Timothy W.
Jones‡
a,
Anna
Osherov‡
b,
Mejd
Alsari‡
c,
Melany
Sponseller
b,
Benjamin C.
Duck
a,
Young-Kwang
Jung
d,
Charles
Settens
b,
Farnaz
Niroui
b,
Roberto
Brenes
b,
Camelia V.
Stan
e,
Yao
Li
ef,
Mojtaba
Abdi-Jalebi
c,
Nobumichi
Tamura
e,
J. Emyr
Macdonald
g,
Manfred
Burghammer
h,
Richard H.
Friend
c,
Vladimir
Bulović
b,
Aron
Walsh
di,
Gregory J.
Wilson
a,
Samuele
Lilliu
j and
Samuel D.
Stranks
*bc
aCSIRO Energy Centre, Mayfield West, NSW 2304, Australia
bResearch Laboratory of Electronics, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, Massachusetts 02139, USA. E-mail: sds65@cam.ac.uk
cCavendish Laboratory, University of Cambridge, JJ Thompson Avenue, Cambridge CB3 0HE, UK
dDepartment of Materials Science and Engineering, Yonsei University, Seoul 03722, Korea
eAdvanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA
fXi’an Jiaotong University, State Key Laboratory for Mechanical Behavior of Materials, Xi’an, China
gSchool of Physics and Astronomy, Cardiff University, Cardiff CF24 3AA, UK
hEuropean Synchrotron Radiation Facility, Grenoble, France
iDepartment of Materials, Imperial College London, Exhibition Road, London SW7 2AZ, UK
jDepartment of Physics and Astronomy, University of Sheffield, Sheffield S3 7RH, UK
First published on 15th January 2019
Halide perovskites are promising semiconductors for inexpensive, high-performance optoelectronics. Despite a remarkable defect tolerance compared to conventional semiconductors, perovskite thin films still show substantial microscale heterogeneity in key properties such as luminescence efficiency and device performance. However, the origin of the variations remains a topic of debate, and a precise understanding is critical to the rational design of defect management strategies. Through a multi-scale investigation – combining correlative synchrotron scanning X-ray diffraction and time-resolved photoluminescence measurements on the same scan area – we reveal that lattice strain is directly associated with enhanced defect concentrations and non-radiative recombination. The strain patterns have a complex heterogeneity across multiple length scales. We propose that strain arises during the film growth and crystallization and provides a driving force for defect formation. Our work sheds new light on the presence and influence of structural defects in halide perovskites, revealing new pathways to manage defects and eliminate losses.
Broader contextMetal halide perovskites are generating enormous excitement for optoelectronic devices including solar cells and light-emitting diodes. However, there are still substantial numbers of defects that trap energized charge carriers, leading to power losses in devices. The origin of these defects, particularly on the micro-scale, remains unknown. In this work, the authors show that these defects are related to complicated strain patterns that appear on multiple length scales in perovskite films – ranging from tens of micrometres down to the tens of nanometre scale. The results suggest that these strain patterns are built into the film upon formation. This work has profound implications on our understanding of the operation of the materials on the micro-scale. It also highlights the need to control strain as a lever to eliminate these problematic defects in order to ultimately attain maximum device performances. |
Macroscopic observations of non-radiative losses in perovskite films are manifested as an emission heterogeneity on the microscale, where the PL lifetime and intensity have been shown to vary between different grains even in the highest-performing polycrystalline films.13 This microscale spatial heterogeneity has also been observed in charge carrier diffusion lengths14 and in the photocurrent and open-circuit voltage in full solar cell devices.15 Together, these measurements demonstrate the impact of the microscale on the resulting macroscopic polycrystalline film properties, suggesting that there is enormous scope for improving device PCE by understanding and then addressing these heterogeneous optoelectronic properties. The literature consensus attributes non-radiative losses to a spatially-heterogeneous population of sub-bandgap electronic states that act as trap-mediated recombination sites.13,16 However, the origin of these traps and their heterogeneity is still unresolved.
Structural defects have a strong influence on the charge-carrier lifetime and recombination in semiconductors and, typically, charged point defects make efficient non-radiative recombination centers.17 Most semiconductors are intolerant to the presence of defects, with typical defect densities in crystalline silicon devices required to be no larger than ∼108–1010 cm−3.18 In contrast, trap densities in polycrystalline perovskite films in high-performing devices have been determined to be relatively large (∼1013–1018 cm−3).16,19,20 This in fact suggests a remarkable defect tolerance in perovskite films that needs to be understood in the context of the nature of the trap states and any residual non-radiative losses. Grain boundaries are a likely locus for defects sites, and large grain sizes, as viewed from scanning electron microscope (SEM) images, are implicitly preferred. Furthermore, polycrystalline perovskite films have been reported to show some degree of strain, which directly influences the macroscopic optoelectronic properties and stability,21–23 though it is unclear how the strain varies on the microscale, affects local recombination or the mechanism of the strain-related effects. Here, we directly probe the heterogeneity of high-quality perovskite thin films across multiple length scales to reveal local strain-related structural defects and their direct impact on optoelectronic behaviour.
In Fig. 1b and c we show the spatial maps of the peak scattering vector q of the azimuthally-integrated 〈222〉 and 〈220〉 peaks (see Fig. S4 for maps of peak intensities, ESI†). These maps reveal local structural heterogeneity on the scale of beam resolution. Fig. 1d and e show the μXRD patterns taken from horizontal slices as indicated by the coloured lines on Fig. 1b and c. Subtle shifts in the peak position and broadening reveal the presence of detailed microscale structural heterogeneity. This heterogeneity in the local q vector for the 〈220〉 orientation of ∼0.15%, corresponding to a spontaneous stress of ∼19 MPa (based on a Young's modulus of 12.8 GPa27), is typical of different regions of the films measured, and is a level of heterogeneity unobservable with laboratory diffraction techniques in both spatial resolution and peak variation. Interestingly, the μXRD slice in Fig. 1d depicts a region with long-range parallel coupling between the two reflections in local q variation, whereas Fig. 1e depicts a region with anti-parallel coupling between the two reflections (Fig. S5, ESI†). These observations are evidence for complex local strain variations,21 and long-range (>10 μm) strain patterns present throughout the polycrystalline film from regions that share similar crystallographic properties, as will be discussed further below.
After subtracting the instrumental contribution towards peak broadening of the line profiles, we consider the extreme cases of contributions from microstrain-only and crystallite-size only, as defined by the Williamson–Hall formalism.28 The variance of the crystallite-size-only is considerably larger and thus we conclude that microstrain is the dominant contributor towards peak breadth in these samples (see Fig. S6 and for further discussion, ESI†). We show the resulting microstrain map of the dominant 〈220〉 reflection in Fig. 1f (see Fig. S7 for 〈222〉, ESI†). This reveals that the microstrain also has a complex local heterogeneity with a typical magnitude of ∼0.1–0.2%, indicating that each microscale region has its own local strain environment. Importantly, there is a strong correlation between q and microstrain for the 〈220〉 peak (Fig. 1g; Fig. S7 for the 〈222〉, ESI†). That is, XRD peaks with the larger local q (lower d-spacing) contain the largest structural broadening due to microstrain (and vice versa). This suggests that the strain in the polycrystalline films is compressive, i.e. acting to reduce the volume of the unit cell. We observe similar correlations in an alloyed ‘triple cation’ MA0.15FA0.79Cs0.06Pb(I0.85Br0.15)3 sample (FA = formamidinium), suggesting this observation can be generalised to other compositions29 (Fig. S8, ESI†).
To directly assess the impact of the observed strain-related defects on the perovskite optoelectronic properties, we now correlate confocal PL measurements with μXRD measurements on the same scan area. For direct comparison of the two measurements, we use the Au fiducial markers (Fig. 1a inset) and an image analysis algorithm for image registration, while also accounting for the spatial resolution differences between the measurements (see ESI† and Fig. S9–S12 for details). We show a spatial map of the local compressive strain of a region of the film in Fig. 2a. Here, we determined the local compressive strain from μXRD using the relative shift of the peak q-value at each local point from the minimum q in the distribution (i.e. strain = (qmin − q)/qmin). We show in Fig. 2b a confocal PL intensity map of the correlation region highlighted in the strain map in Fig. 2a. We note that here we are at sufficiently low excitation fluence that the local PL distribution is dominated by trap states rather than diffusion of carriers out of the local region.30 We show the local time-resolved PL measurements of a bright region and a dim region in Fig. 2c representing the recombination of charge carriers. The bright regions have a longer PL lifetime than the dark regions, which is consistent with increased fractions of trap-limited recombination in the latter. We extract trap densities representative of the regions by fitting the decays with a kinetic model developed previously,16 quantifying the reduction in trap density from the dark (7.5 × 1016 cm−3) to the bright (1 × 1016 cm−3) regions. These dark PL (trap) regions are the cause of significant power losses in solar cells, with devices fabricated from these MAPbI3 films typically showing open-circuit voltages of ∼1.1 V,8 which is ∼0.2 V from the radiative limits.
Fig. 2 Correlating the local structural and time-resolved luminescence properties of MAPbI3 films. (a) Spatial map of the (compressive) strain variation using the relative shift of the peak q-value at each local point from the minimum q in a μXRD map (i.e. strain = (qmin − q)/qmin). The dashed line denotes the correlation region between μXRD and PL. (b) Confocal PL intensity map of a MAPbI3 perovskite film with pulsed 405 nm excitation (0.5 MHz repetition rate, 0.1 μJ per cm2 per pulse) corresponding to the dashed region in a. (c) Time-resolved PL decays of the bright (blue) and dark (red) regions highlighted in (b). The dotted lines are fits to the data using a trap model to extract the electronic trap density.16 Inset: Highlighted 〈220〉 peak diffraction pattern for the bright and dark PL regions. (d) Scatter plots of statistically-significant correlations between local PL lifetime and compressive strain (relative defect density; calculated from relationship in e). See Fig. S13 (ESI†) for other example PL decays. (e) Ratio in concentration of charged iodide vacancies (V+I defects) in 〈110〉 strained perovskite crystals to an unstrained crystal. |
We show in the inset of Fig. 2c the 〈220〉 μXRD peaks corresponding to the region with bright emission (long PL lifetime) and the region with dark emission (short PL lifetime). We find that the region with inferior emission intensity and carrier lifetime dynamics corresponds to a region with compressive-strained 〈220〉 lattice planes (i.e. larger q and increased peak broadening), whereas the region showing brighter emission and longer carrier lifetime is comparatively unstrained (i.e. smaller q and peak widths close to the instrumental broadening). We show in Fig. 2d that these trends appear consistently across the correlated mapped regions: scatter plots of the relevant quantities reveal a statistically-significant decrease in PL lifetime with compressive strain (see Fig. S14 for microstrain, ESI†).
Structural defects have a strong influence on the charge-carrier lifetime and recombination in semiconductors and charged point defects typically make efficient non-radiative recombination centers due to a long-range Coulomb attraction.17 We now explore the impact of a strained crystal on local point defect concentrations. Using a first-principles atomic model, we introduce compressive 〈110〉 strain with magnitude on the order observed in our local measurements (∼0.2%) and probe the effect on the point defect thermodynamics (Fig. 2e). With compressive strain, there is an increase in the charged iodide vacancy (V+I) concentration by a factor of 2 with respect to the unstrained crystal due to a negative defect pressure (see Methods and Fig. S15 and S16, ESI†). Similar behaviour is expected for other negative-pressure (strain releasing) defect species or clusters. This relative increase in defect density for a dark region ascertained from the trap model (Fig. 2c) is consistent with the calculated magnitude of the increase in strain-related halide vacancy concentrations after considering contributions from each direction (Fig. 2e). Using the relationship in Fig. 2e to convert the strain map to a relative defect density map (Fig. S17, ESI†), we show a strong anti-correlation between the PL lifetime and defect (halide vacancy) concentration ratio in Fig. 2d. These results reveal that the observation of local PL heterogeneity is substantially influenced by locally-heterogeneous strain distributions that are associated with defects such as halide vacancies. We note that we do not identify the isolated iodine vacancy as the origin of non-radiative recombination, and indeed other point and extended defects will be generated due to strain and likely also contribute to trap-assisted recombination.31–33
In order to further investigate the long range strain behaviour and its relationship to local grains, we performed scanning nanofocus XRD (nXRD) measurements at the ID13 beamline at the European Synchrotron Radiation Facility (ESRF) (Fig. S18, ESI†).26 A MAPbI3 perovskite film prepared as above was raster scanned (beam spot size 200 × 200 nm2; see Methods for details). We show a quiver plot for the 〈110〉 orientation in Fig. 3a, where the value of the azimuthal angle coordinate χp for each diffraction spot is represented using an arrow with its centre located in the spatial position from which the diffraction spot was acquired, and with an orientation and colour corresponding to χp indicated by the colour map.26 Diffraction spots adjacent both in real and reciprocal space coordinates were considered as belonging to the same cluster, here indicated as ‘super-grain’ (see Methods for details). We highlight in blue and red two super-grains with the largest covered areas calculated as the number of pixels within the super-grain times the pixel area (400 × 400 nm2). This observation of long-range features is consistent with our μXRD results, but here we visualise them with better spatial resolution. Furthermore, the super-grains also exhibit local strain (q) variations within their dimensions (Fig. 3b and c).
The super-grain sizes for the 〈110〉 reflection are plotted in Fig. 3d, showing that the largest regions cover an area of ∼25 μm2, extending well beyond the grain size observed in SEM images (∼1 μm2). We find this disparity is further exaggerated in the triple cation MA0.15FA0.79Cs0.06Pb(I0.85Br0.15)3 samples, which show super-grains as large as ∼250 μm2 (Fig. 3d) despite SEM grain sizes of only ∼0.1 μm2; this is clearly seen in the overlay of an SEM image and a quiver plot highlighting the largest super-grain in Fig. 3e (see Fig. S19 for other super-grains, ESI†). This finding could shed light onto the apparent paradox whereby ‘small-grain’ triple cation perovskite films still attain much higher PCE than these MAPbI3 films.11 Our results suggest that the critical grain size may actually be the longer-range structural super-grains rather than the grains viewed in SEM images.
To now explore crystallinity at a sub-grain resolution we turn to transmission electron microscopy (TEM). In Fig. 4a, we show a cross-sectional bright-field TEM image of the perovskite film. The sample was prepared by thinning down the 0.4 μm long lamella that appeared as an individual grain in the top-view SEM image by focused-ion beam (FIB). A selected-area electron diffraction (SAED) pattern obtained from a 200 nm region within the lamella is outlined by a circle in the micrograph and shown in the inset of Fig. 4a. The SAED pattern indicates a non-single crystalline nature of the “single grain” observed in SEM. Although the d-spacing corresponds to the tetragonal MAPbI3 perovskite structure, the presence of elongated diffraction spots as well as a weak diffraction ring is a strong indicator of imperfections within the lattice that likely originate from strain and/or other defects.
To probe crystallinity at a deeper scale, a high-resolution TEM (HR-TEM) micrograph was collected from a 70 × 70 nm2 region using a low electron dose rate of ∼1–4 electrons Å−2 s−1, in line with previous reports34 (Fig. 4b). The micrograph shows a lack of lattice continuality in the tested area as indicated by the presence of domains and structural defects. Fast Fourier transform (FFT) patterns generated from various 10 × 10 nm2 regions of the HR-TEM micrograph are outlined by the coloured boxes in Fig. 4c and clearly demonstrate structural heterogeneity within a single grain. The regions marked by black and purple boxes possess near-identical diffraction patterns and are highly crystalline as indicated by the sharp diffraction spots. The identical d-spacing of the diffraction spots indicate a similarity in crystallinity and sub-grain crystallite orientation on the 10 × 10 nm2 scale. The highly crystalline regions marked by the black and purple boxes contrast with regions bound by the red and blue boxes. The markedly different patterns indicate the whole grain is not uniformly crystallized. The blue region is well-crystallised but shows the presence of more than one diffraction pattern. This can be a result of multi-grain overlap or can indicate the presence of structural defects such as micro-twins and/or dislocations that can provide additional strain relief.34 In contrast, the red box shows a much weaker pattern indicating a poorly-crystallised or amorphous region within the same grain. While amorphisation due to beam damage cannot be excluded, a lack of homogeneity in the amorphisation signature through the sample makes irradiation-induced amorphisation less plausible, and the local changes in crystallographic orientation, or twinning, are highly unlikely to arise from beam damage.34 These results reveal that each of these grain “entities” are in fact comprised of many sub-crystallites on a ∼10–100 nm scale above that of the unit cell but below that of a single grain. This scale of heterogeneity is consistent with recent reports showing lower symmetry domains below 20 nm35 and substantial spatial variation in the photo-response of polycrystalline perovskite devices even within each grain.15,36,37
Seeking deposition methods that minimise local strain fields will be critical for achieving devices approaching the radiative limits. This includes developing compositions and lower temperature processing methods that do not require the material to proceed through a phase transition during film formation. As an example, the FA-rich alloyed perovskites demonstrating the highest solar cell device performances to date are closer to cubic at room temperature44 and thus do not undergo a phase transition between annealing and room temperature; stabilising such cubic structures at room temperatures will be critical for development of new alloyed analogues. In fact, our measurements performed on the MA0.15FA0.79Cs0.06Pb(I0.85Br0.15)3 films (cf.Fig. 3) reveal that there is reduced local heterogeneity in these samples and that the long range super-grain features exist over a much longer distance than the MAPbI3 counterparts, consistent with a reduced impact of strain. Judicious choice of the solar cell or LED device contacts will also be critical, as these may induce additional strain during perovskite growth (if underneath the perovskite) or during contact deposition (on top of the perovskite). This may require careful matching of the lattice parameters of contact materials to the perovskite; this has not been an issue to date but may become critical when approaching the performance limits in which all strain-related defects must be eliminated. Finally, passivation treatments must target relieving strain either through steric design of chemically passivating molecules or through managing the concentrations and spatial distribution of defects.2,6,7
Our work has revealed that this strain-related heterogeneity proliferates across many length scales throughout the entire film, from long-range super-grain clusters, to grain-to-grain and sub-grain nanoscale variations. This suggests that we need to re-think the conception of a perovskite grain as a single crystalline entity, for example as viewed in electron microscopy. This highlights important new questions on nucleation and grain growth. The large impact of local strain on local optoelectronic properties also opens up the ability to manipulate perovskite optoelectronic behaviour through purposeful engineering of local strain fields in novel device structures. For example, the formation of large super-grain clusters with analogous orientation and facet control may have beneficial properties for carrier transport, and could be utilised in lateral or back-contact device structures. Furthermore, the large perturbations to the crystal symmetry may lead to a Rashba effect in the electronic structure, which will further influence carrier dynamics,45 and this could be induced on a local level such as at a heterojunction interface. Nevertheless, the exceptional performance of perovskites in spite of so many layers of disorder is remarkable. Their behaviour is akin to liquid metals, which have disordered structures, yet maintain excellent charge transport properties.46 This sentiment would also explain the high open-circuit voltages in devices even in the earliest stages of disordered crystallite formation.47
Glass cover slips (PL, micro-XRD measurements) or Si〈100〉 substrates (nano-XRD, TEM measurements) were washed sequentially with soap (Micro 90), de-ionized water, acetone, and isopropanol, and treated under oxygen plasma for 10 minutes. Thin films of MAPbI3 were solution-processed by employing a methylammonium iodide (MAI) and lead acetate Pb(Ac)2·3H2O precursor mixture with a hypophosporous acid (HPA) additive.24 MAI and Pb(Ac)2·3H2O were dissolved in anhydrous N,N-dimethylformamide at a 3:1 molar ratio with final concentration of 37 wt% and HPA added to a HPA:Pb molar ratio of ∼11%. The precursor solution was spin-coated at 2000 rpm for 45 seconds in a nitrogen-filled glovebox, and the substrates dried at room temperature for 10 minutes before annealing at 100 °C for 5 minutes.
The triple-cation-based perovskite MA0.15FA0.79Cs0.06Pb(I0.85Br0.15)3 samples were prepared by dissolving PbI2 (1.2 M, TCI), FAI (1.11 M), MABr (0.21 M) and PbBr2 (0.21 M, TCI) in a mixture of anhydrous DMF:DMSO (4:1, volume ratios) followed by addition of 5 volume percent from CsI (TCI, Japan) stock solution (1.5 M in DMSO). We then spin-coated the perovskite solution using a two-step program at 2000 and 4000 rpm for 10 and 40 seconds, respectively, and dripping 150 μL of chlorobenzene after 30 seconds. We then annealed the films at 100 °C for 1 hour.
We note that the total light dose on each local region during the PL maps (<1 J cm−2) is much lower than that required to see light-induced changes in these samples.49,50 Furthermore, following the PL measurements, the samples were stored in vacuum bags before being loaded directly onto the chilled nitrogen flow for the XRD measurements (see below); control PL maps taken on films stored in a similar manner did not show any significant changes. This suggests that the illumination, storage and transport did not affect the samples.
Diffraction data was analysed as in ref. 26. Briefly, after the construction of an average diffraction pattern for the diffraction collected across the entire scan area for the MAPbI3 and MA0.15FA0.79Cs0.06Pb(I0.85Br0.15)3 films, we create circular regions of interest (ROIs) around the 〈110〉 and 〈210〉 perovskite reflection rings, respectively. Diffraction spots in the ROI-restricted pattern of the scan were analyzed and used to perform grain clustering based on the azimuthal angle coordinate χp extracted from the center of each diffraction spot. We assume that diffraction spots that are adjacent both in spatial coordinates and in reciprocal space coordinates originate from the same grain. This clustering is performed based on the pairwise Euclidean distance between pixels. We used an empirically-determined threshold value for cutting the hierarchical tree. The dataset for the identification of the diffraction spots is bi-dimensional (pixel X and pixel Y), and the dataset for the identification of the super-grains is tri-dimensional (2 spatial coordinates X and Y, and the azimuthal angle). The area of each ‘super-grain’ is simply calculated as the number of pixels within the ‘super-grain’ multiplied by the area of a pixel in microns. This gives a collection of super-grains (clusters) that can be ordered in decreasing size (Fig. 3d). Spatial maps of the q position were obtained from the line profiles extracted for all the diffraction spots in the ROI for each super-grain.
The equilibrium crystal structure was then subject to compressive and tensile strain (up to 0.5%) along the 〈110〉 and 〈111〉 crystallographic orientations. Next, the formation of iodine vacancy defects was probed in a series of 5 structures for each value of strain (ε), with the defect energy calculated from:
ΔED,ε = E[defect]ε − E[bulk]ε |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c8ee02751j |
‡ These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2019 |