Deciphering the interplay between tin vacancies and free carriers in the ion transport of tin-based perovskites†
Received
3rd February 2025
, Accepted 10th April 2025
First published on 11th April 2025
Abstract
Mixed ionic-electronic conduction is a prevalent phenomenon in metal halide perovskites, having a critical impact in multiple optoelectronic applications. In Sn-based halide perovskites, their higher hole density ([p]) owing to the facile formation of Sn vacancies (VSn2−) induces substantial electronic transport differences versus their Pb-based analogues. However, the influence of [p] and VSn2− on their ionic transport properties remains elusive. Herein, the link between electronic and ionic transport is unravelled in a compendium of Sn-based perovskite compositions. Specifically, ionic and electronic conductivities are found to concomitantly rise with higher Sn content. Using a combination of electrical characterization techniques, a rise in [p] and VSn2− is demonstrated to increase mobile ion density, enhancing lateral ion migration and ionic conductivity. First-principles simulations reveal that [p] and VSn2− jointly lower the energy barrier for iodide migration from 0.38 eV to 0.12 eV. Chemical mapping techniques support these observations by identifying the bias-induced migration of iodide and formamidinium ions in compositions with higher [p] and VSn2−. These fundamental insights on the ionic-electronic coupling will enable next-generation of Sn-based perovskite technologies with improved performance and stability.
Broader context
Halide perovskites based on tin-lead mixtures offer unique infrared absorption properties that enable efficient all-perovskite tandem solar cells, an emerging photovoltaic technology combining affordability and high energy output. However, the low stability of these materials continues to hinder their progress towards practical applications. A key phenomenon that limits solar cell lifetime is the migration of constituent ions within the perovskite layer. Hence, understanding the fundamental mechanisms governing ion transport in halide perovskites is paramount to design more stable perovskite photovoltaics. While ion migration in lead-based perovskite has received extensive attention, the study of this process in tin-based analogues remains at its early stages. In this work, we employ a comprehensive set of characterization techniques and first-principles simulations to unveil the mixed ionic-electronic transport mechanism in a wide compositional range of tin-based perovskites. We demonstrate that ionic transport in these materials is governed by the combined effect of tin vacancies and electronic hole carriers in the perovskite. These crucial insights will not only inspire novel design rules for stable tin-based perovskite solar cells, but also for related optoelectronic applications beyond photovoltaics.
|
Introduction
Metal halide perovskites, represented by the ABX3 formula (A = methylammonium (MA+), formamidinium (FA+), and/or Cs+; B = Pb2+ and/or Sn2+; X = I−, Br− and/or Cl−), are promising semiconductors for a myriad of optoelectronic applications. The combination of their remarkable optical properties and their low formation energies result in high performance optoelectronic devices compatible with low-cost fabrication methods.1 However, the relatively soft crystal lattice allows halide ions to easily propagate under operational conditions, such as high temperature, illumination, or electrical bias.2,3 The migration of halide ions as vacancies or interstitials strongly influences the charge transport properties of perovskites, as evidenced by their high ionic conductivity (σion) and mobile ion density (nion) relative to their electronic conductivity (σelec) and charge carrier concentration, respectively.4–6 The mixed ionic-electronic conduction results in several unusual phenomena observed in perovskite devices, including hysteresis in current–voltage (IV) curves, transient responses to external stimuli, formation of charge extraction barriers at interfaces with contacts, and detrimental chemical reactions between mobile ions and contacts.7–12
A critical aspect determining the mixed conduction properties of halide perovskites is their composition.5,9,13 Particularly, the partial or complete substitution of Pb2+ by Sn2+ at the perovskite B-site induces a dramatic increase in the hole carrier concentration [p].14–20 Interestingly, the ion transport properties of Sn-based perovskites exhibit marked differences compared to their Pb counterparts. For instance, the phenomenon of halide segregation, where mixed iodide/bromide (I−/Br−) compositions separate into individual I−-rich and Br−-rich domains under illumination, is not observed in Sn-based perovskites compared to Pb analogues.21–23 Moreover, solar cells and field effect transistors fabricated from mixed Sn–Pb perovskites present suppressed hysteresis regardless of the voltage scan rate.24,25 While these reports suggest that ion transport is reduced by the incorporation of Sn2+, several other studies have highlighted the substantial role of mobile ions in the degradation and performance of Sn-based perovskite solar cells. Specifically, short-circuit current density (Jsc) losses in mixed Sn–Pb perovskite solar cells have been attributed to a screening of the built-in potential caused by a high mobile ion density (nion ∼ 1017 cm−3).8 Furthermore, mobile ion-induced hysteresis has been correlated to lower efficiency and reduced stability.26,27 Photo-induced iodide vacancies (VI+) and iodide interstitials (Ii−) are suggested to initiate the oxidation of I− to molecular iodine (I2), which in turn can detrimentally promote further Sn oxidation.28–31 Importantly, the majority of these reports investigate ion migration through indirect and/or qualitative methods (i.e., time-dependent photoluminescence mapping or hysteresis analysis in IV curves), while focusing on a limited number of Sn-based perovskite compositions. Therefore, a systematic quantification of ion transport properties considering the distinct defect chemistry of Sn-based perovskites in a comprehensive compositional space is yet lacking.
The defect chemistry of Sn-based perovskites is significantly influenced by Sn vacancies (VSn2−), which result from the facile oxidation of Sn2+ to Sn4+ and are therefore prevalent.17 Moreover, VSn2− are known to act as acceptor defects that increase [p].18,32 Considering the mixed ionic-electronic character of halide perovskites, we anticipate that formation of VSn2− and the consequent increase in [p] may impact the ion transport properties. Revealing this link could unfold overlooked transport phenomena in Sn-based perovskites.
In this work, we quantify the ionic-electronic transport properties for a broad selection of Sn-based perovskite compositions (ASnxPb1−xI3, where A = MA+, FA+, and Cs0.15FA0.85 and x = 0, 0.25, 0.5, 0.75 and 1). Our findings show that regardless of the A-site composition, the σion and σelec concomitantly increase as the Sn fraction rises. By tuning [p] and VSn2− in mixed Sn–Pb perovskites with the additive SnF2, we discover that the increase in the σion and mobile ion density (nion) correlates with the formation of VSn2− and the consequent rise in [p]. Utilizing multiple chemical mapping techniques, we identify that lateral bias-induced migration of I− and FA+ is promoted in samples with higher VSn2− and [p]. First-principles calculations reveal that the simultaneous presence of hole charge carriers and VSn2− reduce the activation energy barrier of I− migration from 0.38 eV to 0.12 eV. These results reveal that ion transport in Sn-based perovskites is strongly influenced by the joint effect of VSn2− and hole carriers, which is a critical aspect of the mixed ionic-electronic transport mechanism in this class of materials. The joint effect of VSn2− and hole carriers on ion transport presented herein describe a central aspect of the mixed ionic-electronic transport in these materials, enabling the design of robust Sn-based perovskite optoelectronics.
Results and discussion
Ionic and electronic conductivities of Sn-based perovskites
To understand how the Sn content influences the ionic transport properties of ASnxPb1−xI3 perovskites (where A = MA, FA or Cs0.15FA0.85; we test this last composition due to its relevancy to solar cells and transistors),24,25 we perform galvanostatic polarization measurements on devices with coplanar contacts under N2 atmosphere and dark conditions (device architecture: Au/Cr/ASnxPb1−xI3/Cr/Au; channel width: 150 μm). A schematic of the device is shown in Fig. 1a (inset), and fabrication details are provided in the Experimental section. Galvanostatic polarization has been widely exploited to quantify the σion and σelec of several Pb perovskites.4,5,33–35 The measurement consists of tracking voltage changes through time while applying a constant electrical current (Fig. 1a). Initially, mobile ions and electrons begin to flow, both contributing to the total conductivity of the system (σtot = σion + σelec). This defines a starting voltage V (t ≈ 0), which is used to estimate σtot. Since ions cannot be extracted through the ion-blocking contacts, these will progressively accumulate at the interface with the electrodes. The depletion of mobile ions will cause a decrease in σion (and thus a drop in σtot), resulting in a voltage surge, a consequence of Ohm's law. At t = a, all mobile ions are depleted, and the voltage signal stabilizes at a constant value V (t = a), indicating that the flowing current is now constituted only of electronic carriers. From V (t = a), σelec is calculated and therefore, σion can be inferred. Further details on the measurement assumptions, the effect of electrode metals (Fig. S1–S3, ESI†), and the duration of the galvanostatic polarization measurement (Fig. S4, ESI†) are provided in Note S1 (ESI†).
 |
| Fig. 1 (a) Schematic of a typical polarization curve from a mixed ionic-electronic conductor. The inset shows the device architecture used in these experiments. The calculation of σion is derived from σion = σtot − σelec (b) galvanostatic polarization curves of FASnxPb1−xI3 (x = 0.0, 0.25, 0.50, 0.75 and 1), the electrical current used for the measurement varied according to the σtot of each composition. (c) σelec and σion as a function of Sn fraction for FASnxPb1−xI3 samples. | |
Fig. 1b shows the galvanostatic polarization curves of FASnxPb1−xI3 (x = 0, 0.25, 0.5, 0.75, and 1). We found that all samples exhibited mixed conduction behavior regardless of the Sn content. This observation was further confirmed for different A-site compositions such as MASnyPb1−yI3 (y = 0, 0.25, 0.5, and 0.75; Fig. S5a, ESI†) and Cs0.15FA0.85SnzPb1−zI3 (z = 0, 0.5, and 1; Fig. S6a, ESI†), indicating that mixed conduction prevails irrespectively of the A-cation selection in Sn-based perovskites. From the curves in Fig. 1b, we extracted σion and σelec of FA-based perovskites as shown in Fig. 1c (σion and σelec, trends for MA- and Cs0.15FA0.85-based perovskites presented in Fig. S5b and S6b, ESI†). In agreement with previous reports, we detect an increase in σelec with higher Sn content (x, y, z > 0.5) for all the A-site compositions studied. In FASnxPb1−xI3, σelec increases approximately four orders of magnitude, spanning from 2.07 × 10−8 S cm−1 in FAPbI3 to 3.73 × 10−4 S cm−1 in FASnI3 (a full list of values is provided in Table S1, ESI†). We mainly attribute the observed trend in σelec to the higher density of VSn2− at larger Sn fractions and thus an increase in the [p].14,16,17 Additional factors that may play a minor role on the observed σelec trend (i.e., bandgap narrowing) are discussed in Note S2 (ESI†).
Intriguingly, in the FA-series (Fig. 1c), σion increases up to 2 orders of magnitude (from 1.64 × 10−7 S cm−1 to 2.07 × 10−5 S cm−1) when transitioning from the Sn-lean (x ≤ 0.5) to the Sn-rich compositions (x ≥ 0.75). We notice a similar trend for the MA- and Cs0.15FA0.85-compositions (Fig. S5 and S6, respectively, ESI†). Specifically, we observe that (i) σion is higher than σelec for Sn-lean compositions, and that (ii) σelec dominates the σtot relative to σion for Sn-rich compositions. By considering the significant [p] found in Sn-rich samples and that holes are more mobile than ions,36,37 it is reasonable to expect that σelec exceeds σion in these compositions. However, the higher σelec compared to σion does not infer that ion migration is reduced. Instead, it is apparent in Fig. 1c that σelec and σion are coupled as both rise when the Sn content increases. Based on the magnitude of σion, these results indicate that significant ion migration is happening in Sn-based perovskites, implying substantial ionic-electronic conduction in these materials.
Mixed ionic-electronic transport in Sn-based perovskites
To understand why σion is enhanced at higher Sn
:
Pb ratios, we initially consider how the Sn content regulates the grain size and crystallinity. Earlier reports on Pb perovskites show that higher lattice strain and smaller grain size (i.e., larger density of grain boundaries) correlate to a reduced activation energy for I− migration.38–41 Notably, both of these parameters are heavily influenced by the addition of Sn, as faster crystallization rates are usually observed.42,43 For the majority of the A-site compositions, top-view scanning electron microscopy (SEM) images reveal larger grain sizes at higher Sn fractions (Fig. S7 and S8, ESI†). However, in pure Sn-based compositions (FASnI3, MASnI3, and Cs0.15FA0.85SnI3) we identify the formation of pinholes, which further increase the grain boundaries density and hence, the possible ion migration pathways.40 X-Ray diffraction (XRD) patterns from several compositions (FASnxPb1−xI3 being the exception) present a higher intensity of the characteristic perovskite diffraction peak at ∼14° as the Sn content increases, suggesting enhanced crystallinity (Fig. S9, ESI†). Furthermore, we record a shift to larger 2θ values with added Sn in the lattice, suggesting a lattice contraction due to the smaller radius of Sn2+ compared to Pb2+ (Fig. S10a–c, ESI†).44 We further investigate XRD peak broadening by measuring their full width at half maximum (FWHM). Our results indicated that the FWHM follows an overall decreasing trend as the Sn content increases (Fig. S10d–f, ESI†). Lower FWHM values have been interpreted as evidence of reduced microstrain.45,46 Taken together, the alterations in morphology and crystallinity resulting from the substitution of Sn do not reveal any clear trends behind the observed high σion.
Given the non-obvious relationship between σion and microstructure in our samples, we sought to adjust [p] and VSn2− density (since these variables also correlates with increasing Sn content) without significantly altering the morphology and crystallinity. For this, we fixed the composition to FASn0.5Pb0.5I3 while varying the concentration of SnF2 (0, 2.5, 5, and 10 mol% respect to the SnI2 concentration). This additive has played a pivotal role in the development of state-of-the-art Sn-based perovskite devices.47,48 SnF2 lessens [p] by capturing Sn4+ impurities and reducing the formation of VSn2−, in Note S3 (ESI†) we explain in more detail the working mechanism of SnF2.49–51 Consistent with previous reports, we find that changes in average grain size, crystallinity and FWHM upon addition of SnF2 are much smaller compared to those observed with variations of Sn content (Fig. S11 and S12, ESI†).50,52
Fig. 2a shows the galvanostatic polarization curves of the SnF2 containing samples. As the SnF2 concentration increases, we observe a reduction in σelec and σion (Fig. 2b). After 5 mol% SnF2, σelec and σion present no significant changes. Since σelec = q[p]μh (where q is the elementary charge and μh represents the hole mobility), a lower σelec due to [p] compensation by SnF2 is consistent. We do not anticipate that variations in μh influence σelec, as μh is known to remain relatively constant across different SnF2 concentrations.50 Therefore, the reduction of σion appears to be coupled to the depletion of [p] and VSn2−, as these are the variables that are expected to change the most upon SnF2 addition.
 |
| Fig. 2 (a) Galvanostatic polarization curves from FASn0.5Pb0.5I3 samples with different SnF2 concentrations: 0, 2.5, 5, and 10 mol% (b) σelec and σion as a function of SnF2 concentration. | |
Analogous to σelec, the σion from a particular ionic species is derived from its charge (z), the mobile ion mobility (μion), and nion (σion = zμionnion).53 If a relationship exists between [p] (and consequently VSn2−) and σion upon SnF2 addition, then a coupling between [p] and nion is plausible. It is also possible that [p] influences μion; a more in-depth discussion is provided in Note S4 and Fig. S13 (ESI†). To test if [p] and nion are linked, we quantify both parameters as a function of SnF2 concentration by performing Mott–Schottky (MS) analysis. This measurement records the depletion layer capacitance (Cdepl) of semiconductor junctions as a function of an applied bias, where each potential is superimposed by an AC bias signal. The AC frequency determines what type of species (i.e., ionic or electronic) creates the charge density profile that causes the fluctuations of the depletion layer width. At high frequencies (104–106 Hz), MS analysis can estimate carrier concentrations exceeding 1016 cm−3 for perovskite films with a thickness ≈300 nm.54,55 According to values derived from Hall measurements in MA0.6FA0.4Sn0.5Pb0.5I3 composition (≈1017 cm−3), high-frequency MS represents a suitable option to determine carrier concentration in Sn-rich compositions,56 although we note, that the contribution of mobile ions to the high-frequency results (e.g. due to the much slower voltage sweep during the measurement) cannot be ruled out yet. More recently, it was demonstrated that the extraction of the nion can be performed at low frequencies (∼1 Hz).57 Assuming that one species is dominant (e.g. I− or VI+), the low-frequency capacitance relates to the formation of a depletion layer due to the displacement of the dominant species (a more detailed explanation on the capacitance–voltage behavior in perovskites is found in Note S5, ESI†).
We extract nion and [p] from the FASn0.5Pb0.5I3 samples using the previously selected SnF2 concentrations (details on sample preparation are provided in the Experimental section). In Fig. 3a, we show the low-frequency capacitance–voltage (C–V) data, where we set the AC signal to 4 Hz. From the linear dependency between C−2 and the voltage applied (Vapp), nion is estimated from eqn (1):57
|  | (1) |
where
ε0 is the vacuum permittivity and
ε is the perovskite relative permittivity. Our results indicate that as the SnF
2 concentration increases from 0 to 10 mol%, the
nion gradually reduces from 6.2 × 10
17 to 2.4 × 10
17 cm
−3. To corroborate the obtained
C–
V values of
nion, we conduct bias assisted charge extraction (BACE) measurements. BACE records the temporal evolution of charge density upon switching from open-circuit voltage (
Voc) to 0 V.
8Fig. 3b illustrates the charge density progression, which reveals two distinct features. The initial increase at ≈10
−6 s corresponds to a combination of the injected or doping-induced electronic carriers and the electrode charge. The subsequent rise at ≈10
−1 s is associated with the contribution of the mobile ions. The BACE-derived values exhibit
nion decreasing from 5.8 × 10
17 to 2.6 × 10
17 cm
−3 as the SnF
2 concentration increases. Although, a complete saturation of the transient was not observed, meaning that
nion could be underestimated. The magnitudes of the
nion from the BACE measurements are thus in excellent agreement with the low-frequency
C–
V experiments and in line with previously reported values.
8
 |
| Fig. 3 (a) Low-frequency MS analysis; the AC frequency used is 4 Hz with an amplitude of 10 mV. (b) BACE measurements; the light purple, light blue, and light green-colored regions indicate the contribution from the electrode charge, the electronic charges that were injected at Voc, and the mobile ions, respectively. These measurements were carried out with samples encapsulated with a UV-curated epoxy and dark conditions. (c) High-frequency MS analysis; the AC frequency used is 10 kHz with an amplitude of 10 mV. The samples fabricated for all of these measurements have the following architecture: ITO/PEDOT:PSS (30 nm)/FASn0.5Pb0.5I3 (300 nm)/C60 (25 nm)/BCP (7 nm)/Cu (100 nm), where the selected SnF2 concentrations are: 0, 2.5, 5, and 10 mol%. The capacitance–voltage measurements are carried under a N2 environment and dark conditions. (c) nion, p, σion, and σelec as a function of SnF2 concentration in a FASn0.5Pb0.5I3 perovskite. For this figure, σion, and σelec were derived from the galvanostatic polarization measurements, nion values were selected from the low-frequency MS analysis, and p was obtained from the high-frequency MS analysis. | |
Fig. 3c displays the high-frequency C–V measurements of the same compositions (acquired at 10 kHz). MS analysis shows that [p] drops from 5.85 × 1016 to 1.6 × 1016 cm−3 when the SnF2 concentration increases from 0 to 10 mol%. In Fig. 3d, we converge all the transport measurements performed (nion, σion, [p], and σelec) as a function of the SnF2 concentration. Our findings indicate that the reduction of σion and σelec upon SnF2 addition is caused by the corresponding decrease of nion and [p], respectively. This strongly suggests the formation of VSn2−, and the resulting increase in [p], modulate the ionic transport in Sn-based perovskites.
Atomistic insights on mixed ionic-electronic transport
In Sn-based perovskites, the modulation of [p] and nion by tuning SnF2 concentration is inherently linked to a change of VSn2− density. Hence, it is important to consider how both the absence of Sn2+ in the lattice and the consequent increase in electronic charge carriers influence the possible ion migration pathways at the atomic scale. To this end, we perform density functional theory (DFT) simulations on FAPb0.5Sn0.5I3 to study the influence of VSn2−, holes (h+), and their combined effect on the energy barrier of I− migration (computational details are provided in the Experimental section). We calculate the I− migration energy barriers in the presence of the following defects in the lattice: (i) VI+ (i.e., the pristine case), (ii) VI+ and VSn2−, (iii) VI+ and h+, and (iv) VI+, VSn2−, and 2h+ (Fig. 4a). DFT calculations predict an energy barrier of 0.38 eV for the pristine case (Fig. 4b). This value is closely aligned with previously reported energy barriers in Pb perovskites,58 suggesting that Sn2+ ions do not significantly alter the I− migration dynamics. In contrast, the introduction of VSn2− increases the barrier up to 0.69 eV. This is in agreement with previous reports and has been attributed to a substantial rearrangement of the lattice structure caused by the VSn2−.25 Nonetheless, the energy barrier is reduced to 0.33 eV when we consider the effect of h+ in the calculations. The barrier drops even further to 0.12 eV when 2h+ and VSn2− are involved. Given the p-type character of Sn-based perovskites, we propose that excess of h+ will likely remain localized near the VSn2− vicinity.59 The positive charge of h+ could facilitate the migration of the negatively charged I− due to attractive electrostatic interactions, shortening the migration pathway (Fig. 4c).
 |
| Fig. 4 (a) Theoretical model of the iodide migration pathway in a FAPb0.5Sn0.5I3 lattice under four scenarios: (i) VI+ (i.e. the pristine case), (ii) VI+ and VSn2−, (iii) VI+ and h+, and (iv) VI+, VSn2−, and 2h+. (b) The energy barriers for iodide migration of each proposed scenario. (c) Relative energy landscapes of the system during iodide migration, the x-axis represents the iodide migration step during its transition towards an adjacent VI+. | |
Furthermore, to elucidate the effect of spatial inhomogeneity (regular features in real polycrystalline films) it is important to consider how other defects beyond VSn2− influence the energy barrier for I− migration (Fig. S15, ESI†). We find that I− and Sn2+ interstitial defects (Iint and Snint, respectively) raise the energy barrier above 1 eV, with a highest energy barrier achieved with Iint. Alternatively, I− on Sn2+ sites (ISn) and Sn2+ on I− sites (SnI) defects present energy barriers below 1 eV, where SnI defects show particularly low energy barriers of 0.36 eV. These results suggest that regions where I− poor conditions exist (i.e. regions where Ii is less likely to form but more likely to develop SnI defects) could similarly facilitate I− migration. Additionally, we also evaluate the effect of the surface (i.e., crystal termination such as that potentially found on grain boundary or pinhole) on the formation of several defects by calculating the difference between the defect formation energy in the bulk and on the surface (ΔΔH = ΔHbulk − ΔHsurf) (Fig. S16, ESI†). The ΔΔH of VI and VSn, the defects associated with I− mobility in our calculations, are notably the lowest; suggesting that formation of such defects is potentially favorable in the bulk and surface. While we acknowledge that sample heterogeneity has a significant impact on the ion transport dynamics, the results presented herein emphasize the role of VSn and h+ in the I− migration in Sn-based perovskites, which is in good agreement with the experimental observations.
Visualizing ion migration
To visualize the effect of mobile ions in the FASn0.5Pb0.5I3 samples with different SnF2 concentrations, we perform hyperspectral PL imaging upon biasing lateral devices (Fig. 5a). We track the intensity changes of the PL peak (990 nm) as a function of time while applying a constant bias of 30 V. To discard any light-soaking effect from the excitation source, we initially track the PL without any bias (Fig. S17a, ESI†). While the excitation light does not produce substantial changes in PL in most of the samples, a more detailed comment on this is given in Note S6 (ESI†). In the first minute of biasing, the channel from the 0 mol% SnF2 sample presents a 2% increase in the PL intensity. After 5 minutes, the emission reduces at the centre of the channel and a more intense signal appears near the positive electrode. During the following 10 minutes, the PL signal from the channel decreases 6% of the initial state, while the emission from the anode side increases 1%. PL quenching originates from the propagation of mobile ions towards the contacts, leaving behind an area with high defect density that enhances non-radiative recombination, as demonstrated by earlier reports.24,60 Furthermore, the increase in PL intensity observed near the anode is consistent with the accumulation of mobile anionic species (i.e., I−) and the passivation of ionic vacancies in this region. The 5 mol% SnF2 sample similarly reveals a slight increase of 4% in the emission upon biasing. After 3 min, the PL intensity remains almost unchanged for the duration of the measurement. The 2.5 and 10 mol% SnF2 samples present a similar behaviour as the 5 mol%, sustaining a relatively constant PL emission during the biasing (Fig. S17b, ESI†). These opposing responses to the bias could be related to the higher nion found in the samples without SnF2 (Fig. 3c). In this case, a higher density of defects is created since there are more mobile ions susceptible to the applied electric field, causing a faster reduction of the PL emission in the channel.
 |
| Fig. 5 (a) Hyperspectral PL maps from the Au/Cr/FASn0.5Pb0.5I3 (channel length is 300 μm)/Cr/Au lateral devices upon 30 V bias, the SnF2 concentrations are: 0 mol% and 5 mol%. We tracked the PL emission from mainly two areas: the channel and the anode, these are indicated by the pink and orange squares on the 0 min frames for each sample. Each frame was recorded every 1 minute. For clarity, we only displayed 5 frames per sample. The rest of the data set can be found in Fig. S10b (ESI†). (b) SEM images (top panel) and I EDS elemental maps (bottom panel) from the 0 mol% sample after 30 V bias for 1 hour. The white square highlights the damaged region after biasing. ToF-SIMS x–z depth profile maps of: (c) PbI+ and (d) CH5N+ fragments from the 0 and 5 mol% SnF2 samples before and after bias. | |
To confirm if ionic species are being mobilized by the applied electric field and their possible composition, we perform elemental mapping using energy-dispersive X-ray spectroscopy (EDS) and time-of-flight-secondary ion mass spectroscopy (ToF-SIMS). For both techniques, we track the changes of different species of interest before and after biasing the 0 mol% and 5 mol% SnF2 samples at 30 V bias for 1 hour. After biasing, we notice an evident morphology change near the cathode region that is only present in the 0 mol% SnF2 sample. The EDS mapping reveals a significant absence of iodine in the degraded region (Fig. 5b). Interestingly, we also observe a decrease of Sn in the damaged region (Fig. S18a, ESI†). The absence of I and Sn are consistent with the I− migration-induced formation of I2 and the subsequent degradation of the Sn–Pb perovskite (a more in-depth discussion is provided in Note S7, ESI†). In comparison, the 5 mol% film do not show significant variations in the I, Sn, and Pb content along the channel (Fig. S18b, ESI†). Furthermore, we employ ToF-SIMS depth profiling (3D chemical mapping) to elucidate variations in the distribution of the perovskite constituent secondary ions within the channel and under the electrodes. The process takes place via a non-interlaced depth profiling method employing alternating imaging and sputtering cycles (employing a 30 keV Bi3+ primary ion beam and a 20 keV Ar2000+ sputtering ion beam). The secondary ion images presented in Fig. 5c and d represent an x–z cross-section of the channel between the electrodes as well as the edges of the positive and negative electrodes. An illustration of a complete x–y, x–z and y–z series of ion images are shown for reference in Fig. S19 (ESI†). ToF-SIMS mapping from the 0 mol% samples show a higher concentration of PbI+ near the anode vicinity after biasing (Fig. 5c). Other iodide-based fragments, such as Pb2I3+, and SnI+, display a similar behaviour (Fig. S20, ESI†), compatible with mobile iodide migration. Interestingly, we further detect a higher concentration of CH5N+ fragments near the cathode area, consistent with an accumulation of FA+ ions (Fig. 5d). In contrast, the accumulation of these species near the electrodes is not visible after biasing the 5 mol% samples (Fig. S21, ESI†). The vertical extensions of the PbI+ and CH5N+ secondary ions at the outer edges of the channel represent material beneath the Au and Cr electrode regions which are revealed during depth profiling (this is illustrated using the Au3+ and Cr+ secondary ions for an example system in Fig. S22, ESI†). Additionally, we examine the potential for F− migration after the application of bias. The depth profile maps show an accumulation of F− at the top and bottom surfaces, as shown in previous research.61 However, we do not observe a preferential accumulation of F− at either of the contacts, suggesting that F− migration is an unlikely process (Fig. S23, ESI†). Based on these results, we identify that the migration of I− and FA+ in mixed Sn–Pb perovskites correlates well with a higher concentration of VSn2− and h+.
Implications of the mixed ionic-electronic conduction in the bias stability of Sn–Pb perovskite solar cells
Finally, we investigate the implications of these findings on the stability of Sn-based perovskites solar cells under continuous forward biasing. We monitor the power conversion efficiency (PCE) from IV curves of SnPb perovskite solar cells (ITO/PEDOT:PSS/Cs0.25FA0.75Sn0.5Pb0.5I3/C60/BCP/Cu) upon forward biasing at 1 V under dark (Fig. 6). We test the bias stability of the device with perovskite compositions containing 0 mol% and 5 mol% SnF2 (the photovoltaic parameters of both devices are summarized in Fig. S24, ESI†). The aim of these experiments is to induce device failure by leveraging mechanisms primarily linked to ion migration, as other factors such as temperature or light result in more intricate degradation pathways. Importantly, several reports have shown that aging devices under forward biasing exhibits similar degradation dynamics compared to illumination and outdoor testing.62–64 Effects such as charge collection losses, defect growth, degradation of crystalline structure and electrochemical reactions at the interfaces can be reproduced with forward bias tests. Consequently, results from bias-induced degradation could provide a preliminary understanding on the stability of photovoltaic devices when subjected to real-world conditions, including variations in temperature and light exposure.
 |
| Fig. 6
IV curves from ITO/PEDOT:PSS/Cs0.25FA0.75Sn0.5Pb0.5I3/C60/BCP/Cu solar cells with (a) 0 mol% and (b) 5 mol% SnF2 recorded during 1 V bias stressing for 60 min. After resting the devices for 60 min under dark, another IV is acquired. (c) Normalized PCE as a function of time from the 0 mol% and 5 mol% SnF2 devices. The initial efficiencies are 10.1 and 15.5 for the devices with 0 and 5 mol%, respectively. | |
From the analysis of all the photovoltaic parameters, we notice that the current density (Jsc) experiences the most significant losses in both 0 mol% and 5 mol% SnF2 samples. The PCE from the device containing 0 mol% SnF2 decays exponentially, reaching 33% of its initial efficiency after 60 min (Fig. 6a). The 5 mol% sample shows a rapid PCE decrease during the first 5 minutes of biasing, after which the PCE remains relatively stable at 44% of the original PCE for the rest of the biasing period (Fig. 6b). These fast performance losses are associated with an accumulation of ions at the interfaces with the contacts, inducing a severe screening of the built-in potential.8,65 Notably, we detect a significant difference in the PCE after resting the devices for 1 hour under dark conditions. While the device with 0 mol% SnF2 presents a permanent loss of performance after the resting period, the 5 mol% SnF2 sample almost recovers to their initial efficiency (Fig. 6c). The irreversible degradation in the 0 mol% SnF2 device agrees with the more severe ion migration and the I and Sn loss recorded in our EDS experiments (Fig. 5b). The increase of nion and the concentration of VSn2− in the samples without SnF2 could favor detrimental chemical reactions with the transport layers or within the perovskite, resulting in a permanent degradation at the interfaces or the perovskite bulk (a more detailed discussion on the possible degradation mechanism under bias stress is provided in Note S8, ESI†). Therefore, these results demonstrate that efficiency losses attributed to bias stress in SnPb perovskites are considerable and are correlated to the mixed ion-electronic properties of these materials.
Conclusions
In summary, we investigate the mixed ionic-electronic transport properties of Sn-based perovskites and find that the σelec and σion undergo a simultaneous increase at higher Sn fractions. Our experimental results indicate that this behaviour is due to an enhancement in [p] and VSn2−, which is followed by a rise in nion and prominent lateral ion migration. This implies that the strong p-doping in Sn-based perovskites not only modulates the σelec but also the σion. First-principles simulations reveal that the combined effect of h+ and VSn2− facilitate I− migration. The excess of positively charged h+ near the vicinity of a VSn2− promotes the movement of the negatively charged I− through attractive electrostatic interactions, thereby reducing the migration barrier for I− migration. We verify these insights through elemental mapping, where we identify that I− and FA+ migration exacerbates at a higher concentration of VSn2− and h+. Overall, these results highlight the combined effect of vacancies and electronic carriers on the ion transport in Sn-based perovskites. Considering the tendency of these materials for self p-doping, we demonstrate that the relation between VSn2−, h+, and mobile ions has major implications for the bias stability of Sn-based perovskite solar cells. Hence, the detailed picture of the mixed ionic-electronic transport in Sn-based perovskites provided herein is of great relevance towards high performing and stable Sn-based perovskite technologies.
Experimental section
Materials
CsI (dry beads 99.999%), PbI2 (dry beads, 99.999%), SnI2 (99.99%), SnF2, bathocuproine (BCP), DMSO, DMF, and chlorobenzene (all anhydrous) were obtained from Sigma Aldrich. MAI, FAI, and MACl were provided by Greatcell. Cr (99.95%), Au (99.99%), and Cu (99.9%) pellets were purchased from Kurt J. Lesker. C60 was supplied by nanoC. PEDOT:PSS AI 4083 was obtained from Heraeus.
Perovskite solution preparation
All the perovskite solutions were prepared by weighting the respective precursor salts (CsI, MAI, MACl, FAI PbI2, SnI2, and SnF2, with their ratios corresponding to the stoichiometry of the desired composition) in a single glass vial. MACl was added for the FAPbI3 composition to stabilize the α-phase while only leaving traces of MA, the concentration of MACl used was 32 mg ml−1.66 For the galvanostatic polarization measurements from the MA, FA, and CsFA series, the concentration of SnF2 was set to 10 mol% with respect to the concentration of SnI2 of each composition. The precursors were dissolved in a mixture of DMF:DMSO and stirred for 1 hour before using the solution. The DMF
:
DMSO ratio varied depending on the composition, for pure Pb compositions a 4
:
1 ratio was used, for the rest of the compositions a 3
:
1 ratio was selected. The concentration of all the solutions used to prepare the samples for the galvanostatic polarization, hyperspectral PL, and capacitance voltage measurements was set to 1.2 M. For the fabrication of the perovskite solar cell, a 1.6 M solution was prepared. All solutions were stirred for 1 h. Solution preparation was carried inside a N2 glovebox with trace levels of O2 (<0.1 ppm).
Film characterization
SEM images were taken with a Zeiss Auriga microscope operated at 5 keV. X-Ray diffraction patterns were acquired with a Bruker D8 advanced diffractometer employing Cu-Kα radiation. SEM-EDS based experiment were performed in Thermo Scientific™ (FEI) DualBeam SEM-FIB Helios G5 operated at 10 keV, and equipped with UltraDry EDS Detector with 129 eV energy resolution. Additionally, some SEM images were taken with a Zeiss Auriga microscope (5 keV). X-Ray diffraction (XRD) patterns were acquired with a Bruker D8 advanced diffractometer employing Cu-Kα radiation.
ToF-SIMS elemental analysis
Analysis was conducted using a HybridSIMS instrument (IONTOF, GmbH). Dual-beam dynamic SIMS was conducted using a 20 keV Ar2000 gas cluster ion beam (GCIB) with a target current of 1 nA as a sputter beam and 30 keV Bi3+ analysis beam with a pulsed target current of ∼0.03 pA. Non-interlaced depth profiling was employed, whereby the sputter and analysis ion beams operate at alternating intervals with 5 frames of sputtering per 5 analysis frame and a 0.5 s pause between the sputtering and analysis. A region of 350 μm × 350 μm was sputtered, and a 200 μm × 200 μm region was analyzed with pixel density 256 × 256. The data was collected in positive polarity for 393 seconds. The biased samples were subjected to an electric field of 27 V for 1 hour inside a N2 glovebox.
Sample preparation for galvanostatic polarization and hyperspectral PL measurements
Glass substrates (Xinyan Technology, 2.5 × 2.5 cm2) were sequentially sonicated in a Hellmanex aqueous solution, deionized water, acetone, and isopropanol for 10 min each. Substrates were dried with N2 and treated with UV ozone for 15 min. A two-step spin-coating procedure of 1000 rpm for 10 s and 4000 rpm for 40 s was used for the deposition of the perovskite films. Chlorobenzene (100 μl) is used as antisolvent in all compositions, the timing at which CB is added varied according to the composition. After the antisolvent quenching, most compositions were annealed at 100 °C for 10 min except for the MAPbI3, and FAPbI3 sample, which were annealed at 100 and 150 °C for 30 and 10 min, respectively. The perovskite deposition was followed by the sequential thermal evaporation of Cr (10 nm) and Au (60 nm), this was carried under high vacuum by an Angstrom Engineering system. All processing was performed inside a N2 glovebox with O2 levels below 0.1 ppm and continuous purging.
Sample preparation for the capacitance–voltage measurements and solar cell fabrication
Pre-patterned ITO/glass substrates (Xinyan Technology, 10/15 Ω sq−1, CN) were sequentially sonicated in a Hellmanex aqueous solution, deionized water, acetone, and isopropanol for 10 min each. Substrates were dried with N2 and treated with UV ozone for 15 min. A solution of PEDOT
:
PSS and isopropanol (1
:
1) was coated on the clean ITO at 4000 rpm for 50 s, then annealed at 150 °C for 10 min. Then, samples were transferred immediately to a N2 glovebox with O2 levels below 0.1 ppm. A two-step spin-coating procedure of 1000 rpm for 10 s and 4000 rpm for 40 s was used for the deposition of the perovskite films. CB was added at the 15th s of the second step. Samples with the compositions FASn0.5Pb0.5I3 and Cs0.25FA0.75Sn0.5Pb0.5I3 are annealed at 100 °C and 120 °C for 10 min, respectively. Subsequently, C60 (30 nm) and BCP (7 nm) were thermally evaporated at a rate of 0.1 A s−1 under high vacuum by using an Angstrom Engineering system. Finally, 100 nm of Cu stripes were thermally evaporated to result in devices with an area of 0.1 cm2.
Galvanostatic polarization measurements
Samples were transfer from inside the glovebox to a Nextron chamber with a 2 probe configuration, with ultra-pure N2 (99.999%) being continuously injected to the chamber. The temperature of the sample was actively fixed to 40 °C with a Peltier module controlled by the software provided by the company. Before the measurement started, a waiting time of 10 min was given to stabilize the temperature of the sample. The electrical measurements were performed with a Keithley 4200 semiconductor parameter analyzer. The injected current was selected to produce low voltages of ∼100 mV. At such voltages, we reduce the risk of any irreversible reaction (in the perovskite bulk or the perovskite/contacts interface) that could be induced by the constant flow of current. The σtot is calculated from the following equation:
, where I is the applied current, Vt=0 is the voltage measured at the initial point in time (t), L is the channel length (150 μm), d is the thickness of the perovskite (∼300 nm), and w is the width of the electrode (3 μm). The σelec is estimated from the following equation:
, where Vt=a is the voltage after some time a, which was characterized by the stabilization of the voltage. The stabilization time varied according to the composition, but it was typically observed in the range of 102 seconds. The σion was deduced from the equation: σion = σelec − σtot.
Capacitance–voltage measurements
The capacitance–voltage measurements were performed with a Fluxim Paios characterization suite inside a N2 glovebox. The AC frequency for the low frequency measurements was set to 4 Hz with an amplitude of 20 mV. The measurement was executed in a voltage range from −0.2 to 0.3 V with scan speed of 0.7 V s−1. The nion was extracted from the linear region of this range by the following formula:
.57 The value of ε = 37 was extracted from the geometrical capacitance obtained from capacitance vs. frequency plots (Fig. S14, ESI†). The relative permittivity is deduced from the geometrical capacitance (the capacitance at 10 kHz) with the following relation:
.54 For the high frequency measurements, an AC signal of 10 kHz with an amplitude of 10 mV was selected. The chosen voltage range was from −0.2 to 0.8 with a scan rate of 0.7 V s−1. The p was estimated from the Mott Schottky analysis given by:
, where Vbi is the built-in voltage. All measurements are taken under N2 and dark conditions.
BACE measurements
In dark BACE, the device was initially held at a voltage close to the open-circuit voltage in the dark, where the injected charge equals the short-circuit current. After a pre-set delay time, a bias of 0 V was applied to extract the injected and capacitive charge in the device. The delay times for the fresh devices were chosen to be typically five times longer than the extraction time of charges observed under the collection bias (typically ∼5–10 s) to allow ionic charges to distribute throughout the active layer. The extracted charge was obtained by integrating the current transient and the charge carrier density by dividing the total charge by the elementary charge and the cell volume. The measurements were performed using the ion characterization system from FastChar UG.
Hyperspectral PL measurements
Hyperspectral PL imaging was carried out with a IMA hyperspectral microscopy system (photon etc). A bias of 30 V was applied to samples in the dark during every time step, after which bias was stopped and PL was acquiered by exciting the sample with a continuous-wave 405 nm laser. PL images where then normalized to t = 0 min to visualize changes in PL induced by ion migration. Samples were encapsulated with a glass slide and a polyurethane sealant.
Solar cell characterization
A Xe arc lamp solar simulator (ABET Sun 2000, class A) calibrated to ±5% of AM1.5G between 350–1100 nm was used to obtain current–voltage curves. To record the J–V response, a Keithley 2400-series source-measure unit was used with a scan speed of 0.15 V s−1. For the stability measurements, a J–V was recorded for specific times while devices were under 1 V forwards bias. For the entire duration of the measurement, devices were kept under a N2 glovebox (<0.1 ppm).
DFT calculations
First-principles calculations were employed using density functional theory (DFT) based on plane-wave basis set and the projected augmented wave method, as implemented in the VASP package.67,68 SCAN+rvv10 type meta-GGA with vdW functional was used to describe the exchange–correlation functionals.69,70 All starting geometries are generated based on the experimental lattice parameters of FAPbI3. A cut-off energy of 400 eV for the plane-wave basis sets and 4 × 4 × 4 for gamma-centered k-mesh are used for the self-consistent field calculations and geometry optimizations. In geometry optimizations, the ionic positions and cell dimensions were allowed to relax, using a conjugate gradient algorithm, until all residual forces are smaller than 0.02 eV Å−1.
Migration barriers were determined from the energy profiles of iodine migration along the reaction path from one site towards the nearest iodine-vacancy defect. The initial and final structures were relaxed first, followed by a linear interpolation scheme to generate intermediate structures along the migration pathway. The energy profiles were calculated using the nudged elastic band and constrained energy minimization methods. 16 grid points were considered for energy profile calculations. The energy barriers associated with the iodide migrations were determined from the energy difference between the minimum energy and the saddle point.
Author contributions
L. H. H fabricated all the samples, measured XRD and SEM, and performed the galvanostatic polarization, capacitance–voltage, and solar cells characterization measurements. L. L. assisted with sample preparation and the hyperspectral PL measurements. A M. K., M. P., D. J. S. conducted the ToF-SIMS measurements. I. Y. performed the DFT calculations. N. K. and M. S. performed the BACE characterizations. B. V. and S. W. assisted with SEM and EDS measurements. M. G. R. participated in the development of the experimental set-ups. L. H. H., L. L., and D. B. conceived the idea, directed the project, and wrote the manuscript. All authors read and commented on the manuscript.
Data availability
The data supporting the article have been included as part of the ESI.†
Conflicts of interest
The authors declare no competing interests.
Acknowledgements
Part of the research reported in this publication was supported by funding from King Abdullah University of Science and Technology (KAUST) - Center of Excellence for Renewable Energy and Storage Technologies, under award number 5937. All authors thank the KAUST Solar Platform (KSP) for granting access to the characterization and device fabrication facilities and equipment employed herein. M. S. acknowledges the Vice-Chancellor Early Career Professorship Scheme from CUHK for funding.
References
- B. A. Rosales, K. Schutt, J. J. Berry and L. M. Wheeler, ACS Energy Lett., 2023, 8, 1705–1715 CrossRef CAS.
- C. Eames, J. M. Frost, P. R. F. Barnes, B. C. O’Regan, A. Walsh and M. S. Islam, Nat. Commun., 2015, 6, 7497 CrossRef CAS PubMed.
- E. Mosconi and F. De Angelis, ACS Energy Lett., 2016, 1, 182–188 CrossRef CAS.
- T.-Y. Yang, G. Gregori, N. Pellet, M. Grätzel and J. Maier, Angew. Chem., Int. Ed., 2015, 54, 7905–7910 CrossRef CAS PubMed.
- G. Y. Kim, A. Senocrate, Y.-R. Wang, D. Moia and J. Maier, Angew. Chem., Int. Ed., 2021, 60, 820–826 CrossRef CAS PubMed.
- R. A. Kerner and B. P. Rand, J. Phys. Chem. Lett., 2018, 9, 132–137 CrossRef CAS PubMed.
- P. Calado, A. M. Telford, D. Bryant, X. Li, J. Nelson, B. C. O’Regan and P. R. F. Barnes, Nat. Commun., 2016, 7, 13831 CrossRef CAS PubMed.
- J. Thiesbrummel, V. M. Le Corre, F. Peña-Camargo, L. Perdigón-Toro, F. Lang, F. Yang, M. Grischek, E. Gutierrez-Partida, J. Warby, M. D. Farrar, S. Mahesh, P. Caprioglio, S. Albrecht, D. Neher, H. J. Snaith and M. Stolterfoht, Adv. Energy Mater., 2021, 11, 2101447 CrossRef CAS.
- V. M. Le Corre, J. Diekmann, F. Peña-Camargo, J. Thiesbrummel, N. Tokmoldin, E. Gutierrez-Partida, K. P. Peters, L. Perdigón-Toro, M. H. Futscher, F. Lang, J. Warby, H. J. Snaith, D. Neher and M. Stolterfoht, Sol. RRL, 2022, 6, 2100772 CrossRef CAS.
- J. Carrillo, A. Guerrero, S. Rahimnejad, O. Almora, I. Zarazua, E. Mas-Marza, J. Bisquert and G. Garcia-Belmonte, Adv. Energy Mater., 2016, 6, 1502246 CrossRef.
- D. Lin, T. Shi, H. Xie, F. Wan, X. Ren, K. Liu, Y. Zhao, L. Ke, Y. Lin, Y. Gao, X. Xu, W. Xie, P. Liu and Y. Yuan, Adv. Energy Mater., 2021, 11, 2002552 CrossRef CAS.
- D. Moia, I. Gelmetti, P. Calado, W. Fisher, M. Stringer, O. Game, Y. Hu, P. Docampo, D. Lidzey, E. Palomares, J. Nelson and P. R. F. Barnes, Energy Environ. Sci., 2019, 12, 1296–1308 RSC.
- S. Reichert, Q. An, Y.-W. Woo, A. Walsh, Y. Vaynzof and C. Deibel, Nat. Commun., 2020, 11, 6098 CrossRef CAS PubMed.
- M. A. Haque, L. H. Hernandez, B. Davaasuren, D. R. Villalva, J. Troughton and D. Baran, Adv. Energy Sustainability Res., 2020, 1, 2000033 CrossRef.
- T. Liu, X. Zhao, J. Li, Z. Liu, F. Liscio, S. Milita, B. C. Schroeder and O. Fenwick, Nat. Commun., 2019, 10, 5750 CrossRef CAS PubMed.
- R. Tounesi, L. H. Hernandez, L. Lanzetta, S. Jang, B. Davaasuren, M. A. Haque and D. Baran, ACS Appl. Electron. Mater., 2023, 6, 2826–2831 CrossRef.
- D. Meggiolaro, D. Ricciarelli, A. A. Alasmari, F. A. S. Alasmary and F. De Angelis, J. Phys. Chem. Lett., 2020, 11, 3546–3556 CrossRef CAS PubMed.
- L. Lanzetta, T. Webb, J. M. Marin-Beloqui, T. J. Macdonald and S. A. Haque, Angew. Chem., Int. Ed., 2023, 62, e202213966 CrossRef CAS PubMed.
- M. A. Haque, T. Zhu, L. H. Hernandez, R. Tounesi, C. Combe, B. Davaasuren, A.-H. Emwas, F. P. García de Arquer, E. H. Sargent and D. Baran, Cell Rep. Phys. Sci., 2023, 4, 101703 CrossRef CAS.
- L. Lanzetta, L. Gregori, L. H. Hernandez, A. Sharma, S. Kern, A. M. Kotowska, A.-H. Emwas, L. Gutiérrez-Arzaluz, D. J. Scurr, M. Piggott, D. Meggiolaro, M. A. Haque, F. De Angelis and D. Baran, ACS Energy Lett., 2023, 8, 2858–2867 CrossRef CAS.
- N. Li, Z. Zhu, J. Li, A. K.-Y. Jen and L. Wang, Adv. Energy Mater., 2018, 8, 1800525 CrossRef.
- K. O. Ighodalo, W. Chen, Z. Liang, Y. Shi, S. Chu, Y. Zhang, R. Khan, H. Zhou, X. Pan, J. Ye and Z. Xiao, Angew. Chem., Int. Ed., 2023, 62, e202213932 CrossRef CAS PubMed.
- S. Martani, Y. Zhou, I. Poli, E. Aktas, D. Meggiolaro, J. Jiménez-López, E. L. Wong, L. Gregori, M. Prato, D. Di Girolamo, A. Abate, F. De Angelis and A. Petrozza, ACS Energy Lett., 2023, 8, 2801–2808 CrossRef CAS PubMed.
- S. P. Senanayak, K. Dey, R. Shivanna, W. Li, D. Ghosh, Y. Zhang, B. Roose, S. J. Zelewski, Z. Andaji-Garmaroudi, W. Wood, N. Tiwale, J. L. MacManus-Driscoll, R. H. Friend, S. D. Stranks and H. Sirringhaus, Nat. Mater., 2023, 22, 216–224 CrossRef CAS PubMed.
- K. Dey, D. Ghosh, M. Pilot, S. R. Pering, B. Roose, P. Deswal, S. P. Senanayak, P. J. Cameron, M. S. Islam and S. D. Stranks, Energy Environ. Sci., 2024, 17, 760–769 RSC.
- C. Li, Z. Song, C. Chen, C. Xiao, B. Subedi, S. P. Harvey, N. Shrestha, K. K. Subedi, L. Chen, D. Liu, Y. Li, Y.-W. Kim, C.-S. Jiang, M. J. Heben, D. Zhao, R. J. Ellingson, N. J. Podraza, M. Al-Jassim and Y. Yan, Nat. Energy, 2020, 5, 768–776 CrossRef CAS.
- J. Thiesbrummel, S. Shah, E. Gutierrez-Partida, F. Zu, F. Peña-Camargo, S. Zeiske, J. Diekmann, F. Ye, K. P. Peters, K. O. Brinkmann, P. Caprioglio, A. Dasgupta, S. Seo, F. A. Adeleye, J. Warby, Q. Jeangros, F. Lang, S. Zhang, S. Albrecht, T. Riedl, A. Armin, D. Neher, N. Koch, Y. Wu, V. M. Le Corre, H. Snaith and M. Stolterfoht, Nat. Energy, 2024, 9, 664–676 CrossRef CAS.
- J. Wang, M. A. Uddin, B. Chen, X. Ying, Z. Ni, Y. Zhou, M. Li, M. Wang, Z. Yu and J. Huang, Adv. Energy Mater., 2023, 13, 2204115 CrossRef CAS.
- J. Sanchez-Diaz, R. S. Sánchez, S. Masi, M. Kreĉmarová, A. O. Alvarez, E. M. Barea, J. Rodriguez-Romero, V. S. Chirvony, J. F. Sánchez-Royo, J. P. Martinez-Pastor and I. Mora-Seró, Joule, 2022, 6, 861–883 CrossRef CAS PubMed.
- L. Lanzetta, T. Webb, N. Zibouche, X. Liang, D. Ding, G. Min, R. J. E. Westbrook, B. Gaggio, T. J. Macdonald, M. S. Islam and S. A. Haque, Nat. Commun., 2021, 12, 2853 CrossRef CAS PubMed.
- A. Alsulami, L. Lanzetta, L. Huerta Hernandez, D. Rosas Villalva, A. Sharma, S. P. Gonzalez Lopez, A.-H. Emwas, A. Yazmaciyan, F. Laquai, I. Yavuz and D. Baran, J. Am. Chem. Soc., 2024, 146, 22970–22981 CrossRef CAS PubMed.
- S. Gupta, D. Cahen and G. Hodes, J. Phys. Chem. C, 2018, 122, 13926–13936 CrossRef CAS.
- G. Y. Kim, A. Senocrate, T.-Y. Yang, G. Gregori, M. Grätzel and J. Maier, Nat. Mater., 2018, 17, 445–449 CrossRef CAS PubMed.
- A. Senocrate, I. Moudrakovski, G. Y. Kim, T.-Y. Yang, G. Gregori, M. Grätzel and J. Maier, Angew. Chem., Int. Ed., 2017, 56, 7755–7759 CrossRef CAS PubMed.
- A. Senocrate, I. Spanopoulos, N. Zibouche, J. Maier, M. S. Islam and M. G. Kanatzidis, Chem. Mater., 2021, 33, 719–726 CrossRef CAS.
- F. Peña-Camargo, J. Thiesbrummel, H. Hempel, A. Musiienko, V. M. Le Corre, J. Diekmann, J. Warby, T. Unold, F. Lang, D. Neher and M. Stolterfoht, Appl. Phys. Rev., 2022, 9, 021409 Search PubMed.
- M. García-Batlle, S. Deumel, J. E. Huerdler, S. F. Tedde, A. Guerrero, O. Almora and G. Garcia-Belmonte, ACS Appl. Mater. Interfaces, 2021, 13, 35617–35624 CrossRef PubMed.
- J. Zhao, Y. Deng, H. Wei, X. Zheng, Z. Yu, Y. Shao, J. E. Shield and J. Huang, Sci. Adv., 2017, 3, eaao5616 CrossRef PubMed.
- L. McGovern, I. Koschany, G. Grimaldi, L. A. Muscarella and B. Ehrler, J. Phys. Chem. Lett., 2021, 12, 2423–2428 CrossRef CAS PubMed.
- M. Ghasemi, B. Guo, K. Darabi, T. Wang, K. Wang, C.-W. Huang, B. M. Lefler, L. Taussig, M. Chauhan, G. Baucom, T. Kim, E. D. Gomez, J. M. Atkin, S. Priya and A. Amassian, Nat. Mater., 2023, 22, 329–337 CrossRef CAS PubMed.
- Y. Shao, Y. Fang, T. Li, Q. Wang, Q. Dong, Y. Deng, Y. Yuan, H. Wei, M. Wang, A. Gruverman, J. Shield and J. Huang, Energy Environ. Sci., 2016, 9, 1752–1759 RSC.
- F. Hao, C. C. Stoumpos, P. Guo, N. Zhou, T. J. Marks, R. P. H. Chang and M. G. Kanatzidis, J. Am. Chem. Soc., 2015, 137, 11445–11452 CrossRef CAS PubMed.
- G. Nasti, M. H. Aldamasy, M. A. Flatken, P. Musto, P. Matczak, A. Dallmann, A. Hoell, A. Musiienko, H. Hempel, E. Aktas, D. Di Girolamo, J. Pascual, G. Li, M. Li, L. V. Mercaldo, P. D. Veneri and A. Abate, ACS Energy Lett., 2022, 7, 3197–3203 CrossRef CAS.
- S. Kahmann, Z. Chen, O. Hordiichuk, O. Nazarenko, S. Shao, M. V. Kovalenko, G. R. Blake, S. Tao and M. A. Loi, ACS Appl. Mater. Interfaces, 2022, 14, 34253–34261 CrossRef CAS PubMed.
- D. Kim, J. S. Yun, P. Sharma, D. S. Lee, J. Kim, A. M. Soufiani, S. Huang, M. A. Green, A. W. Y. Ho-Baillie and J. Seidel, Nat. Commun., 2019, 10, 444 CrossRef CAS PubMed.
- K. Nishimura, D. Hirotani, M. A. Kamarudin, Q. Shen, T. Toyoda, S. Iikubo, T. Minemoto, K. Yoshino and S. Hayase, ACS Appl. Mater. Interfaces, 2019, 11, 31105–31110 CrossRef CAS PubMed.
- T. J. Macdonald, L. Lanzetta, X. Liang, D. Ding and S. A. Haque, Adv. Mater., 2023, 35, 2206684 CrossRef CAS PubMed.
- M. Pitaro, E. K. Tekelenburg, S. Shao and M. A. Loi, Adv. Mater., 2022, 34, 2105844 CrossRef CAS PubMed.
- M. H. Kumar, S. Dharani, W. L. Leong, P. P. Boix, R. R. Prabhakar, T. Baikie, C. Shi, H. Ding, R. Ramesh, M. Asta, M. Graetzel, S. G. Mhaisalkar and N. Mathews, Adv. Mater., 2014, 26, 7122–7127 CrossRef CAS PubMed.
- K. J. Savill, A. M. Ulatowski, M. D. Farrar, M. B. Johnston, H. J. Snaith and L. M. Herz, Adv. Funct. Mater., 2020, 30, 2005594 CrossRef CAS.
- J. Pascual, M. Flatken, R. Félix, G. Li, S.-H. Turren-Cruz, M. H. Aldamasy, C. Hartmann, M. Li, D. Di Girolamo, G. Nasti, E. Hüsam, R. G. Wilks, A. Dallmann, M. Bär, A. Hoell and A. Abate, Angew. Chem., Int. Ed., 2021, 60, 21583–21591 CrossRef CAS PubMed.
- Q. Chen, J. Luo, R. He, H. Lai, S. Ren, Y. Jiang, Z. Wan, W. Wang, X. Hao, Y. Wang, J. Zhang, I. Constantinou, C. Wang, L. Wu, F. Fu and D. Zhao, Adv. Energy Mater., 2021, 11, 2101045 CrossRef CAS.
- H. Lee, S. Gaiaschi, P. Chapon, D. Tondelier, J.-E. Bourée, Y. Bonnassieux, V. Derycke and B. Geffroy, J. Phys. Chem. C, 2019, 123, 17728–17734 CrossRef CAS.
- O. Almora, C. Aranda, E. Mas-Marzá and G. Garcia-Belmonte, Appl. Phys. Lett., 2016, 109, 173903 CrossRef.
- S. Ravishankar, Z. Liu, U. Rau and T. Kirchartz, PRX Energy, 2022, 1, 013003 CrossRef.
- J. Tong, Q. Jiang, A. J. Ferguson, A. F. Palmstrom, X. Wang, J. Hao, S. P. Dunfield, A. E. Louks, S. P. Harvey, C. Li, H. Lu, R. M. France, S. A. Johnson, F. Zhang, M. Yang, J. F. Geisz, M. D. McGehee, M. C. Beard, Y. Yan, D. Kuciauskas, J. J. Berry and K. Zhu, Nat. Energy, 2022, 642–651 CrossRef CAS.
- J. Diekmann, F. Peña-Camargo, N. Tokmoldin, J. Thiesbrummel, J. Warby, E. Gutierrez-Partida, S. Shah, D. Neher and M. Stolterfoht, J. Phys. Chem. Lett., 2023, 14, 4200–4210 CrossRef CAS.
- Y. Zhao, I. Yavuz, M. Wang, M. H. Weber, M. Xu, J.-H. Lee, S. Tan, T. Huang, D. Meng, R. Wang, J. Xue, S.-J. Lee, S.-H. Bae, A. Zhang, S.-G. Choi, Y. Yin, J. Liu, T.-H. Han, Y. Shi, H. Ma, W. Yang, Q. Xing, Y. Zhou, P. Shi, S. Wang, E. Zhang, J. Bian, X. Pan, N.-G. Park, J.-W. Lee and Y. Yang, Nat. Mater., 2022, 21, 1396–1402 CrossRef CAS.
- M. Y. Toriyama, M. K. Brod and G. J. Snyder, ChemNanoMat, 2022, 8, e202200222 CrossRef CAS.
- C. Li, A. Guerrero, S. Huettner and J. Bisquert, Nat. Commun., 2018, 9, 5113 CrossRef PubMed.
- K. P. Marshall, M. Walker, R. I. Walton and R. A. Hatton, Nat. Energy, 2016, 1, 16178 CrossRef CAS.
- U. Erdil, M. Khenkin, M. Remec, Q. Emery, V. Sudhakar, R. Schlatmann, A. Abate, E. A. Katz and C. Ulbrich, ACS Energy Lett., 2025, 10, 1529–1537 CrossRef PubMed.
- M. A. Torre Cachafeiro, E. L. Comi, S. Parayil Shaji, S. Narbey, S. Jenatsch, E. Knapp and W. Tress, Adv. Energy Mater., 2024, 15, 2403850 CrossRef.
- D. Di Girolamo, N. Phung, F. U. Kosasih, F. Di Giacomo, F. Matteocci, J. A. Smith, M. A. Flatken, H. Köbler, S. H. Turren Cruz, A. Mattoni, L. Cinà, B. Rech, A. Latini, G. Divitini, C. Ducati, A. Di Carlo, D. Dini and A. Abate, Adv. Energy Mater., 2020, 10, 2000310 CrossRef CAS.
- L. Huerta Hernandez, M. A. Haque, A. Sharma, L. Lanzetta, J. Bertrandie, A. Yazmaciyan, J. Troughton and D. Baran, Sustainable Energy Fuels, 2022, 6, 4605–4613 CAS.
- T. Zhu, D. Zheng, M.-N. Rager and T. Pauporté, Sol. RRL, 2020, 4, 2000348 CrossRef CAS.
- G. Kresse and J. Hafner, Phys. Rev. B:Condens. Matter Mater. Phys., 1993, 47, 558–561 CAS.
- G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15–50 CrossRef CAS.
- J. Klimeš, D. R. Bowler and A. Michaelides, Phys. Rev. B:Condens. Matter Mater. Phys., 2011, 83, 195131 Search PubMed.
- H. Peng, Z.-H. Yang, J. P. Perdew and J. Sun, Phys. Rev. X, 2016, 6, 041005 Search PubMed.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.