+ All Categories
Home > Documents > Origin of Reversible Photoinduced Phase Separation in Hybrid … · 2017. 3. 8. · Origin of...

Origin of Reversible Photoinduced Phase Separation in Hybrid … · 2017. 3. 8. · Origin of...

Date post: 27-Apr-2021
Category:
Upload: others
View: 5 times
Download: 0 times
Share this document with a friend
6
Origin of Reversible Photoinduced Phase Separation in Hybrid Perovskites Connor G. Bischak, Craig L. Hetherington, ,Hao Wu, Shaul Aloni, §,D. Frank Ogletree, §,David T. Limmer, ,§,and Naomi S. Ginsberg* ,,,§,,# Department of Chemistry, University of California, Berkeley, California 94720, United States Molecular Biophysics and Integrative Bioimaging Division, § Materials Science Division, and Molecular Foundry, Lawrence Berkeley National Laboratory, Berkeley, California 94720, United States Kavli Energy NanoScience Institute, Berkeley, California 94720, United States # Department of Physics, University of California, Berkeley, California 94720, United States * S Supporting Information ABSTRACT: The distinct physical properties of hybrid organicinorganic materials can lead to unexpected non- equilibrium phenomena that are dicult to characterize due to the broad range of length and time scales involved. For instance, mixed halide hybrid perovskites are promising materials for optoelectronics, yet bulk measurements suggest the halides reversibly phase separate upon photoexcitation. By combining nanoscale imaging and multiscale modeling, we nd that the nature of halide demixing in these materials is distinct from macroscopic phase separation. We propose that the localized strain induced by a single photoexcited charge interacting with the soft, ionic lattice is sucient to promote halide phase separation and nucleate a light-stabilized, low-bandgap, 8 nm iodide-rich cluster. The limited extent of this polaron is essential to promote demixing because by contrast bulk strain would simply be relaxed. Photoinduced phase separation is therefore a consequence of the unique electromechanical properties of this hybrid class of materials. Exploiting photoinduced phase separation and other nonequilibrium phenomena in hybrid materials more generally could expand applications in sensing, switching, memory, and energy storage. KEYWORDS: Photoinduced phase transition, hybrid mixed halide perovskite, multiscale simulations, cathodoluminescence imaging, polaron P hotovoltaic and light-emitting devices typically operate under conditions far from equilibrium. As such, elucidating the response of functional materials to nonequilibrium driving forces is vital to understanding their fundamental physical properties and to determining their suitability for device applications. In particular, photoinduced dynamic processes are of major importance to the performance of hybrid perovskite- based devices. 14 Hybrid perovskites are low-cost, solution processable materials that are promising for many device applications, including photovoltaics 510 and light-emitting diodes (LEDs). 11 The high device eciencies have been attributed partly to their long charge carrier migration lengths 7,12 and tolerance of structural defects. 13,14 The chemical formula of hybrid perovskites is APbX 3 , where A is an organic cation, typically methylammonium (CH 3 NH 3 + , MA) or formamidinium (HC(NH 2 ) 2 + , FA), and X is either iodide, bromide, chloride, or iodide/bromide or bromide/chloride mixtures. By varying the halide ratios in hybrid perovskites, the bandgap can be tuned across the visible spectrum. 1,15,16 Precise control of the bandgap presents promising opportunities for color-tuning perovskite-based LEDs and lasers, and for incorporating hybrid perovskites in tandem solar cells. 5,17 Light-induced eects, however, restrict the practical use of mixed halide hybrid perovskites. 1,2,18,19 Photoluminescence (PL) and X-ray diraction (XRD) measurements suggest that MAPb(I x Br 1x ) 3 (0.1 < x < 0.8) undergoes reversible phase separation into iodide-rich and bromide-rich regions when photoexcited. 1 Such demixing is detrimental to photovoltaic performance because it leads to charge carrier trapping in the iodide-rich regions. Determining the microscopic mechanism behind phase separation is essential for furthering approaches to mitigate adverse photoinduced eects in devices and should expand the range of their functional applications into areas such as optical memory storage and sensing. 20,21 Unfortunately, the microscopic mechanism behind this eect has been elusive because of the multiple length and time scales involved in Received: October 24, 2016 Revised: January 23, 2017 Published: January 30, 2017 Letter pubs.acs.org/NanoLett © 2017 American Chemical Society 1028 DOI: 10.1021/acs.nanolett.6b04453 Nano Lett. 2017, 17, 10281033
Transcript
Page 1: Origin of Reversible Photoinduced Phase Separation in Hybrid … · 2017. 3. 8. · Origin of Reversible Photoinduced Phase Separation in Hybrid ... mixed halide hybrid perovskites.1,2,18,19

Origin of Reversible Photoinduced Phase Separation in HybridPerovskitesConnor G. Bischak,† Craig L. Hetherington,†,‡ Hao Wu,† Shaul Aloni,§,∥ D. Frank Ogletree,§,∥

David T. Limmer,†,§,⊥ and Naomi S. Ginsberg*,†,‡,§,⊥,#

†Department of Chemistry, University of California, Berkeley, California 94720, United States‡Molecular Biophysics and Integrative Bioimaging Division, §Materials Science Division, and ∥Molecular Foundry, Lawrence BerkeleyNational Laboratory, Berkeley, California 94720, United States⊥Kavli Energy NanoScience Institute, Berkeley, California 94720, United States#Department of Physics, University of California, Berkeley, California 94720, United States

*S Supporting Information

ABSTRACT: The distinct physical properties of hybridorganic−inorganic materials can lead to unexpected non-equilibrium phenomena that are difficult to characterize due tothe broad range of length and time scales involved. Forinstance, mixed halide hybrid perovskites are promisingmaterials for optoelectronics, yet bulk measurements suggestthe halides reversibly phase separate upon photoexcitation. Bycombining nanoscale imaging and multiscale modeling, we findthat the nature of halide demixing in these materials is distinctfrom macroscopic phase separation. We propose that thelocalized strain induced by a single photoexcited chargeinteracting with the soft, ionic lattice is sufficient to promotehalide phase separation and nucleate a light-stabilized, low-bandgap, ∼8 nm iodide-rich cluster. The limited extent of this polaronis essential to promote demixing because by contrast bulk strain would simply be relaxed. Photoinduced phase separation istherefore a consequence of the unique electromechanical properties of this hybrid class of materials. Exploiting photoinducedphase separation and other nonequilibrium phenomena in hybrid materials more generally could expand applications in sensing,switching, memory, and energy storage.

KEYWORDS: Photoinduced phase transition, hybrid mixed halide perovskite, multiscale simulations, cathodoluminescence imaging,polaron

Photovoltaic and light-emitting devices typically operateunder conditions far from equilibrium. As such, elucidating

the response of functional materials to nonequilibrium drivingforces is vital to understanding their fundamental physicalproperties and to determining their suitability for deviceapplications. In particular, photoinduced dynamic processes areof major importance to the performance of hybrid perovskite-based devices.1−4 Hybrid perovskites are low-cost, solutionprocessable materials that are promising for many deviceapplications, including photovoltaics5−10 and light-emittingdiodes (LEDs).11 The high device efficiencies have beenattributed partly to their long charge carrier migrationlengths7,12 and tolerance of structural defects.13,14 The chemicalformula of hybrid perovskites is APbX3, where A is an organiccation, typically methylammonium (CH3NH3

+, MA) orformamidinium (HC(NH2)2

+, FA), and X is either iodide,bromide, chloride, or iodide/bromide or bromide/chloridemixtures. By varying the halide ratios in hybrid perovskites, thebandgap can be tuned across the visible spectrum.1,15,16 Precisecontrol of the bandgap presents promising opportunities for

color-tuning perovskite-based LEDs and lasers, and forincorporating hybrid perovskites in tandem solar cells.5,17

Light-induced effects, however, restrict the practical use ofmixed halide hybrid perovskites.1,2,18,19 Photoluminescence(PL) and X-ray diffraction (XRD) measurements suggest thatMAPb(IxBr1−x)3 (0.1 < x < 0.8) undergoes reversible phaseseparation into iodide-rich and bromide-rich regions whenphotoexcited.1 Such demixing is detrimental to photovoltaicperformance because it leads to charge carrier trapping in theiodide-rich regions. Determining the microscopic mechanismbehind phase separation is essential for furthering approachesto mitigate adverse photoinduced effects in devices and shouldexpand the range of their functional applications into areas suchas optical memory storage and sensing.20,21 Unfortunately, themicroscopic mechanism behind this effect has been elusivebecause of the multiple length and time scales involved in

Received: October 24, 2016Revised: January 23, 2017Published: January 30, 2017

Letter

pubs.acs.org/NanoLett

© 2017 American Chemical Society 1028 DOI: 10.1021/acs.nanolett.6b04453Nano Lett. 2017, 17, 1028−1033

Page 2: Origin of Reversible Photoinduced Phase Separation in Hybrid … · 2017. 3. 8. · Origin of Reversible Photoinduced Phase Separation in Hybrid ... mixed halide hybrid perovskites.1,2,18,19

characterizing the behavior both experimentally and theoret-ically.We find that these novel photoinduced processes arise

because hybrid materials, including inorganic−organic perov-skites, have physical properties that differ substantially fromboth traditional inorganic and organic semiconductors. Inparticular, hybrid perovskites have elastic moduli that fall inbetween those of pure inorganic and organic solids,22,23 aidingin their tolerance of structural defects while maintaining long-range order. Hybrid perovskites also have a high static dielectricconstant, resulting in strong electron−phonon coupling, andlarge polarizabilities, resulting in small exciton bindingenergies.24−26 These macroscopic properties reflect the broadrange of time scales characterizing molecular motion in thesematerials. For example, motions associated with freely rotatingorganic cations and facile halide migration have been implicatedin the large dielectric constant of the material27,28 and inhysteresis upon charging.29,30 Understanding the interplaybetween and molecular origin of the mechanical and electronicproperties of these hybrid materials provides insight into how ahierarchy of bonding energies can result in multifacetedcollective phenomena. These insights will lend predictivepower to the development of new high-performance functionalmaterials.Here we describe how photoinduced phase separation in

mixed halide hybrid perovskites can be mediated throughstrain-induced phase separation at the locations of polaronsphotogenerated charge carriers and their accompanying latticedistortions. The heterogeneous nature of this nonequilibriumhalide phase separation that we reveal starkly contrasts themacroscopic domains that would form under equilibriumconditions. We combine cathodoluminescence (CL) imaging, apowerful tool for characterizing the nanoscale optical propertiesof hybrid perovskite films,14,31,32 and multiscale modeling, a setof techniques to bridge molecular and mesoscopic scales, toobserve and explain the dynamic process of photoinducedphase separation. After prolonged illumination, small clustersenriched in one halide species are observed to localize neargrain boundaries, which is consistent with the effects ofpolaronic strain in both our molecular and mean-fieldphenomenological models and the associated phase diagramthat we construct. The transient dynamics of cluster formationare characterized by an initial latency and stochastic fluctuationsin their formation. The process of cluster formation is capturedwith the phenomenological theory at intermediate electron−phonon coupling and consists of photogenerated polaronsfinding, stabilizing, and subsequently becoming trapped inhalide composition fluctuations. Additional imaging andcalculations validate the proposed requirement that polaronsreside specifically within the lower bandgap phase-separatedregions to ensure their stability and suggest a path toward newoptical memory or detection applications.To observe the nanoscale features of photoinduced phase

separation, we fabricated mixed halide hybrid perovskite filmsfor CL imaging. We made a series of films with varying iodide/bromide ratios (Figure 1a) and ultimately selected MAPb-(I0.1Br0.9)3 for further investigation because of its spectrallydistinct iodide-rich and bromide-rich regions. When initiallyphotoexciting the MAPb(I0.1Br0.9)3 film, we observe a singleemission peak at 540 nm, corresponding to well-mixed halides.Upon continued illumination, a second spectral feature appearsat 690 nm (Figure 1b), which corresponds to iodide-richcontent and is thus indicative of phase separation (Figure 1c).

When placed in the dark, the halides remix and the redemission peak disappears on a similar time scale. We illuminatethe sample with a 405 nm LED for light soaking inside ascanning electron microscope (SEM) fitted with CL collectioncapabilities (Figure 1d) and then scan the electron beam toprobe photoinduced iodide-rich cluster formation at thenanoscale. The electron beam by itself cannot induce phaseseparation (see Supporting Information). Figure 1e shows asecondary electron (SE) image, CL image, and a SE/CL overlaycollected after 5 min illumination at 100 mW/cm2, demonstrat-ing that the iodide-rich regions localize to grain boundaries insteady state. Although the size of the bright spots in the CLimage results from a convolution of iodide cluster size and a

Figure 1. Band gap tunability, PL spectra before and after lightsoaking, CL experimental setup, and CL/SE images at steady state.(A) PL spectra of MAPb(IxBr1−x)3 films with x = 0 (green), x = 0.2(yellow), x = 0.5 (orange), x = 0.7 (red), and x = 1 (black). (B) PLspectra of MAPb(I0.1Br0.9)3 before (green) and after (red) lightsoaking for 5 min. (C) Schematic of phase separation and reversibilityin MAPb(IxBr1−x)3 where yellow and blue spheres represent I− andBr−, respectively, the red and white pill shapes represent MA, and thelead atoms (not shown) are located in the center of the octahedra. (D)Schematic of CL acquisition and light soaking. (E) SE, CL, and SE/CLoverlay after light soaking for 5 min at 100 mW/cm2. The scale bar is 2μm.

Nano Letters Letter

DOI: 10.1021/acs.nanolett.6b04453Nano Lett. 2017, 17, 1028−1033

1029

Page 3: Origin of Reversible Photoinduced Phase Separation in Hybrid … · 2017. 3. 8. · Origin of Reversible Photoinduced Phase Separation in Hybrid ... mixed halide hybrid perovskites.1,2,18,19

large carrier migration length, individual clusters are still visiblewithin each grain.Molecular simulations suggest that photoinduced phase

separation arises when charged excitations generate sufficientlattice strain to destabilize the solid solution, favoring demixing.Using a classical point charge model27,33 we find that mixed I/Br perovskites undergo demixing transitions as a function oftemperature with a critical temperature of 190 K (Figure S1),consistent with the well-mixed films made experimentally atroom temperature. Analysis of Br/Cl perovskites find asimilarly low critical solution temperature of 140 K. For I/Clperovskites, however, the critical solution temperature (1800K) is found to be well above room temperature, consistent withexperimental observations that such mixtures are unstable.34

While direct experimental evidence of temperature-induceddemixing is hampered experimentally by low ion conductivities,these predictions are in reasonable agreement with othertheoretical estimates.35 By analyzing the relative energeticcontributions to the heat of mixing, we find that elastic effectsfrom lattice mismatch are much larger than specific chemicalinteractions, which leads to a demixing transition that dependsstrongly on strain (Figure S2). As illustrated in Figure 2a, uponlight absorption weakly bound electron−hole pairs, withbinding energies Ex ≈ 0.03 eV,25 rapidly dissociate, easilycreating free charges. These charges deform the surroundinglattice through electron−phonon coupling, which is expected tobe significant given the ionic nature of the material. Using free

energy calculations and path integral molecular dynamics of apseudopotential-based model of an excess charge,36 shown inFigure 2b and Video S1, we find that high spatial overlapbetween the lattice and a single-charge density distributiongenerates sufficient strain to drive local phase separation atroom temperature. By contrast, uniform bulk strain could berelaxed through global volume changes, precluding phaseseparation. Figure 2c shows the free energy with and withoutthe excess charge as a function of composition, βΔF(x), withina volume characteristic of the charge distribution’s extent. Thischarge and the lattice deformation field that surrounds ittogether form a polaron that we predict to have an average sizeof 8 nm and binding energy Ep = 0.08 eV (Figure 2a). Asdiscussed below, the lower bandgap of the iodide-enrichedphase energetically stabilizes and spatially traps the polaron.To relate these findings to the experimental observations, we

have distilled them into a simple analytical theory. Modelingthe phase separation with a Landau−Ginzburg Hamiltonianwith linear coupling between strain and composition fields37

and applying a semiclassical description of the excess charge,38

we parametrize and evaluate a theory for photoinduced phaseseparation (see Supporting Information). As shown in Figure2c, this model is capable of describing the underlying freeenergy surfaces predicted from molecular dynamics simulations.Within a mean-field approximation, we determine the fulltemperature−composition phase diagram for both ground- andphotoexcited states (Figure 2d). We confirm experimentally

Figure 2. Steady-state stabilities and dynamic mechanism for photoinduced phase separation. (A) Photoinduced polaron trapping and associatedenergy scales associated with phase separation. Color scheme is the same as in Figure 1 with the addition of lead atoms represented by gray circles.(B) Snapshot of the 99% isosurface of excess charge density taken from the molecular dynamics (MD) simulation. (C) Free energies per unit cell forMAPb(IxBr1−x)3 with varying composition in the ground (red) and photoexcited (blue) states, computed from MD simulations (circles) and meanfield theory (solid lines). (D) Mean field theory temperature−composition phase diagram in the ground (red) and photoexcited state (blue) with thepath through the phase diagram from initial state (star) to demixed state (circle) observed experimentally. Areas beneath the red and bluecoexistence curves indicate demixed states. (E) The extent of demixing (i.e., purity of demixed regions found by tracing the blue coexistence curve inD) as a function of electron−phonon coupling and temperature computed from mean field theory in the photoexcited state.

Nano Letters Letter

DOI: 10.1021/acs.nanolett.6b04453Nano Lett. 2017, 17, 1028−1033

1030

Page 4: Origin of Reversible Photoinduced Phase Separation in Hybrid … · 2017. 3. 8. · Origin of Reversible Photoinduced Phase Separation in Hybrid ... mixed halide hybrid perovskites.1,2,18,19

that varying the film temperature over a range of 50 K isinsufficient to induce demixing, but does increase the demixingrate by increasing halide mobility (Figure S3). We also map thedegree of demixing as a function of temperature and electron−phonon coupling, α, in the photoexcited state (Figure 2e).Previous PL studies confirm that the stable iodide-rich clusterphase is at approximately MAPb(I0.8Br0.2)3,

1 which agrees withour predicted phase diagram and PL measurements.To confirm the validity of our thermodynamic model for

photoinduced phase separation, we demonstrate experimentallyboth the ability of polarons to stabilize iodide-rich clusters andthat large electron−phonon coupling is required to cause halidephase separation. For the latter, we show both experimentallyand with the phenomenological model that reducing theelectron−phonon coupling in the system by replacing MA+

with less polar Cs+ reduces the tendency to phase separate(Figure S4), consistent with the phase diagram in Figure 2e. Todemonstrate that the local presence of a polaron stabilizesiodide-rich clusters in the hybrid perovskites, specifically inMAPb(I0.1Br0.9)3, we first generate iodide-rich clusters with a405 nm excitation source. We then illuminate the film with 635nm light, generating charge carriers only in pre-existing iodide-rich clusters. The 635 nm light stabilizes these clusters but doesnot generate new ones (Figure S5 and S6), confirming that thecontinued presence of photogenerated carriers is required forcluster stability.Having established a model for the thermodynamics of halide

phase separation, we focus on the dynamics of the formationand evolution of iodide-rich clusters, which further constrains

our model. To monitor the formation of iodide-rich clusters inbulk, we collected a series of PL emission spectra underconstant light-soaking at 50 mW/cm2 (Figure 3a) and plot theintegrated intensity of iodide-rich emission as a function of timeas it grows to a finite value proportional to the illuminationintensity (Figure 3b, red). To resolve the emergence andevolution of iodide-rich clusters, we alternate between lightsoaking and CL imaging (Figure 3c). Figure 3d and 3e (VideosS2 and S3) show overlaid CL and SE image sequences with 10and 30 s intervals of light soaking, respectively. After somelatency, iodide-rich clusters begin to emerge and grow primarilyin number. They quickly reach a maximum size, suggesting thatcluster nucleation rather than growth limits the rate of phaseseparation. We also find using CL microscopy that the rate ofcluster formation increases with illumination intensity (FigureS7, Video S4 and S5).To gain insight into the microscopic dynamics leading to

cluster formation, we examine the photoinduced progressionwithin single domains with both CL (Figure 3f and Figure S8)and simulations (Figure 3g). Simulations of the clusteringprocess are performed by solving our phenomenological theorynumerically on a lattice (Video S6). We observe similardynamics in both experiment and simulation, such as theformation of both transient and stabilized clusters. An ensembleaverage of independent simulated clustering events yields acurve that characterizes the clustering process (Figure 3b,black), which agrees with the experimentally observed PLintensity growth in time (red) and its illumination powerdependence (Figures S9 and S10). We find that the 5−10 s

Figure 3. Formation and evolution of iodide-rich clusters. (A) PL spectra after different light soaking times at 50 mW/cm2. (B) Normalized PLintensity versus time (red) and normalized simulated cluster size versus time (black) with standard error values (blue). (C) Duty cycle for CL imageseries. CL image series with (D) 10 and (E) 30 s of light soaking between each CL image. The scale bars are 2 μm. (F) CL image series of a singledomain with 10 s of light soaking between each CL image. The color scale indicates iodide-rich CL intensity, which convolves carrier migrationlength with feature size. The scale bar is 200 nm. (G) A series of snapshots from a cluster formation simulation of a 100 nm region with iodide-richregions in yellow and bromide-rich regions in blue.

Nano Letters Letter

DOI: 10.1021/acs.nanolett.6b04453Nano Lett. 2017, 17, 1028−1033

1031

Page 5: Origin of Reversible Photoinduced Phase Separation in Hybrid … · 2017. 3. 8. · Origin of Reversible Photoinduced Phase Separation in Hybrid ... mixed halide hybrid perovskites.1,2,18,19

initial lag time corresponds to a characteristic time scale for apolaron to become trapped in a spontaneous fluctuation ofhigher iodide concentration, which is limited by the ions’diffusivity rather than the polaron’s diffusivity. During thefollowing 10−90 s, a trapped polaron stabilizes an iodide-richcluster and accumulates more iodide, and more clusters formwithin the film. Subsequently, cluster growth stops, as clustersize is limited to the deformation region of the polaron, andcluster number is limited by the total number of photo-generated charges. Although the time scales for these dynamicsare considerably longer than the lifetime of a single polaron,continuous illumination enables newly generated polarons toreplace recombining ones.39 Newly generated polaronscontinue to stabilize clusters due to the much larger mobilityof the excess charge relative to the iodide, resulting in acharacteristic localization time for the polaron that is muchsmaller than the characteristic time for a cluster to relax. Bymeasuring the PL of the iodide-rich clusters at differentintensities (Figure S9), which correspond to different steady-state polaron densities, we estimate the size of the clusters to be8−10 nm in diameter (Figure S11), which is in agreement withour theoretical prediction. During later times, the clustersmigrate to grain boundaries to relieve strain (Figure 1e). Werule out that the enhanced CL emission near the grainboundaries comes from outcoupling effects due to the filmtopography because previous CL studies on lead halideperovskite thin films did not observe such effects.14

The confluence of multiple physical processes that coupledisparate length and time scales gives rise to the decidedlypeculiar photoinduced phase separated state. For example,because the strain field applied by each polaron is limited inspatial extent, the photoexcited steady state of the systemconsists of a series of isolated nanoscale clusters rather than theexpected large, single iodide-rich domain of a low-temperatureequilibrium state in the dark (Figure 3d). Further, becausepolarons preferentially localize in iodide-rich regions, thecomposition of the remaining material is neither predictednor observed to deviate from its preillumination halide mixtureratio (Figure 1b), as though it were still in the dark. Last,because the iodide-rich clusters are limited to nanoscale sizes,their coarsening dynamics do not abide by familiar, universalscaling laws.We have established that the unusual interplay between the

electronic and mechanical properties in hybrid perovskitematerials results in local, excited-state phase behavior thatdiffers substantially from the ground state due to polaronicstrain. Photoinduced phase separation is a general phenomenonin hybrid perovskites that has been observed in bothpolycrystalline films and single crystals and with differentorganic cations.1 We further find both experimentally andtheoretically that it occurs with other halide mixtures, such as inMAPb(BrxCl1−x)3 thin films (Figures S12 and S13). Our studyshows that the unique combination of mobile halides,substantial electron−phonon coupling, and long-lived chargecarriers is required for photoinduced phase separation.Decreasing defect concentrations to reduce vacancy-mediatedhalide migration or lowering electron−phonon coupling couldsignificantly reduce photoinduced effects and improve compat-ibility with device applications at ambient conditions. Forinstance, it has been shown that Cs-doped FAPb(IxBr1−x)3 filmsare much more phase-stable under illumination than undopedfilms, likely due to a decrease in the electron−phononcoupling.5

Taking advantage of the generality of photoinduced phaseseparation could provide new opportunities for expanding thefunctional applications of mixed halide hybrid perovskites.Because of their sensitive spectral photoresponse, thesematerials could be used in sensing, switching, or memoryapplications. In fact, we have demonstrated a first step towardmemory storage by transiently patterning the local halidecomposition (Figure S6). Because we show that a single excitedcharge is responsible for locally changing the structure andcomposition of the material, truly nanoscale memory elementscould even be realized via single charge injection. More broadly,the commonality in the electronic and mechanical propertiesamong a rapidly growing library of new hybrid materialssuggests that nonequilibrium processes in these materials aresimilar and could be exploited for new device applications. Thecoordinated multiscale methodologies that we have developedcould also uncover the nature of other nonequilibriumphenomena in other dynamic materials, for example in energystorage40 or electronically correlated materials,41for which theircomplexity necessitates a multifaceted approach, capable ofbridging molecular and mesoscopic length scales.

■ ASSOCIATED CONTENT*S Supporting InformationMaterials and methods, an explanation for why the electronbeam does not cause demixing, Figures S1−S13, Videos S1−S6,Tables S1−S2, and supporting references. This materials isavailable free of charge via the Internet at The SupportingInformation is available free of charge on the ACS Publicationswebsite at DOI: 10.1021/acs.nanolett.6b04453.

Materials and methods, an explanation for why theelectron beam does not cause demixing, Figures S1−S13,Tables S1 and S2, and supporting references (PDF)Molecular dynamics simulation of a 99% isosurface ofexcess charge density on the perovskite lattice (MPG)Sequence of CL images of phase separation with 10 slight soaking intervals at 50 mW/cm2 (AVI)Sequence of CL images of phase separation with 30 slight soaking intervals at 50 mW/cm2 (AVI)Sequence of CL images of phase separation with 10 slight soaking intervals at 1100 mW/cm2 (AVI)Sequence of CL images of phase separation with 30 slight soaking intervals at 1100 mW/cm2 (AVI)Cluster formation simulation (MPG)

■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected] S. Ginsberg: 0000-0002-5660-3586Present Address(H.W.) Department of Chemistry and Chemical Biology,Harvard University, Cambridge, MA 02138, United States.Author ContributionsC.G.B. and N.S.G. designed the CL and PL experiments. C.G.Bfabricated the perovskite films, acquired the CL and PL data,and performed all experimental data analysis. D.T.L. designedthe models and performed all simulations. C.L.H., H.W., S.A.,and D.F.O. produced the CL acquisition software and providedinstrument support and helpful discussion. C.G.B., D.T.L., and

Nano Letters Letter

DOI: 10.1021/acs.nanolett.6b04453Nano Lett. 2017, 17, 1028−1033

1032

Page 6: Origin of Reversible Photoinduced Phase Separation in Hybrid … · 2017. 3. 8. · Origin of Reversible Photoinduced Phase Separation in Hybrid ... mixed halide hybrid perovskites.1,2,18,19

N.S.G. wrote the manuscript and all authors revised andapproved the manuscript.

NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTS

CL characterization was supported by a David and LucilePackard Fellowship for Science and Engineering to N.S.G. CLand PL at the Lawrence Berkeley Lab Molecular Foundry wereperformed as part of the Molecular Foundry user program,supported by the Office of Science, Office of Basic EnergySciences, of the U.S. Department of Energy under Contract No.DE-AC02-05CH11231. C.G.B. acknowledges an NSF GraduateResearch Fellowship (DGE 1106400) and N.S.G. acknowledgesan Alfred P. Sloan Research Fellowship and a Camille DreyfusTeacher-Scholar Award. We thank Z. Luo for assistance withfilm deposition. D.T.L was supported initially through thePrinceton Center of Theoretical Science.

■ REFERENCES(1) Hoke, E. T.; Slotcavage, D. J.; Dohner, E. R.; Bowring, A. R.;Karunadasa, H. I.; McGehee, M. D. Chem. Sci. 2015, 6 (1), 613−617.(2) Yoon, S. J.; Draguta, S.; Manser, J. S.; Sharia, O.; Schneider, W.F.; Kuno, M.; Kamat, P. V. ACS Energy Lett. 2016, 1 (1), 290−296.(3) Gottesman, R.; Zaban, A. Acc. Chem. Res. 2016, 49 (2), 320−329.(4) deQuilettes, D. W.; Zhang, W.; Burlakov, V. M.; Graham, D. J.;Leijtens, T.; Osherov, A.; Bulovic, V.; Snaith, H. J.; Ginger, D. S.;Stranks, S. D. Nat. Commun. 2016, 7, 11683.(5) McMeekin, D. P.; Sadoughi, G.; Rehman, W.; Eperon, G. E.;Saliba, M.; Horantner, M. T.; Haghighirad, A.; Sakai, N.; Korte, L.;Rech, B.; Johnston, M. B.; Herz, L. M.; Snaith, H. J. Science 2016, 351(6269), 151−155.(6) Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith,H. J. Science 2012, 338 (6107), 643−647.(7) Stranks, S. D.; Eperon, G. E.; Grancini, G.; Menelaou, C.;Alcocer, M. J. P.; Leijtens, T.; Herz, L. M.; Petrozza, A.; Snaith, H. J.Science 2013, 342 (6156), 341−344.(8) Burschka, J.; Pellet, N.; Moon, S.-J.; Humphry-Baker, R.; Gao, P.;Nazeeruddin, M. K.; Gratzel, M. Nature 2013, 499 (7458), 316−319.(9) Zhou, H.; Chen, Q.; Li, G.; Luo, S.; Song, T.; Duan, H.-S.; Hong,Z.; You, J.; Liu, Y.; Yang, Y. Science 2014, 345 (6196), 542−546.(10) Nie, W.; Tsai, H.; Asadpour, R.; Blancon, J.-C.; Neukirch, A. J.;Gupta, G.; Crochet, J. J.; Chhowalla, M.; Tretiak, S.; Alam, M. A.;Wang, H.-L.; Mohite, A. D. Science 2015, 347 (6221), 522−525.(11) Tan, Z.-K.; Moghaddam, R. S.; Lai, M. L.; Docampo, P.; Higler,R.; Deschler, F.; Price, M.; Sadhanala, A.; Pazos, L. M.; Credgington,D.; Hanusch, F.; Bein, T.; Snaith, H. J.; Friend, R. H. Nat. Nanotechnol.2014, 9 (9), 687−692.(12) Dong, Q.; Fang, Y.; Shao, Y.; Mulligan, P.; Qiu, J.; Cao, L.;Huang, J. Science 2015, 347 (6225), 967−970.(13) de Quilettes, D. W.; Vorpahl, S. M.; Stranks, S. D.; Nagaoka, H.;Eperon, G. E.; Ziffer, M. E.; Snaith, H. J.; Ginger, D. S. Science 2015,348 (6235), 683−686.(14) Bischak, C. G.; Sanehira, E. M.; Precht, J. T.; Luther, J. M.;Ginsberg, N. S. Nano Lett. 2015, 15 (7), 4799−4807.(15) Noh, J. H.; Im, S. H.; Heo, J. H.; Mandal, T. N.; Seok, S. I. NanoLett. 2013, 13 (4), 1764−1769.(16) Sadhanala, A.; Ahmad, S.; Zhao, B.; Giesbrecht, N.; Pearce, P.M.; Deschler, F.; Hoye, R. L. Z.; Godel, K. C.; Bein, T.; Docampo, P.;Dutton, S. E.; De Volder, M. F. L.; Friend, R. H. Nano Lett. 2015, 15(9), 6095−6101.(17) Bailie, C. D.; Christoforo, M. G.; Mailoa, J. P.; Bowring, A. R.;Unger, E. L.; Nguyen, W. H.; Burschka, J.; Pellet, N.; Lee, J. Z.;Gratzel, M.; Noufi, R.; Buonassisi, T.; Salleo, A.; McGehee, M. D.Energy Environ. Sci. 2015, 8 (3), 956−963.

(18) Beal, R. E.; Slotcavage, D. J.; Leijtens, T.; Bowring, A. R.; Belisle,R. A.; Nguyen, W. H.; Burkhard, G. F.; Hoke, E. T.; McGehee, M. D. J.Phys. Chem. Lett. 2016, 7 (5), 746−751.(19) Jaffe, A.; Lin, Y.; Beavers, C. M.; Voss, J.; Mao, W. L.;Karunadasa, H. I. ACS Cent. Sci. 2016, 2, 201−209.(20) Lin, Q.; Armin, A.; Lyons, D. M.; Burn, P. L.; Meredith, P. Adv.Mater. 2015, 27 (12), 2060−2064.(21) Sutherland, B. R.; Johnston, A. K.; Ip, A. H.; Xu, J.; Adinolfi, V.;Kanjanaboos, P.; Sargent, E. H. ACS Photonics 2015, 2 (8), 1117−1123.(22) Rakita, Y.; Cohen, S. R.; Kedem, N. K.; Hodes, G.; Cahen, D.MRS Commun. 2015, 5 (4), 623−629.(23) Sun, S.; Fang, Y.; Kieslich, G.; White, T. J.; Cheetham, A. K. J.Mater. Chem. A 2015, 3 (36), 18450−18455.(24) Onoda-Yamamuro, N.; Matsuo, T.; Suga, H. J. Phys. Chem.Solids 1992, 53 (7), 935−939.(25) Johnston, M. B.; Herz, L. M. Acc. Chem. Res. 2016, 49 (1), 146−154.(26) Galkowski, K.; Mitioglu, A.; Miyata, A.; Plochocka, P.; Portugall,O.; Eperon, G. E.; Wang, J. T.-W.; Stergiopoulos, T.; Stranks, S. D.;Snaith, H. J.; Nicholas, R. J. Energy Environ. Sci. 2016, 9 (3), 962−970.(27) Mattoni, A.; Filippetti, A.; Saba, M. I.; Delugas, P. J. Phys. Chem.C 2015, 119 (30), 17421−17428.(28) Leguy, A. M. A.; Frost, J. M.; McMahon, A. P.; Sakai, V. G.;Kockelmann, W.; Law, C.; Li, X.; Foglia, F.; Walsh, A.; O’Regan, B. C.;Nelson, J.; Cabral, J. T.; Barnes, P. R. F. Nat. Commun. 2015, 6, 7124.(29) Unger, E. L.; Hoke, E. T.; Bailie, C. D.; Nguyen, W. H.;Bowring, A. R.; Heumuller, T.; Christoforo, M. G.; McGehee, M. D.Energy Environ. Sci. 2014, 7 (11), 3690−3698.(30) Snaith, H. J.; Abate, A.; Ball, J. M.; Eperon, G. E.; Leijtens, T.;Noel, N. K.; Stranks, S. D.; Wang, J. T.-W.; Wojciechowski, K.; Zhang,W. J. Phys. Chem. Lett. 2014, 5 (9), 1511−1515.(31) Dou, L.; Wong, A. B.; Yu, Y.; Lai, M.; Kornienko, N.; Eaton, S.W.; Fu, A.; Bischak, C. G.; Ma, J.; Ding, T.; Ginsberg, N. S.; Wang, L.-W.; Alivisatos, A. P.; Yang, P. Science 2015, 349 (6255), 1518−1521.(32) Xiao, C.; Li, Z.; Guthrey, H.; Moseley, J.; Yang, Y.; Wozny, S.;Moutinho, H.; To, B.; Berry, J. J.; Gorman, B.; Yan, Y.; Zhu, K.; Al-Jassim, M. J. Phys. Chem. C 2015, 119 (48), 26904−26911.(33) Lewis, G. V.; Catlow, C. R. A. J. Phys. C: Solid State Phys. 1985,18 (6), 1149.(34) Yang, B.; Keum, J.; Ovchinnikova, O. S.; Belianinov, A.; Chen,S.; Du, M.-H.; Ivanov, I. N.; Rouleau, C. M.; Geohegan, D. B.; Xiao, K.J. Am. Chem. Soc. 2016, 138 (15), 5028−5035.(35) Brivio, F.; Caetano, C.; Walsh, A. J. Phys. Chem. Lett. 2016, 7(6), 1083−1087.(36) Kuharski, R. A.; Bader, J. S.; Chandler, D.; Sprik, M.; Klein, M.L.; Impey, R. W. J. Chem. Phys. 1988, 89 (5), 3248.(37) Cahn, J. W. Acta Metall. 1962, 10 (3), 179−183.(38) Kalosakas, G.; Aubry, S.; Tsironis, G. P. Phys. Rev. B: Condens.Matter Mater. Phys. 1998, 58 (6), 3094−3104.(39) Yang, X.; Yan, X.; Wang, W.; Zhu, X.; Li, H.; Ma, W.; Sheng, C.Org. Electron. 2016, 34, 79−83.(40) Li, Y.; El Gabaly, F.; Ferguson, T. R.; Smith, R. B.; Bartelt, N.C.; Sugar, J. D.; Fenton, K. R.; Cogswell, D. A.; Kilcoyne, A. L. D.;Tyliszczak, T.; Bazant, M. Z.; Chueh, W. C. Nat. Mater. 2014, 13 (12),1149−1156.(41) Schmitt, F.; Kirchmann, P. S.; Bovensiepen, U.; Moore, R. G.;Rettig, L.; Krenz, M.; Chu, J.-H.; Ru, N.; Perfetti, L.; Lu, D. H. Science2008, 321 (5896), 1649−1652.

Nano Letters Letter

DOI: 10.1021/acs.nanolett.6b04453Nano Lett. 2017, 17, 1028−1033

1033


Recommended