Reducing energy loss in WBG perovskite subcells
PBP consists of an anionic part where the central bromine atom is bonded to two adjacent bromine atoms and a cationic part from pyridinium salts (Fig. 1a). It can occupy a halide site or equivalently fill a halide vacancy, coordinating with the perovskite. As illustrated in Fig. 1a, the pure solution of PBP, the mixed solution of PBP and PbI2, and the mixed solution of PBP and CsI exhibited the color of orange, red-brown, and brown, respectively. These noticeable color differences indicate strong interactions between PBP and the perovskite precursors. Proton nuclear magnetic resonance (1H NMR) spectroscopy was employed to investigate the existing interaction. For clarity in analysis, we labeled the hydrogen atoms at different positions of the PBP as Ha, Hb, and Hc, respectively. As depicted in Fig. 1b, compared to the pure PBP solution, all proton signals shift upfield in the PBP and PbI2 mixed solution. However, it remains unclear whether this shift is due to the formation of N − H···I hydrogen bonds, coordination between N − H and Pb2+ or a combination of both effects. To elucidate this, we compared the 1H NMR differences between the PBP and CsI mixed solution and the PBP solution. Only the peaks of Ha and Hb shift upfield, while the peak of Hc remains almost unchanged. Since PBP and CsI interact mainly through hydrogen bonding, the upfield shift of Ha and Hb is attributed to the formation of N − H···I hydrogen bonds19, 20. Therefore, the upfield shift of Hc in the mixed solution of PBP and PbI2 is solely due to the deshielding effect from the coordination of N − H with Pb2 + 21. Moreover, combining the effects of hydrogen bonding and coordination, the shift of Ha and Hb peaks in the mixed solution of PBP and PbI2 is significantly greater than that in the mixed solution of PBP and CsI. These results reveal the existence of N − H···I hydrogen bonds and coordination between PBP and perovskite precursors.
Further confirmation of the interactions between PBP and perovskites was obtained through X-ray photoelectron spectroscopy (XPS). The Pb 4f XPS spectra (Fig. 1c) show that the Pb 4f5/2 and 4f7/2 peaks of the perovskite surface shift from 143.04 and 138.22 eV (untreated) to 142.85 and 138.00 eV (PBP-treated), suggesting alterations in the electron cloud density surrounding Pb. This change could be attributed to electron sharing between the − NH − group in PBP and Pb2+ in the perovskite22. Similarly, as shown in the I 3d XPS spectrum in Fig. 1d, we find that the I 3d3/2 and 3d5/2 peaks of the untreated perovskite surface are located at 630.44 and 618.97 eV, respectively. After PBP optimization, these peaks shift to 630.09 and 618.63 eV, indicating the formation of N − H···I hydrogen bonds between PBP and I− in the perovskite23. In contrast, the Cs 3d XPS spectra (Supplementary Fig. 1) exhibit no notable shift, indicating the unchanged chemical environment around Cs, which is consistent with results obtained from the 1H NMR analysis.
As for the uniformity and defect distribution details of the untreated and PBP-treated perovskite films, photoluminescence (PL) mapping displays the corresponding information (Fig. 1e). From the statistical histograms depicting the spatial distribution of luminescence intensity (Supplementary Fig. 2), it can be seen that the PL intensity distribution in the untreated perovskite film is more dispersed with poor distribution uniformity, indicating a low and uneven luminescence intensity that suggests the presence of numerous defects or non-uniform regions24. In contrast, the PL intensity distribution in the PBP-treated perovskite film is more concentrated, reflecting improved uniformity and higher film quality with fewer defects.
The influence of PBP on perovskite morphology was verified through scanning electron microscopy (SEM) and atomic force microscopy (AFM) measurements. As shown in the SEM images (Fig. 1f), it is evident that while both untreated and PBP-treated perovskite films exhibit distinct perovskite grains and grain boundaries, the grain size of untreated perovskite films is uneven, with sizes primarily ranging from 350 to 500 nm, showing a broad distribution. In contrast, the grain size of the PBP-treated perovskite is concentrated within a narrower range of 450 to 500 nm, indicating a more uniform internal structure of the film. This helps reduce grain boundary defects caused by size differences, facilitates smoother carrier transport between grains and reduces recombination losses at the grain boundaries. Figure 1g illustrates the AFM images of untreated and PBP-treated perovskite films, where the perovskite film after PBP treatment possessed slightly larger grain domain sizes, leading to the surface roughness increase from 13 to 18 nm, which is conducive to forming better physical contact and reduce defects.
To exclude the defect evolution of WBG perovskite, PL spectra under continuous illumination were performed (Fig. 2a). During the initial 10 min of illumination, the PL intensity of the pristine perovskite film exhibits a rapid decline, indicating an increase in recombination centers on the perovskite surface. Following this period, the PL intensity continues to decrease, albeit at a slower rate, implying a gradual accumulation of defect states within the perovskite film. In contrast, the perovskite film with PBP treatment shows no significant reduction in PL intensity throughout the entire duration of illumination.
The trap state density (nt) and carrier mobility (µ) can be obtained by fitting the space-charge-limited current (SCLC) model based on electron-only and hole-only devices. The trap-filled limit voltage (VTFL) provides information about the nt, calculated using the formula25:
V TFL = \(\:\frac{\text{e}{L}^{2}{n}_{t}}{2{\epsilon\:}{{\epsilon\:}}_{0}}\) (1)
where ε is the relative permittivity of the perovskite material, ε0 is the vacuum permittivity, e is the elementary charge, and L is the thickness of the perovskite active layer. For the electron-only devices (Supplementary Fig. 3a), the VTFL values for untreated and PBP-treated perovskite devices are 0.25 and 0.16 V, respectively. The electron trap state densities are calculated to be 5.2 × 1015 and 3.3×1015 cm− 3, showing a reduction of 1.9 × 1015 cm− 3 after PBP treatment. For the hole-only devices (Supplementary Fig. 3b), the VTFL values for untreated and PBP-treated perovskite devices are 0.64 and 0.26 V, respectively. The corresponding hole trap state densities are 1.3 × 1016 and 5.4 × 1015 cm− 3, representing a decrease of 7.6 × 1015 cm− 3 after PBP modification. The significant reduced nt of the PBP-treated perovskite devices indicates that the introduction of PBP can effectively improve the crystalline quality and mitigate defects in the perovskite films (Fig. 2b). Next, the µ was calculated by using the Mott-Gurney formula:
µ = \(\:\frac{8}{9}\)JDεε0\(\:\frac{{L}^{3}}{{V}^{2}}\) (2)
where JD is the current density, and V is the applied voltage. For the single electron devices (Supplementary Fig. 4a), the electron mobility increased from 1.17 × 10− 3 cm2 V− 1 s− 1 for untreated perovskite devices to 1.70 × 10− 3 cm2 V− 1 s− 1 for PBP-treated perovskite devices, an enhancement of 0.53 × 10− 3 cm2 V− 1 s− 1. For the hole-only devices (Supplementary Fig. 4b), the hole mobility of the PBP-treated perovskite film is 1.65 × 10− 3 cm2 V− 1 s− 1, much higher than that of the untreated film (9.14 × 10− 4 cm2 V− 1 s− 1), indicating faster carrier transport in PBP-treated devices. Additionally, the hole-to-electron mobility ratio in untreated and PBP-treated devices is 0.78 and 0.97, respectively, suggesting more balanced carrier transport in PBP-treated perovskites. This balance minimizes recombination and asymmetric transport losses in the devices.
Time-resolved photoluminescence (TRPL) spectroscopy was used to analyze the carrier dynamics and recombination kinetics in perovskite films. As depicted in Supplementary Fig. 5, a double exponential decay model was used to fit the TRPL decay curves, obtaining carrier lifetime parameters. Detailed fitting parameters are listed in Supplementary Table 1. For the untreated perovskite film, the fast recombination process τ1 is 50.0 ns, while for the PBP-treated perovskite film, τ1 increases to 102.3 ns, suggesting that the shallow traps are reduced and carriers can exist for a longer duration without undergoing fast recombination in the PBP-treated film. Additionally, Compared to the untreated film, which has a slow recombination process τ2 of 204.3 ns, the PBP-treated film exhibits a significantly higher τ2 of 563.0 ns, indicating a reduction in deep defects and more efficient carrier transport. Furthermore, the average decay lifetime of the untreated perovskite devices is 187.3 ns, increasing to 550.2 ns for the PBP-treated devices, which implies a reduced defect density and suppressed non-radiative recombination26. This indicates that the introduction of PBP into perovskite films effectively improves charge separation efficiency and minimizes charge recombination losses.
Sequentially, electrochemical impedance spectroscopy (EIS) was performed to study the impact of PBP treatment on carrier transport and recombination mechanisms in PerSCs. Supplementary Fig. 6 shows the Nyquist plots obtained from the EIS measurements of untreated and PBP-treated PerSCs under dark conditions and zero bias. The equivalent circuit model used for fitting is provided as the inset of Supplementary Fig. 6. The Nyquist plots include a semicircle in the high-frequency region and an arc in the mid-frequency region. The semicircle in the high-frequency region reflects charge transfer resistance (Rct), while the arc in the mid-frequency region reflects recombination resistance (Rrec). The fitting parameters are listed in Supplementary Table 2. With the reduction in Rct from 9.6 × 104 Ω in untreated perovskite solar cells to 6.3 × 104 Ω after PBP treatment, the interfacial charge transport efficiency is improved, enabling carriers to traverse the interface more quickly and effectively. The increase in Rrec from 0.76 × 106 Ω in untreated PerSCs to 1.12 × 106 Ω in PBP-treated PerSCs indicates a more efficient charge separation process, resulting in reduced charge recombination27. Supplementary Fig. 7 displays the current density-voltage (J-V) characteristics of PerSCs tested under dark conditions. The dark current density of PBP-treated PerSCs is significantly lower than that of untreated PerSCs, suggesting reduced leakage current paths and fewer recombination centers28. Subsequently, to investigate the carrier dynamics in devices, we measured the J-V characteristics under varying light intensities from 1 to 100 mW/cm2. The relationship between light intensity and VOC can be used to assess the recombination mechanisms in PerSCs. Theoretically, the slope of the curve corresponds to the ideality factor (n), which can be calculated using the following formula29:
n = \(\:\frac{e}{{K}_{B}T}\frac{d{V}_{OC}}{dlnI}\) (3)
where KB is the Boltzmann constant, T is the absolute temperature, I is the light intensity. When n approaches 1, it indicates that bimolecular radiative recombination is the dominant recombination mechanism. When n approaches 2, it implies that Shockley-Read-Hall (SRH) recombination and non-radiative recombination caused by trap states are dominant. As shown in Supplementary Fig. 8a, the value of n decreases from 1.74 in untreated PerSCs to 1.61 in PBP-treated PerSCs, reflecting a significant suppression of charge recombination and a reduction in defect state density. Next, we investigated the trend of JSC with varying light intensities to further verify the carrier recombination mechanisms. Generally, the relationship between JSC and light intensity can be described by the following formula: JSC ∝ \(\:{P}_{light}^{{\alpha\:}}\), where α is an exponent related to the recombination mechanism. An α value less than 1 indicates the presence of defect-assisted non-radiative recombination. As illustrated in Supplementary Fig. 8b, the α values for untreated and PBP-treated perovskite solar cells are 0.97 and 0.99, respectively, indicating that both devices exhibit defect state-induced SRH recombination.
On the basis of the above-mentioned results, we fabricated inverted WBG perovskite devices to study the impact of PBP treatment on device performance. The optimal concentration of PBP was found to be 0.5 mg/mL (Supplementary Fig. 9 and Supplementary Table 3). Figure 2c describes the J-V curves of the best-performing untreated and PBP-treated PerSCs. The corresponding optimal performance parameters are given in Supplementary Table 4. The untreated PerSC achieves a VOC of 1.29 V, a JSC of 17.59 mA cm− 2, an FF of 77.6%, and a PCE of 17.73%. In comparison, the optimized PBP-treated PerSC exhibits a higher VOC of 1.33 V, a JSC of 17.72 mA cm− 2, and an improved FF of 81.2%, yielding a high PCE of 19.08%. The external quantum efficiency (EQE) spectra of the PerSCs are shown in Fig. 2d. The integrated current density for the untreated PerSC is 16.73 mA cm− 2, while that for the PBP-treated PerSC is 17.34 mA cm− 2, which are consistent with the JSC values obtained from the J-V curve. The bandgap of the target device is determined to be 1.79 eV by differentiating its EQE spectrum, which exhibited an inflection point at 691 nm (Supplementary Fig. 10). Figure 2e makes a clear comparison of energy loss. Energy losses are 0.5 and 0.46 eV for untreated and PBP-treated PerSCs, respectively. Supplementary Fig. 11 presents the statistical photovoltaic parameters of 15 individual PerSCs from different batches. The significant improvement in the photovoltaic parameters of PBP-treated PerSCs can be attributed to the passivation effect of PBP on perovskite defects, which results in more efficient charge extraction and reduced recombination losses. These results demonstrate the reliability of our method in improving the performance of PerSCs. Supplementary Fig. 12 shows the operational stability of the PerSCs. The PBP-treated devices exhibit significantly enhanced stability, retaining 80% of their initial efficiency after 160 hours of continuous LED illumination. Additionally, we also achieved improved thermal stability for the PBP-treated PerSC, which retained 90% of its initial PCE after heating at 65°C for 600 h in the N2 atmosphere (Supplementary Fig. 13).
Design of ICL with high transmittance and conductivity
In PO-TSCs, the ICLs control charge collection and recombination from subcells and are critical in determining the performance of PO-TSCs. The design of high-quality ICLs aims to avoid or minimize reverse junction formation, parasitic light absorption, and junction resistance. Here, we designed a V2O5-based ICL structure with low optical/electrical loss and compared it with the traditional MoO3-based ICLs. To investigate the differences in optical and electrical properties of the V2O5-based and MoO3-based ICLs, we conducted a series of characterizations, including transmittance, conductivity, and hole mobility. First, we compared the UV-Vis spectra of the MoO3-based and V2O5-based ICLs. As displayed in Fig. 3a, within the wavelength range of 550 to 1100 nm, the transmittance of the MoO3-based ICLs is significantly lower than that of V2O5-based ICLs. The low optical loss of the V2O5-based ICLs in the NIR region indicates that the front junction perovskite subcell and the ICL have lower utilization efficiency for the NIR part of the solar spectrum, favoring the transmission of photons in the NIR region. This contributes to the enhancement of current in the organic bottom cells, which promotes current matching between the front and rear subcells in the PO-TSCs.
Regarding conductivity (σ), both MoO3-based and V2O5-based ICLs show a linear relationship between current density and voltage, indicating the formation of good ohmic contacts (Fig. 3b)30. The σ of the ICLs was calculated using the formula31: σ = Jd/V, where d represents the thickness of the ICL, J represents the current density, and V represents the voltage. The V2O5-based ICLs show a higher σ value of 2.23×10− 3 mS cm− 1 compared with MoO3-based ICLs (1.84×10− 3 mS cm− 1). Obviously, the hole mobility of V2O5 is also superior to that of MoO3 (Fig. 3c). These results indicate that V2O5 can effectively facilitate rapid and efficient charge transport, which enhances hole selectivity32, 33.
The ultraviolet photoelectron spectroscopy (UPS) measurement of V2O5 shows the work function (WF) and the valence band maximum (VBM) of 5.10 eV and 7.87 eV (Supplementary Fig. 14). Then the conduction band minimum (CBM) of the V2O5 film is determined to be 4.72 eV based on its bandgap of 3.15 eV (Supplementary Fig. 15). Supplementary Fig. 16 illustrates the energy level diagram. The detailed energy diagrams following Fermi level alignment are depicted in Fig. 3d. A significant band bending occurs at the C60/MoO3 interface due to the considerable work function mismatch. This bending is expected to induce a large Schottky barrier at the interface, which hampers efficient charge recombination within the ICL. Replacing MoO3 with V2O5 shifts the energy levels upward, substantially lowering the barrier and facilitating barrier-free electron transport from the WBG perovskite front subcells into the ICL, where the electrons efficiently recombine with the holes from the organic rear subcells.
We selected the combination of PBDB-T-2F (PM6), BTP-eC9 and PC71BM as the photoactive layer for the organic subcells34, 35. The detailed chemical structures of the donor and acceptor materials are illustrated in Supplementary Fig. 17. UV-Vis absorption spectroscopy was then used to compare the absorbance and transmittance of MoO3 and V2O5 thin films. As seen in Supplementary Fig. 18, within the absorption range of 600 to 1100 nm, the absorption intensity of the MoO3 films is higher than that of V2O5 films. The higher absorption intensity of MoO3 films in the NIR region reduces the effectiveness of the organic subcell in capturing low-energy photons, leading to current mismatch in the PO-TSCs. Correspondingly, the transmittance of the MoO3 films is lower than that of V2O5 films. We then established an optical transfer matrix model to analyze the light field distribution within the absorber layers. The device structure used for the simulation is ITO/MoO3 or V2O5/PM6:BTP-eC9/TPMA/Ag. As presented in Fig. 3e, the MoO3-based and V2O5-based OSCs exhibit similar simulated JSC within the 300–600 nm range owing to the comparable absorption capacity of both for high-energy photons. However, the V2O5-based OSCs improve the utilization of low-energy photons in the near-infrared range of 600 to 1000 nm, leading to a higher simulated JSC, which is beneficial for the current balancing between the front and rear subcells in PO-TSCs.
Transient photovoltage (TPV) and transient photocurrent (TPC) experiments were performed to gain insight into the charge recombination and extraction kinetics process36. As for TPC decay lifetimes (Supplementary Fig. 19), V2O5-based OSCs exhibit a shorter extraction time of 1.91 µs compared to 2.13 µs for MoO3-based OSCs, suggesting enhanced charge transport capacity. Photogenerated carrier lifetimes obtained from TPV measurements are 5.26 and 8.27 µs for MoO3-based and V2O5-based OSCs, respectively, indicating charge recombination has been suppressed in the V2O5-based OSCs. The charge generation properties were characterized by measuring their exciton dissociation probabilities (Pdiss), which were estimated by surveying the photocurrent densities versus effective voltage curves. As indicated in Supplementary Fig. 20, the calculated Pdiss values were 94.47% for MoO3-based OSCs and 97.13% for V2O5-based OSCs, with the higher Pdiss partly explaining the improvements in JSC and FF37.
Utilizing AFM, we studied the surface morphology of the films. The AFM images (Supplementary Fig. 21) reveal stripe-like structures for both MoO3 and V2O5 films, likely caused by anisotropic growth during thermal evaporation38. Additionally, the MoO3 films show a calculated root mean square roughness (RMS) of 0.50 nm, while V2O5 films exhibit an RMS of 0.39 nm. The extremely low roughness and smooth surface morphology of V2O5 films facilitate the formation of well-ordered molecular arrangements and crystallization in the absorber layer. Subsequently, absorber layers were fabricated on ITO/V2O5 and ITO/MoO3 films to validate the above conclusions. The smoother surface with a roughness of 2.25 nm is achieved for the V2O5-based absorber layer, while a rougher surface with a roughness of 3.53 nm is observed for the MoO3-based absorber layer.
According to our previous work, the interface plays an important role in controlling the vertical phase distribution39. We then studied the vertical phase distribution of the absorber layer. The vertical phase distribution is relevant to miscibility between the donor and acceptor. We used the Flory-Huggins interaction parameter (χ) to quantize the miscibility of the donor and acceptor, where larger χ values indicate poorer miscibility40, 41. To calculate χ, we carried out the contact angle tests on water and diiodomethane, and the results are presented in Supplementary Fig. 22. Through the contact angles, we could obtain the surface tensions (γ) for donor and acceptor materials. The γ was calculated using the Owens-Wendt-Rabel-Kaelble (OWRK) method. The relevant calculation results are listed in Supplementary Table 5. The γ of the V2O5-based absorber layer (39.5 mN m− 1) is closer to that of the acceptor on V2O5 films (41.3 mN m− 1), suggesting that the acceptor phase tends to enrich on the surface of the absorber layer. This vertical phase distribution is conducive to more efficient charge transport. Then χ can be calculated using the formula:
χ = (\(\:\sqrt{{{\gamma\:}}_{D}}\)−\(\:\sqrt{{{\gamma\:}}_{A}}\))2 (4)
where D represents the donor, and A represents the acceptor. It was found that χ values were 0.2 and 0.3 based on MoO3 and V2O5, respectively. The decreased miscibility may avoid the excessive mixing of donor and acceptor, leading to the formation of the desired domain size and purity.
Based on the above analysis, we investigate the impact of V2O5 or MoO3 on the device performance. The corresponding J-V curves are exhibited in Supplementary Fig. 23, and the photovoltaic parameters are summarized in Supplementary Table 6. Figure 3f displays the J-V curves of the best-performing MoO3-based and V2O5-based OSCs. The MoO3-based OSCs show a PCE of 17.34% with a VOC of 0.844 V, a JSC of 26.75 mA cm− 2, and an FF of 76.8%. In comparison, the V2O5-based OSC exhibits a VOC of 0.859 V, a JSC of 27.29 mA cm− 2, an FF of 78.4%, and a PCE of 18.39%. Supplementary Fig. 24 presents the statistical photovoltaic parameters of 15 individual OSCs from different batches, showing significant enhancement in V2O5-based OSCs compared to MoO3-based OSCs. Moreover, we observed an enhanced EQE in V2O5-based OSCs from 300 to 850 nm (Fig. 3g). The integrated current density derived from the EQE curves increased from 25.67 mA cm− 2 for the MoO3-based device to 26.31 mA cm− 2 for the V2O5-based device.
Photovoltaic performance of Monolithic PO-TSCs
Encouraged by the reduced energy loss of subcells and minimized optical/electrical loss of ICL, we fabricated PO-TSCs with the architecture of ITO/NiOx/4PADCB/perovskite/EDADI/C60/PEI/ITO/V2O5 or MoO3/PM6:BTP-eC9:PC71BM/TPMA/Ag (Fig. 4a). We compared the performance of PO-TSCs prepared with different thicknesses of ICLs. The J-V curves of the PO-TSCs are depicted in Supplementary Fig. 25. Supplementary Table 7 summarizes the corresponding photovoltaics parameters. The control PO-TSCs yield a PCE of 23.20%, with a VOC of 2.07 V, a JSC of 14.41 mA cm–2 and an FF of 77.7%. In contrast, the optimized PO-TSCs exhibited an FF of 81.1%, a JSC of 14.68 mA/cm², a VOC of 2.10 V, and a PCE of 25.07% (Fig. 4b). As shown in Fig. 4c, the optimized PO-TSCs show good current matching, where the JSC values of 14.11 and 13.98 mA cm− 2 are integrated from the EQE spectra of the perovskite and organic subcells, respectively. For the control PO-TSCs, the integrated current densities are 13.97 and 13.56 mA cm− 2, respectively. Additionally, the steady-state power output of the optimized PO-TSCs was 25.02% at a bias of 1.77 V (Fig. 4d). The statistical photovoltaic parameters of 30 individual PO-TSCs from different batches confirmed the advantages of the PO-TSCs with V2O5-based ICL (Fig. 4e and Supplementary Fig. 26). The significant enhancement in the performance of PO-TSCs can be attributed to the reduced energy loss of perovskite subcells and the low electrical and optical loss of V2O5-based ICL. Figure 4f depicts the light field distribution simulated by the optical transfer matrix model. The simulated JSC values of the PO-TSCs with V2O5-based ICL are higher than those with MoO3-based ICL, which could be due to the higher light field intensity in the 650 to 850 nm wavelength range for the PO-TSCs with V2O5-based ICL.
Subsequently, we evaluated the stability of the PO-TSCs. The thermal stability of the unencapsulated PO-TSCs in a nitrogen atmosphere was verified. After 200 hours of storage in a nitrogen atmosphere at 65°C fixed temperature (Fig. 4g), the PCE of the optimized PO-TSCs decreased by 9%, while the PCE of the control PO-TSCs decreased by 31%. As depicted in Fig. 4h, we also assessed the operational stability of the PO-TSCs. The stability of the TSCs under continuous operation is mainly determined by that of the perovskite subcell. The optimized PO-TSCs still maintained 82% of their initial efficiency after 460 hours of continuous LED light illumination, while the efficiency of the control PO-TSCs decreased to 80% of their original value after 180 hours.