Abstract
Free full text
Vertically oriented low-dimensional perovskites for high-efficiency wide band gap perovskite solar cells
Abstract
Controlling crystal growth alignment in low-dimensional perovskites (LDPs) for solar cells has been a persistent challenge, especially for low-n LDPs (n<3, n is the number of octahedral sheets) with wide band gaps (>1.7eV) impeding charge flow. Here we overcome such transport limits by inducing vertical crystal growth through the addition of chlorine to the precursor solution. In contrast to 3D halide perovskites (APbX3), we find that Cl substitutes I in the equatorial position of the unit cell, inducing a vertical strain in the perovskite octahedra, and is critical for initiating vertical growth. Atomistic modelling demonstrates the thermodynamic stability and miscibility of Cl/I structures indicating the preferential arrangement for Cl-incorporation at I-sites. Vertical alignment persists at the solar cell level, giving rise to a record 9.4% power conversion efficiency with a 1.4V open circuit voltage, the highest reported for a 2eV wide band gap device. This study demonstrates an atomic-level understanding of crystal tunability in low-n LDPs and unlocks new device possibilities for smart solar facades and indoor energy generation.
Introduction
Low dimensional perovskites (LDPs), which have been known since the 1990s for transistor applications, represent an interesting family of materials for exploring light-matter interactions and quantum-confined physics and related applications. These materials consist of self-assembled bulky crystals composed of alternating inorganic layers spaced by large organic cations (R), arranged into a R2An–1BnX3n+1 structure, where n is the number of inorganic layers stacked together defining the so-called LDP dimensionality1. Typically, for A = organic cation, B = lead and X = iodine, if n<3, they are referred to as low-n LDPs, which possess unique optical features such as tuneable band gaps from 2.6eV to 1.7eV going from n=1 to n=3, and remarkably stable structures and phase purity2. Conversely, high-n LDPs have an absorption edge down to 1.5eV and are often constituted by a mixture of phases where APbX3 co-exists2. With their narrower band gap, high-n LDPs have been successfully integrated into efficient perovskite solar cells (PSCs)3. Indeed, from a structural perspective, high-n LDPs partially lose their layered structure and are able to efficiently transport the photoinduced charges, assisted by the co-presence of APbX3 phases (as revealed by the edge of external quantum efficiency spectra close to 1.5eV)4. Moreover, induced vertical growth of high-n LDP crystal structures has been demonstrated, further boosting devices’ power conversion efficiency (PCE)5–13. However, this does not hold true for low-n LDPs where the inorganic frame usually grows parallel to the substrate14 or self-organizes in a random orientation15. Consequently, vertical charge transport and extraction are hindered, strongly limiting the PCE (typically <2%) when integrated into a typical planar device stack16,17. This imposes a severe limit to low-n LDP applications, despite their wide band gap being ideal for semi-transparent devices and building integrated photovoltaics (PVs)18,19. By solving these limitations, low-n LDPs would have advantages over the conventional high band gap solutions based on mixed halide (Br/I) APbX3 perovskites such as methylammonium lead halides (MAPbX3)20, which still suffer from structural instability and light-induced halide segregation, causing open-circuit voltage (VOC) losses and performance reduction21–24. Therefore, to go beyond the state-of-the-art, it is imperative to develop a method to control the tilting and alignment of the inorganic framework of the low-n LDPs with respect to the substrate15, which would be a major advance for wide band gap PSCs.
In this work, we use thiophenemethylammonium iodide (TMAI) as the R cation to form a vertically oriented n=2 LDP with a band gap of ~2eV, which we successfully integrate into a planar device stack. We then demonstrate a highly efficient 2eV wide band gap vertically oriented LDP solar cell, overcoming the current transport limits. This has been obtained by elucidating the critical role of Cl in dictating vertical crystallization. In contrast to the current knowledge on 3D MAPbX3 perovskites, we found that a small degree of Cl substitutes for I in the PbI6 octahedra leading to an expansion of the unit cell in one direction, which is critical in inducing vertical alignment as corroborated by joint experimental and ab initio modelling analysis. Integrated into a solar cell, the vertical orientation is preserved through the whole device thickness, ensuring efficient charge transport and resulting in a champion 9.4% PCE, which is, to the best of our knowledge, the highest PCE reported for a PSC based on a low-n 2eV band gap LDP. The concept is also validated on a semi-transparent solar cell, demonstrating the potential exploitation of low-n LDPs in coloured and transparent PVs.
Results
Thin film fabrication and GIWAXS analysis
Figure 1 schematically represents the preparation procedures for low-n LDPs with the chemical formula TMA2MAPb2I7 (n=2). The standard method (Fig. 1a) involves the dissolution of the stoichiometric ratio of precursor powders in dimethylformamide (DMF) and the subsequent spin coating and annealing steps. The structure and texture of the resulting thin films were investigated by grazing incident wide-angle X-ray scattering (GIWAXS), whose results are reported in Fig. 1b. The unit cell can be identified as being very close to orthorhombic with axes 8.8×41.5×8.7 Å and with β=87°, in close agreement with previous reports25. Since the (0k0) reflections are predominantly oriented along the qz direction, the crystallites orient with their (0k0) planes parallel to the substrate, meaning the inorganic slabs are oriented horizontally (Fig. 1c). This is the most common orientation reported for low-n LDPs (named horizontal hereafter)14.
In Fig. 1d the developed method is outlined, in which we employ a partial substitution of methylammonium iodide (MAI) with methylammonium chloride (MACl) at different concentrations (see Supplementary Fig. 1); this was further optimized at a 30:70 ratio (MAI:MACl) and by adding an annealing step of the mixed precursor solution at 60°C for 2h, before its deposition. The MAI:MACl ratio corresponds to an optimal 10% total amount of Cl substituting I in the perovskite formulation. When 10% Cl is added, the GIWAXS pattern (Fig. 1e) drastically changes revealing that the (0k0) reflections are now oriented along the qy direction. This indicates a switch of crystallite orientation such that the inorganic slabs are now vertically oriented with respect to the substrate. Notably, the signals appear more arch-like because of a certain degree of structural disorder and misalignment with respect to the pure TMA2MAPb2I7 phase. As reported in Fig. 1f, two main orientations of the unit cell are detected for the films produced with our formulation: crystallites are now oriented either with the (001) or with the (101) planes parallel to the substrates. In both orientations, the inorganic slabs are vertically aligned (sample named vertical hereafter), as depicted in Fig. 1f.
To ultimately confirm the different crystal structure orientation of horizontal and vertical samples, we perform a texture analysis through advanced X-Ray Diffraction (XRD) Pole Figure (PF) investigation, whose results are reported in Fig. 1g. By comparing the experimental PFs with the Wulff Stereographic Projections (WSPs) for a single crystal perfectly aligned along the [101] (in red) and [010] (in yellow) crystallographic directions, we investigate the orientation of three different lattice planes: (202) (Fig. 1g), (151) and (080) (Supplementary Fig. 2) that correspond to reflections at 28.73°, 17.83°, and 17.11°, respectively. In the case of the horizontal LDP, the PF (202) presents higher intensity in the outer ring coinciding with the diffraction of the [101] family of planes. These are responsible for the central peak of PF (080), confirming that in the horizontal sample, the vertical orientation of the inorganic layer is not present, in agreement with the GIWAXS results. Conversely, the strong peak at the centre of PF (202) for the vertical sample confirms such preferential growth along the [101] direction in which crystals tend to grow with a fibre-textured disposition perpendicularly to the substrate, which is also confirmed by the ring observed in the PF (151) corresponding to the tilted plane reflections. Overall, the structural analysis shows the effectiveness of our strategy in directing the vertical growth on the low-n LDP thin films.
Optical and photoconductivity analysis of LDP thin films
The optical properties of the low-n LDP thin films do not depend on the material crystallization. Indeed, the different samples absorption spectra are identical, displaying a sharp onset at 2.1eV (as calculated from the Tauc plot in Supplementary Fig. 3) and a clear excitonic peak at 567nm (Fig. 2a), typical of a n=2 LDP. We note that the horizontal sample shows a minor shoulder at longer wavelengths, indicative of a small fraction of n=3 phase, which also appears in the steady state photoluminescence (PL) spectra together with the peak related to the n=2 phase (see Supplementary Fig. 4). To gain a demonstration of the material stability, we monitor the PL peak position exposing the films under continuous 1 Sun AM1.5G illumination and thermal stress at 60°C (Supplementary Fig. 5). In both cases no variations are observed, indicative of a high robustness and phase integrity against UV light and thermal stressors.
To investigate the effect of vertical alignment on the charge-carrier transport properties, we start by analysing the electronic structure of TMA2MAPb2I7 (shown in Fig. 2b), obtained using density functional theory (DFT) calculations. The primitive cell of TMA2MAPb2I7, along with the Brillouin zone showing the high symmetry directions, are shown in Supplementary Fig. 6. TMA2MAPb2I7 has a direct band gap at the Γ-point. The flat bands along the high symmetry direction Γ-Y indicate heavy carriers perpendicular to the PbI6 plane, while the more dispersive bands along Γ- Δ0, and Γ-Z directions suggest lighter carrier transport within the plane. This is corroborated by the effective mass values (inset Fig. 2b) along the three axes of the conventional unit-cell, which are higher along the vertical b-axis and moderate along the horizontal a- and c-axes for both electron and hole carriers. The calculated effective masses are in good agreement with previous reported values for LDPs26. These results demonstrate that carrier transport can be significantly enhanced by growing these LDPs with vertical alignment.
In order to further assess how the different orientations of the LDP planes impact charge-carrier transport at the nanoscale level, we employ optical-pump THz-probe photoconductivity spectroscopy. A sub-ps laser pulse at 3.1eV excites the LDP films, and the resulting photoconductivity is recorded for a range of fluences, as shown in Fig. 2c and d for samples with horizontal and vertical layer orientation with respect to the substrate, respectively. We may further extract the effective electron-hole sum mobility from the conductivity recorded immediately after photoexcitation, under knowledge of the absorbed photon density (see methods section). We note that since the electric field of the THz probe pulse oscillates at right angle with respect to its propagation through the substrate, such mobility values reflect charge-carrier motion within the thin-film plane, parallel to the substrate. Figure 2e and f show that these films indeed exhibit the expected behaviour for highly oriented LDP films. A high in-plane mobility of 4.9 cm2V−1s −1 is observed when the THz electric field oscillates fully along the LDP planes, for the horizontal samples. For the vertical sample, on the other hand, the E-field vector lies on average half along the LDP planes, and half in the direction vertical to the planes. The observed mobility of 2.2 cm2 V−1 s −1 is indeed lowered by nearly a factor of 2, as would be expected given the insignificant mobility of charges across spacer layers (in line with the effective mass values in Fig. 2b), confirming the highly oriented nature of the LDP films27,28.
Insights into chlorine effects on vertical crystallisation
We now focus our attention on the fundamental question: why does the addition of Cl have such a strong impact on guiding the vertical crystallization? For standard MAPbX3 perovskites, it is well known that MACl-assisted perovskite crystallization favours the α-phase (with respect to non-perovskite phases, as in the case of formamidinium-based perovskites)29,30, and that Cl sublimates during the process31. In order to verify the presence of Cl and map its content, we performed time-of-flight secondary ion mass spectroscopy (ToF-SIMS), as shown in Fig. 3a and b for the horizontal and vertical samples, respectively. For the horizontal sample Cl is absent, comparable in intensity to the background signal. In contrast, in the vertical sample there is a gradient distribution of Cl which maximizes at the indium tin oxide (ITO)/perovskite interface. In addition, for both horizontal and vertical LDPs, the CN, I and S signals are uniformly distributed across the whole film thickness.
To further deepen our understanding on Cl inclusion in the film, we conducted a combined experimental and computational modelling analysis, whose results are reported in Fig. 3c–f. First, by closely comparing the GIWAXS profiles of the two samples, a shift in the (0k0) peaks is observed (Fig. 3c). The pattern of the vertical sample can be indexed by a slightly larger orthorhombic unit cell, especially along the axis perpendicular to the octahedra, with axes 8.5 × 42.5 × 9.0 Å (with respect to 8.8 × 41.5 × 8.8 Å for the horizontal sample). The expansion of the long axes can be the result of Cl incorporation inside the crystal unit, leading to structural distortion and lattice expansion. While Cl inclusion in the unit cell is known to be energetically unfavourable in MAPbX3 perovskites, the situation appears to be different for low-n LDPs, due to their layered, quantum-confined structure. Cl partially substitutes I in the equatorial position, creating a strain in the octahedra and thus elongating the unit cell along the vertical axes. Such vertical compression induces a switch in crystal orientation, responsible for the vertical alignment. A similar structural effect, where Cl incorporates in the equatorial position substituting I, has been recently demonstrated in LDP quantum confined nanocrystals, strongly corroborating our results32.
To verify our hypothesis and gain additional atomic-scale insights into Cl incorporation in TMA2MAPb2I7, we have used DFT-based methods to investigate the thermodynamic stability and miscibility of Cl-mixing at I-sites. The positive mixing energies for all the structures between 0 < x<0.21 and 0.79 < x<1 in I1–xClx (Supplementary Fig. 7) indicate a low tendency for Cl/I mixing, although the structures near the end points (x=0, 1) can be accessible at finite temperatures. The thermodynamic stability phase diagram is shown in Fig. 3d. Three important features emerge. First, a stability region exists close to the I-rich (x=0) and Cl-rich (x=1) regions over 300K, which grows with an increase in temperature. This indicates that Cl can be incorporated into I-sites in small amounts (~2% at 373K). Second, small metastable regions exist within 10% of the I-rich and Cl-rich regions (x<0.1 and x>0.9) for the entire temperature range. Being thermodynamically metastable, these states may be accessed and made kinetically stable via alternate synthesis methods. Third, the I/Cl-mixing states are unstable in the wide range of 0.1 < x<0.9. Overall, these results show that it is possible to incorporate Cl at I-sites, but only in small amounts.
Next, we examined the b-lattice parameter variation of all the structures between 0 < x<0.21 and 0.79 < x<1 in I1–xClx to investigate the structural distortion upon Cl addition (Fig. 3e; further analysis of the structural and energetic variation with Cl content is shown in Supplementary Figs. 8,9). Interestingly, compared to the structures with the minimum b-lattice parameter at each concentration, the structures with the maximum b-parameter (i) are energetically favoured, (ii) have a preferred equatorial distribution of Cl at I-sites (Fig. 3f), and (iii) have larger spacer cation distances. Examining the energetically most stable structures at each concentration, Cl ions are found to be predominantly distributed at the equatorial I-sites. The structures with the highest b-lattice parameter have lower unit cell volumes due to compression along the a- and c-lattice parameters. Analysis of the N-C-C and C-S-C bond angles of the spacer cations from 0 <x<0.21 indicate that there is no major stretching of the spacer cation ring, whereas the terminal N-C-C bonds are affected by the presence of highly electronegative Cl- ions in close vicinity. In summary, the equatorial distribution of Cl is energetically preferred, which causes compressive strain on the unit cell along the a- and c-lattice parameters. This creates steric hindrance among the spacer cations, leading to an elongation of the b-parameter, which is in line with our GIWAXS results.
Finally, by testing the effects on crystal orientation of other Cl- sources such as thiophenemethylammonium chloride (TMACl) and lead chloride (PbCl2) (Supplementary Fig. 10), we could ultimately confirm the role of the anion on triggering verticalization.
Device characteristics and performance
We integrate our vertical low-n thin film into a p-i-n PSC, with the key aim of demonstrating the proof-of-concept of an efficient wide band gap LDP solar cell and ultimately breaking current limits. First, to verify that the vertical orientation is preserved when the perovskite is sandwiched in a device stack composed of a self-assembled monolayer (SAM)-modified oxide and a fullerene derivative electron acceptor, we performed high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM), whose results are reported in Fig. 4a. From the cross-section images, all the different layers comprising the device stack are visible (elemental mapping is shown in Supplementary Fig. 11). The grains of the vertical LDP layer are arranged in a columnar fashion, perpendicularly to the substrate, with a length of the nanopillars of the order of 300nm covering the entire thickness of the active layer. By increasing the resolution through the high-resolution TEM (HR-TEM) technique, we could further investigate the alignment of the inorganic backbone and assess the purity of the low-n LDP phase. As shown by the central panel of Fig. 4a, the vertical features can be ascribed to the layered structure of the LDP, in which the bright areas represent the organic sheets (with a spacing of about 13.06Å) and the dark lines corresponds to the stack of two PbI6- octahedra layers. Furthermore, the right image of Fig. 4a shows HR-TEM and corresponding Fast Fourier Transform (FFT) along zone axis <001>, which confirms that the lattice parameters of the material correspond to those of a pure n=2 LDP phase (a = b=7.36Å) (no higher n phase are detected)33.
A schematic of the device is presented in Fig. 4b. Figure 4c shows the characteristic current density (J)-voltage (V) curve for champion PSCs fabricated with horizontal and vertical LDPs in both forward and reverse scan. The horizontal PSC displays a PCE of 0.8%, strongly limited by the short-circuit current density (JSC), as commonly observed in the literature, due to the charge transport barrier34. In contrast, the vertical device shows a PCE of 9.4%, strongly outperforming its horizontal counterpart, with a JSC of 10.6 mAcm−2, a fill factor (FF) of 63.2%, and a VOC of 1.40V. To the best of our knowledge, this proof-of-concept device shows the highest PV parameters reported thus far for a 2eV band gap LDP (i.e. composed of a low-n LDP, see Supplementary Table 1 and Supplementary Fig. 12 for comparison with literature results). A statistical analysis by box charts for the PV parameters of horizontal and vertical PSCs is reported in Supplementary Fig. 13.
Figure 4d shows the external quantum efficiency (EQE) and integrated current density (Jint) curves for horizontal and vertical devices. At lower wavelengths, the EQE signal is significantly different for the two samples. The vertical device achieves values higher than 80%, whereas the horizontal barely overcomes 25% over the visible range. As the absorption spectra of the two samples are very similar, the reason behind the higher photocurrent is ascribed to the different crystal orientation which provides a route for efficient charge transport. Moreover, the EQE spectrum, and specifically its sharp onset at 2eV, provides strong evidence of the phase purity of our low-n LDP active layer, setting the demonstration for a wide band gap low-n LDP to generate such a high photocurrent. Furthermore, we tested the stability of both vertical and horizontal devices by maximum power point tracking (MPPT) under 1 Sun AM1.5G illumination. As shown in Supplementary Fig. 14, stability measurements reported in the literature show that devices fabricated with 2eV band gap, 3D mixed halide perovskites reach the 80% of their initial PCE in a few minutes (T80=400s and 1h for refs. 35 and36, respectively). In contrast, T80 for our vertical device is 150h, demonstrating that the vertical orientation of the crystal structure is beneficial to stability.
Finally, to broaden the scope of our concept, we tested our vertical low-n LDP active layer in semi-transparent devices (Fig. 4e), by substituting the opaque Ag electrode with an indium zinc oxide (IZO) transparent electrode (Fig. 4f, inset). The light transmission analysis, represented in Fig. 4f, shows the high transmittance of the device that reaches values higher than 80% in the region > 620nm, resulting in a calculated average visible transmittance (AVT) for the device of 31%, which meets the requirements for window-integrated PV applications. As presented in Fig. 4g and h and in Supplementary Fig. 15, devices exhibit average values of 1.26V for VOC, 7.3 mAcm−2 for JSC, 50% for FF and 4.6% for PCE, resulting in a light utilization factor (LUE) of 1.86%, which is similar to 3D perovskites-based semi-transparent devices previously reported37,38.
Discussion
In conclusion, our work demonstrates that low-n LDPs can effectively represent a breakthrough strategy for wide-band gap solar cell applications. This approach is made possible by achieving a phase pure low-n LDP thin film featuring vertical crystallization of its inorganic backbone, which allows efficient charge transport. The orientational switch, as demonstrated here by a combined experimental and modelling study, is triggered by Cl incorporation in the unit cell structure which substitutes for I in the equatorial position. As a result, the unit cell is compressed featuring a vertical strain which induces the crystal orientation. In contrast to 3D ABX3 perovskites including MAPbI3, such substitution is thermodynamically favoured in LDPs possibly due to their intimate layered nature, representing a fundamental advance in atomistic understanding. When integrated into a working solar cell the verticalization is preserved, demonstrating a highly efficient low-n LDP solar cell with a 2eV band gap. This concept represents a viable alternative for visible light harvesting wide band gap applications, ranging from building integrated PVs to indoor light harvesting and multi-junction PV devices.
Methods
Materials
Solvents were purchased from Sigma-Aldrich (highest purity and anhydrous); lead iodide (PbI2, 99.99% purity) and (2-(3,6-Dimethoxy-9H-carbazol-9-yl)ethyl)phosphonic acid (MeO-2PACz) were purchased from TCI; organic iodide salts (MAI, MACl, TMAI) were purchased from GreatCell Solar. (6,6)-Phenyl C61 butyric acid methyl ester (PCBM, >99.5%) was purchased from Lumtec; Bathocuproine (BCP, sublimed grade, 99.99% purity) was purchased from Sigma Aldrich. Unless otherwise specified, materials and solvents have been used as received. All the solutions/dispersions were prepared into an Ar-filled glovebox and the spin coating steps were carried out into a N2-filled glovebox (in both case, O2 concentration <0.1 ppm and H2O concentration <0.1 ppm).
Low-dimensional perovskite thin film and solar cell fabrication
The precursor solution (0.7M) for the horizontal TMA2MAPb2I7 (n=2) LDP was prepared by mixing stoichiometric ratio of TMAI, MAI and PbI2 powders in DMF. In the case of the vertically oriented LDP, 70mol% of MAI was replaced by MACl, without changing the total stoichiometric amount of MA in the solution. After complete dissolution, the solution was subjected to a thermal annealing step of 2h at 60°C. The solution was deposited onto glass substrates via single-step spin coating procedure at 4000rpm, 2000 rpm s−1 acceleration for 40s. 120µl of toluene was dropped onto the spinning substrate at 20s before the end of the programme for antisolvent procedure. Subsequently, samples were annealed at 100°C for 30min.
For PSCs fabrication, prepatterned ITO/glass substrates were sequentially cleaned with acetone and 2-propanol (IPA) by ultrasonication for 15min in each solvent. ITO/glass substrates were then dried with N2 and treated with oxygen plasma at 100mW for 10min. Then, substrates were moved into a N2-filled glovebox for subsequent layers deposition. MeO-2PACz self-assembly monolayer was deposited by spin coating (3000rpm, 30s, 1500rpms−1) a 0.335mgmL−1 solution in ethanol, followed by annealing at 100°C for 10min. The LDP layer was fabricated following the same procedure employed to prepare thin films. A PCBM solution (20mgmL−1 in chlorobenzene) was deposited onto the perovskite by dynamic spin coating at 2000 rpm, with an acceleration of 1000rpms−1 for 30s, followed by 10min annealing at 100°C. Finally, 20nm of SnO2 were deposited by atomic layer deposition (ALD, Picosun) and 80nm of Ag were thermally evaporated, producing pixels with 0.045 cm2 active area (shadow mask aperture area estimated by optical microscopy).
For semi-transparent devices, 100nm of IZO was sputtered (Angstrom EvoVac) onto the SnO2 buffer layer as top contact electrode (pixels active area ~0.1 cm2).
Grazing incident wide-angle X-ray scattering
GIWAXS experiments have been performed at the MINA instrument at the University of Groningen. The instrument is built around a Cu rotating anode source, delivering a high-power X-ray beam with a photon energy of 8keV (
Pole figure
Pole Figure (PF) technique is based on the acquisition of XRD data by varying the incidence angle along two different axes to build PF maps. In such way, different orientations of crystalline samples are proved and compared with the Wulff Stereographic Projection (WSP) calculated from single crystal data. Pole Figure (PF) analysis was performed with the detector set in 0D mode, using 0.04rad soller slits and 0.28° parallel plate collimator (PPC) on the diffracted beam. This configuration allows a maximum sample tilt equal to χ=85°. Simulated pattern for the TMA2MAPb2I7 n=2 LDP perovskite was obtained with CrystalDiffract 6.9.4 from CrystakMaker Software Ltd. Pole Figure analysis was carried out using X’Pert Texture 1.3 software from PANalytical. The Wulff stereographic projections (WSP) were created using WinWulff 1.6 from JCrystalSoft.
UV-Vis absorption
The absorption spectra of the perovskite thin films were collected using a UV-NIR spectrophotometer from Perkin Elmer (Lambda 1050+).
Photoluminescence measurements
The photoluminescence (PL) spectra of the different samples were collected using a customized optical setup40 in which the excitation was provided by pulsed/CW laser at 470nm (PicoQuant), while the luminescence was analysed with an interferometer (GEMINI by Nireos) and recorded with a single photon detector (IDQuantique) coupled with a Time Tagger (Swabian Instrument).
Computational modelling
First-principles simulations were performed within the framework of density functional theory (DFT) as implemented in the Vienna Ab-initio Simulation Package (VASP)41–43 using a projector augmented-wave (PAW)44 basis. The generalized gradient approximation (GGA) exchange-correlation functional of Perdew Burke-Ernzerhof (PBE)45 was used for all the structures. VASP-recommended PAW pseudopotentials were used for all the atomic species (C:2s22p2, N:2s22p3, H:1s1, S:3s23p4, Pb:5d106s26p2, I:5s25p5, and Cl:3s23p5). Grimme’s DFT-D3 dispersion correction46 was considered due to the presence of organic molecules. A plane wave energy cut-off of 500eV and a Brillouin zone sampling using Γ-centred k-mesh are employed done REMOVE DONE for all the calculations. A k-point grid of 3 × 1 × 3 with a Gaussian smearing of 0.1 was used for structure optimization whereas a k-point grid of 6 × 2 × 6 with Gaussian smearing of 0.01 was used to obtain accurate charge densities to calculate the electronic structure. The cell volume, shape, and atomic positions were fully relaxed using the conjugate gradient algorithm until the force on each atom converged below 0.01eVÅ−1. The self-consistent energy convergence of 10−6eV was used for electronic degrees of freedom.
The carrier effective masses were calculated using the Effective Mass Calculator (EMC) code47. We used the Site-Occupation Disorder (SOD) code48 to generate symmetrically inequivalent structures in the orthorhombic space group Aba2 (#41) for 0 < x<0.21 and 0.79 < x<1 in I1–xClx. It is unfeasible to enumerate the thousands of symmetrically inequivalent structures for 0.21 < x<0.79 in I1–xClx. Therefore we used the icet code49 to generate randomly occupied structures in this compositional range. The thermodynamic stability phase diagram for Cl mixing at I-sites was calculated using the Generalized Quasi-Chemical Approximation (GQCA)50, which has been effectively utilized in the thermodynamic analysis of perovskite materials51–53. 235 lowest energy-optimized structures distributed uniformly over each possible concentration from 0 < x<1 in I1–xClx were used to generate the thermodynamic stability phase diagram. 1121 different structures chosen randomly and distributed uniformly between 0 < x<0.21 and 0.79 < x<1 in I1–xClx were optimized to analyse the preferred distribution of Cl and their effects on the lattice distortion.
THz photoconductivity and mobility measurements
An amplified laser system (Spectra Physics, MaiTai – Ascend – Spitfire), with a 5kHz repetition rate, centre wavelength of 800nm and pulse duration of 35fs is used to generate the THz radiation using a spintronic emitter. The THz probe is then focused onto the sample, overlaid with a 400nm excitation pump that is generated using a Beta Barium Borate (BBO) crystal. The THz radiation transmitted through the sample is then detected via free-space electro-optical sampling in a ZnTe (110) crystal of thickness 200 μm54.
The variations in the THz transmission ΔT T−1 through the sample can therefore be measured, and temporally resolved by varying the time delay between the optical pump and the probing THz pulses in order to obtain the temporal evolution of the photo-induced conductivity of the sample51,52.
Additionally, for low enough laser fluences and charge-carrier densities, the sum mobility of free charge carriers will have a linear dependence on photoconductivity values and can therefore lead to an extraction of the intrinsic effective sum mobility based on the early time conductivity values post photoexcitation55,56.
We also ascertained that the THz mobility values extracted within the first picoseconds indeed solely reflected the motion of charge carriers within the n=2 LDP phase, by also recording time-resolved photoluminescence spectra, using a time-correlated single photon counting (TCSPC) device following excitation by a 398nm picosecond pulsed diode laser (PicoHarp, LDH-D-C-405M). These PL measurements showed that over the first nanoseconds of charge-carrier lifetimes for which the THz photoconductivity transients were recorded, the PL solely reflected emission from an n=2 LDP phase.
Time-of-flight secondary ion mass spectroscopy
ToF-SIMS analyses were acquired with ToF-SIMS 5–100, with the pulsed primary ions from a Cs+ (2keV) ion gun for sputtering and a Bi+ ion beam (30keV) for analysis.
Transmission Electron Microscopy
Transmission electron microscopy (TEM)-based experiments were conducted in both parallel beam mode high-resolution TEM (HRTEM) and converge beam scanning TEM (STEM) mode. The HRTEM experiments were performed in Cs spherical aberration image-corrected Thermo Fisher Titan 60–300 Cubed TEM microscope. In STEM mode, High-Angle Annular Dark-Field STEM HAADF-STEM) imaging combined with Energy Dispersive Spectroscopy (STEM-EDS) was executed for elemental mapping on Cs-aberration-probe-corrected Thermo Fisher Titan 60–300 Cubed TEM microscope operating at 300kV. TEM data processing was performed using Gatan™ Digital Micrograph and Thermo Scientific™ Velox suites. For the TEM study, cross-sectional electron-transparent lamella was prepared in a focused ion beam (FIB) equipped scanning electron microscope (SEM-FIB Helios G4 DualBeam, FEI) with the help of an EasyLift nanomanipulator and Ga ion source. To protect the region of interest during FIB, two types of protective coatings were deposited: a 0.5µm layer of Pt coating deposited by the e-beam, followed by a 3µm layer of Pt coating deposited by the ion beam for final protection. Step-by-step ion beam milling procedure with beam current (2.4, 0.44, 0.26, 0.045, 0.025nA for 30 to 5kV) was carried out by cutting and thinning down lamella to 60nm while decreasing beam current (till 0.025nA at 2kV) to avoid ion beam damage. Additionally, a low current cleaning process (5–2kV, 81-28 pA) was performed to remove any potential contamination.
Photovoltaic device characterization
Current density (J)-voltage (V) measurements were performed in ambient conditions under simulated AM1.5G light with an intensity of 100mWcm−2 (Wavelabs-Sinus 70). Intensity was calibrated using a calibrated Si reference cell (KG-3 filter Centronics LCE-50). Cells were scanned using a Keithley 2450 source-metre backward and forward from 1.5V to −0.1V, with a 100 mVs−1 scan speed (no dwell time or preconditioning applied). A shadow mask with an aperture of 0.03 cm2 (measured through optical microscopy) was employed during the measurement. Operational stability tests were carried out by maximum power point tracking (MPPT) under AM1.5G illumination (100mWcm−2) in N2 atmosphere at room temperature. External quantum efficiency (EQE) measurements were performed with a Cicci Research s.r.l ARKEO steady-state tests module. Wavelength scan range was set between 300 and 900nm with a scan step of 10nm.
Average visible transmittance and light utilization efficiency computation
AVT has been calculated using Eq. 1 through the widely reported approach57:
where λ is the wavelength, T is the device light transmission, P is the photopic response function, and S is the AM1.5G solar irradiation.
LUE has been calculated as the product of AVT value and device power conversion efficiency (PCE), as presented in Eq. 2:
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Supplementary information
Acknowledgements
The authors thank the University of Pavia through the programme “Dipartimenti di Eccellenza 2023–2027” and Fondazione Cariplo Economia Circolare 2021 Project (FLHYPER, no 20201067), funded under the “Circular Economy-2020” call. G.G. acknowledges the “HY-NANO” project that received funding from the European Research Council (ERC) Starting Grant 2018 under the European Union’s Horizon 2020 research and innovation programme (Grant Agreement no. 802862) and the GOPV project (CSEAA_00011) which received funds from Bando Ricerca di Sistema - CSEA - TIPO A Piano triennale 2019 −2021 Decreto direttoriale 27 ottobre 2021 del Ministero della Transizione Ecologica; MASE-(ex MITE). The authors acknowledge the bilateral project CRG2022 - HERO2D/3D Perovskite Heterojunction Solar Cells: Pairing Performance with Stability. The authors acknowledge funding by the Engineering and Physical Sciences Research Council (EPSRC); U.K. M.S.I. and V. acknowledge the Engineering and Physical Sciences Research Council (EPSRC) Programme Grant (EP/X038777/1) for a postdoctoral fellowship and are grateful to the UK’s HEC Materials Chemistry Consortium (EP/X035859/1) for the use of the ARCHER2 high-performance computing facilities. The authors acknowledge the use of the KAUST Core Labs facilities for some parts of device fabrication and TEM measurements. The authors are grateful for the technical support in TOF-SIMS measurement from Nano-X vacuum interconnected nanotech workstation (Nano-X) of Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences (SINANO). L.M.H acknowledged financial support by the EPSRC. G.G.acknowledges Dr. Stefano Toso (IIT, Italy) for useful discussion.
Author contributions
Conceptualization: A.Z., V.L., V., S.D.W, M.S.I., G.G. Investigation: A.Z., V.L., V., F.T., B.V, K.A.E., J.D, A.S., F.F., G.P., M.P., S.M., C.D., B.K.Y., M. B., E.U., E.A. Resources: V.P. Data curation: S.M. Writing – original draft: A.Z., V.L., V., K.A.E., G.P. Writing – review & editing: M. B., E.U., E.A., C.Q.M., M.A.L., L.M.H., S.D.W., M.S.I., G.G. Visualization: V.L., V. Supervision: F.D., M.A.L., M.D.B., L.M.H., G.P., S.D.W., M.S.I., G.G.
Peer review
Peer review information
Nature Communications thanks the anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Data availability
The data generated in this study are available upon request from the corresponding author.
Competing interests
A patent on the fabrication methodology of low-dimensional perovskites has been filed by the University of Pavia, with G.G. and A.Z. as inventors. The application number is 102022000025854. The patent is currently under examination. The other Authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
These authors contributed equally: Andrea Zanetta, Valentina Larini.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-53339-6.
References
Articles from Nature Communications are provided here courtesy of Nature Publishing Group
Citations & impact
This article has not been cited yet.
Impact metrics
Alternative metrics
Discover the attention surrounding your research
https://www.altmetric.com/details/169551597
Similar Articles
To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.
Double Perovskite Interlayer Stabilized Highly Efficient Perovskite Solar Cells.
ACS Appl Mater Interfaces, 16(34):44988-44996, 19 Aug 2024
Cited by: 0 articles | PMID: 39160138
Ethylenediammonium-Based "Hollow" Pb/Sn Perovskites with Ideal Band Gap Yield Solar Cells with Higher Efficiency and Stability.
J Am Chem Soc, 141(21):8627-8637, 14 May 2019
Cited by: 8 articles | PMID: 31063361
Band Gap Tuning via Lattice Contraction and Octahedral Tilting in Perovskite Materials for Photovoltaics.
J Am Chem Soc, 139(32):11117-11124, 04 Aug 2017
Cited by: 58 articles | PMID: 28704048
Nanostructured Perovskite Solar Cells.
Nanomaterials (Basel), 9(10):E1481, 18 Oct 2019
Cited by: 5 articles | PMID: 31635204 | PMCID: PMC6835749
Review Free full text in Europe PMC
Funding
Funders who supported this work.
European Research Council (1)
Grant ID: 802862
Fondazione Cariplo (Cariplo Foundation) (1)
Grant ID: 20201067