, D.P., Bhattarai, S.: Numerical analysis in DFT and SCAPS-1D on the influence of different charge transport layers of CsPbBr3 perovskite solar cells. Energy Fuels37(8), 6078–6098(2023)
[18]
Lan, S., Pan, B., Liu, Y., Zhang, Z., Zhang, L., Yu, B., Fang, Y., Wang, P.: Preparation and promising optoelectronic applications of lead halide perovskite patterned structures: A review. Carbon Energy5(10), e318(2023)
[19]
Rideout, V.L., Crowell, C.R.: Effects of image force and tunneling on current transport in metal-semiconductor (Schottky barrier) contacts. Solid-State Electron13(7), 993–1009(1970)
[20]
Xu, J., Huang, S., Wang, Z.: First principle study on the electronic structure of fluorine-doped SnO2. Solid State Commun149(13-14), 527–531(2009)
[21]
Kong, M., Garriga, M., Reparaz, J.S., Alonso, M.I.: Advanced optical characterization of PEDOT:PSS by combining spectroscopic ellipsometry and raman scattering. ACS Omega7(43), 39429–39436(2022)
[22]
Michael, S., Bates, A.: The design and optimization of advanced multijunction solar cells using the Silvaco ATLAS software package. Sol. Energy Mater. Sol. Cells87(1-4), 785–794(2005)
[23]
Hur, J., Jeong, W.J., Shin, M., Choi, Y.K.: Schottky tunneling effects in a tunnel FET. IEEE Trans. Electron Dev64(12), 5223–5229(2017)
[24]
Miller, S.C., Good, R.H.: A WKB-type approximation to the Schrödinger equation. Phys. Rev91(1), 174–179(1953)
[25]
Lang, N.D., Kohn, W.: Theory of metal surfaces: Induced surface charge and image potential. Phys. Rev., B, Solid State7(8), 3541–3550(1973)
[26]
Skriver, H.L., Rosengaard, N.M.: Surface energy and work function of elemental metals. Phys. Rev. B Condens. Matter46(11), 7157–7168(1992)
[27]
Mundhaas, N., Yu, Z.J., Bush, K.A., Wang, H.P., Häusele , J. , Kavadiya, S., McGehee, M.D., Holman, Z.C.: Series resistance measurements of perovskite solar cells using JSC–VOC measurements. Sol. RRL3(4), 1800378(2019)
RIGHTS & PERMISSIONS
The author(s)
AI Summary 中Eng×
Note: Please be aware that the following content is generated by artificial intelligence. This website is not responsible for any consequences arising from the use of this content.