You are on page 1of 5

R ES E A RC H | R E PO R TS

that generally depends on the asperity size and 27. C. H. Scholz, Geophys. Res. Lett. 42, 1399–1402 (2015). 45. J. H. Shaw et al., Earth Planet. Sci. Lett. 415, 1–15 (2015).
stress drop and on the resistance of the matrix. 28. C. H. Scholz, Bull. Seismol. Soc. Am. 58, 399–415 (1968). 46. C. Tape, A. Plesch, J. H. Shaw, H. Gilbert, Seismol. Res. Lett.
29. M. Spada, T. Tormann, S. Wiemer, B. Enescu, Geophys. Res. 83, 728–735 (2012).
This effective radius Re controls the range of inter- Lett. 40, 709–714 (2013). 47. Y. Ma, R. Clayton, Geophys. J. Int. 206, 1645–1651 (2016).
action between asperities. The ratio between Re 30. T. Watanabe, Y. Hiramatsu, K. Obara, Geophys. Res. Lett. 34, 48. Z. Yan, R. W. Clayton, J. Geophys. Res. 112, B05311
and the interasperity distance D determines the L07305 (2007). (2007).
ability of asperities to break together in seismic 31. D. R. Shelly, J. L. Hardebeck, Geophys. Res. Lett. 37, L14301
(2010). AC KNOWLED GME NTS
events, despite the intervening creep, and thus 32. J. R. Sweet, K. C. Creager, H. Houston, Geochem. Geophys. We thank Signal Hill Petroleum and NodalSeismic for granting us
influences the statistics of the earthquake catalog. Geosyst. 15, 3713–3721 (2014). permission to use the Long Beach Array data, and we thank
When Re/D is large, ruptures can involve multiple 33. M. G. Bostock, A. M. Thomas, G. Savard, L. Chuang, Breitburn Energy and LA Seismic for permission to use the
asperities. This strong interaction regime poten- A. M. Rubin, J. Geophys. Res. 120, 6329–6350 (2015). Rosecrans Array data. We acknowledge J. P. Avouac, R. Bürgmann,
34. B. Schmandt, R. W. Clayton, J. Geophys. Res. 118, 5320–5338 Y. Ma, and W. Frank for helpful discussions. This research was
tially leads to a scale-free, power-law earthquake (2013). supported by NSF awards EAR-1214912 and EAR-1520081 and by
size distribution (Fig. 3A) and temporal clustering 35. E. Hauksson, J. Geophys. Res. 105, 13875–13903 (2000). the Terrestrial Hazard Observation and Reporting Center at
(Fig. 3C), as observed at shallow depths. When 36. T. Romanyuk, W. D. Mooney, S. Detweiler, J. Geodyn. 43, Caltech. The seismic data are property of Signal Hill Petroleum and
Re/D is small, asperities tend to break in isolation. 274–307 (2007). Breitburn Energy. Data are available for noncommercial use
37. E. Hauksson, Bull. Seismol. Soc. Am. 77, 539–561 (1987). through a license agreement with the data owners that includes
In this weak interaction regime, seismicity is 38. R. Porter, G. Zandt, N. McQuarrie, Lithosphere 3, 201–220 (2011). but is not limited to a nondistribution agreement. Please contact
temporally uncorrelated and, if asperities have 39. P. Audet, J. Geophys. Res. 120, 3527–3543 (2015). the authors for additional information.
a characteristic size, the earthquake size distri- 40. D. S. H. King, C. Marone, J. Geophys. Res. 117, B12203 (2012).
bution is scale-bound, as observed in the deep NIF 41. E. K. Mitchell, Y. Fialko, K. M. Brown, Geochem. Geophys. SUPPLEMENTARY MATERIALS
Geosyst. 16, 4006–4020 (2015).
beneath LB. A systematic decrease of Re/D with www.sciencemag.org/content/354/6308/88/suppl/DC1

Downloaded from http://science.sciencemag.org/ on October 6, 2016


42. T. Ueda, M. Obata, G. Di Toro, K. Kanagawa, K. Ozawa, Geology
increasing depth may result from several processes, Materials and Methods
36, 607–610 (2008).
Figs. S1 to S7
which are not necessarily independent. One possi- 43. A. K. Matysiak, C. A. Trepmann, Tectonophys. 530–531,
References (49, 50)
bility is a rheological control: Re may decrease 111–127 (2012).
44. A. J. Getsinger, G. Hirth, H. Stunitz, E. T. Goergen, Geochem. 23 December 2015; accepted 31 August 2016
with depth due to increasing velocity strengthen- Geophys. Geosyst. 14, 2247–2264 (2013). 10.1126/science.aaf1370
ing of the creeping matrix or decreasing stress
drop within the asperities. Another possibility is
a geometrical (or structural) control: At larger
depths, the range of asperity sizes (and, hence, SOLAR CELLS
of Re) may be narrower or D may be larger (e.g.,
due to lithological variations).

RE FE RENCES AND N OT ES Quantum dot–induced phase


stabilization of a-CsPbI3 perovskite
1. A. Maggi, J. Jackson, D. McKenzie, K. Priestley, Geology 28,
495–498 (2000).
2. R. Bürgmann, G. Dresen, Annu. Rev. Earth Planet. Sci. 36,
531–567 (2008).
3. C. Thurber et al., Bull. Seismol. Soc. Am. 96, S38–S49 (2006).
4. M. Kahraman et al., Earth Planet. Sci. Lett. 430, 129–139 (2015).
for high-efficiency photovoltaics
5. J. H. Shaw et al., Earth Planet. Sci. Lett. 415, 1–15 (2015).
6. R. J. Norris, A. F. Cooper, J. Struct. Geol. 25, 2141–2157 Abhishek Swarnkar,1,2 Ashley R. Marshall,1,3 Erin M. Sanehira,1,4
(2003).
7. J. C. White, J. Struct. Geol. 38, 11–20 (2012).
Boris D. Chernomordik,1 David T. Moore,1 Jeffrey A. Christians,1
8. J. R. Rice, J. Geophys. Res. 111, B05311 (2006). Tamoghna Chakrabarti,5 Joseph M. Luther1*
9. A. Inbal, R. Clayton, J.-P. Ampuero, Geophys. Res. Lett. 42,
6314–6323 (2015). We show nanoscale phase stabilization of CsPbI3 quantum dots (QDs) to low temperatures that
10. A. S. Bryant, L. M. Jones, J. Geophys. Res. 97, 437–447 (1992).
11. H. Magistrale, Geophys. Res. Lett. 29, 87-1–87-4 (2002).
can be used as the active component of efficient optoelectronic devices. CsPbI3 is an
12. E. Hauksson, Geophys. J. Int. 186, 82–98 (2011). all-inorganic analog to the hybrid organic cation halide perovskites, but the cubic phase of bulk
13. L. C. Price, M. J. Pawlewicz, T. A. Daws, “Organic CsPbI3 (a-CsPbI3)—the variant with desirable band gap—is only stable at high temperatures.
metamorphism in the California petroleum basins: Chapter A, We describe the formation of a-CsPbI3 QD films that are phase-stable for months in ambient air.
Rock-Eval and vitrinite reflectance” (U.S. Geological Survey
The films exhibit long-range electronic transport and were used to fabricate colloidal perovskite
Bulletin 2174-A, 1999).
14. E. Hauksson, W. Yang, P. M. Shearer, Bull. Seismol. Soc. Am.
QD photovoltaic cells with an open-circuit voltage of 1.23 volts and efficiency of 10.77%.
102, 2239–2244 (2012). These devices also function as light-emitting diodes with low turn-on voltage and
15. L. B. Grant, J. T. Waggoner, T. K. Rockwell, C. vonStein, tunable emission.
Bull. Seismol. Soc. Am. 87, 277–293 (1997).

H
16. D. L. Kohlstedt, B. Evans, S. J. Mackwell, J. Geophys. Res. 100,
17587–17602 (1995). ybrid organic-inorganic halide perovskites, all-inorganic Pb-halide perovskite with the most
17. G. Hirth, N. M. Beeler, Geology 43, 223–226 (2015). with the common formulation ABX3 (where appropriate band gap Eg for PV applications is
18. E. O. Lindsey, Y. Fialko, J. Geophys. Res. 118, 689–697 A is an organic cation, B is commonly Pb2+, cubic (a) CsPbI3 (Eg = 1.73 eV) because geomet-
(2013).
and X is a halide), were first applied to rical constraints of the perovskite structure re-
19. J. Gazdag, Geophys. 43, 1342–1351 (1978).
20. IRIS, Incorporated Research Institutions for Seismology; photovoltaics (PVs) as methylammonium quire a large +1 A-site cation, and Cs+ is the most
www.iris.edu/hq. lead triiodide (CH3NH3PbI3) in 2009 (1). Perov- feasible. However, below 320°C, the orthorhombic
21. Materials and methods are available as supplementary skite PV devices processed from solution inks now (d) phase (Eg = 2.82 eV) is thermodynamically
materials on Science Online. convert >22% of incident sunlight into electricity,
22. J. Crouch, J. Suppe, Geol. Soc. Am. Bull. 105, 1415–1434 (1993).
which is on par with the best thin-film chalco-
23. T. L. Wright, ”Structural geology and tectonic evolution of 1
Chemical and Materials Science, National Renewable Energy
the Los Angeles Basin, California,“ in American Association of genide and silicon devices, but durability of the Laboratory (NREL), Golden, CO 80401, USA. 2Department of
Petroleum Geologists Memoir 52, K. T. Biddle, Ed. (1991), semiconductor presents a major technical hurdle Chemistry, Indian Institute of Science Education and Research
pp. 35–134. to commercialization. Under environmental stress, (IISER), Pune 411008, India. 3Department of Chemistry and
24. B. M. Kennedy et al., Science 278, 1278–1281 (1997). CH3NH3PbI3 dissociates into PbI2 and CH3NH3I, Biochemistry, University of Colorado, Boulder, CO 80309, USA.
4
25. J. R. Boles, G. Garven, H. Camacho, J. E. Lupton, Department of Electrical Engineering, University of Washington,
Geochem. Geophys. Geosyst. 16, 2364–2381 (2015).
the latter of which is volatile (2). Seattle, WA 98195, USA. 5Metallurgical and Materials
26. V. Lekic, S. W. French, K. M. Fischer, Science 334, 783–787 Thus, an all-inorganic structure without a vol- Engineering, Colorado School of Mines, Golden, CO 80401, USA.
(2011). atile organic component is highly desired. The *Corresponding author. Email: joey.luther@nrel.gov

92 7 OCTOBER 2016 • VOL 354 ISSUE 6308 sciencemag.org SCIENCE


RE S EAR CH | R E P O R T S

preferred (3). Nevertheless, groups have explored


CsPbX3 compounds as PV materials, but films of
a-CsPbI3 undergo immediate transformation to
the orthorhombic phase when exposed to ambi-
ent conditions (4). Attempts to stabilize the cubic
phase through alloying with Br– have been explored
because CsPbIBr2 shows a much reduced d to a
phase transition temperature of 100°C (3). How-
ever, the composition change leads to an undesired
increase in the band gap. We show that nano-
crystal surfaces can be used to stabilize a-CsPbI3
at room temperature, far below the phase transition
temperature for thin film or bulk materials. We
further show that we can control the electronic
coupling of quantum dots (QDs) to produce air-
stable, efficient PV cells (initial efficiency above
10%) based on this all-inorganic material.
Many physical properties differ between
nanometer-sized and bulk crystalline materials

Downloaded from http://science.sciencemag.org/ on October 6, 2016


of the same chemical compound. One such exam-
ple is the structural phase in which the constituent
atoms are arranged. For example, the semicon-
ductors CdS and CdSe embody a rock salt struc-
ture at high pressure. However, the solid-solid
phase transition point between the rock salt phase
and the hexagonal wurtzite phase can vary greatly
in temperature and pressure as a function of crystal
size (5, 6). Manipulated size-dependent phase
diagrams have been explored in a variety of mate-
rial systems, with advantageous properties of the
crystals emerging at reduced dimensions in oxides
(such as TiO2), lanthanides (such as NaYF4) (7),
metals (such as Ag) (8), and ferroelectrics (such as
the perovskite BaTiO3) (9).
Synthetic protocols of colloidal halide perov-
skite QDs have recently been reported (10–17). Fig. 1. Characterization of CsPbI3 QDs. (A) Normalized UV-visible absorption spectra and photographs
CsPbX3 QDs exhibit improved room-temperature of CsPbI3 QDs synthesized at (a) 60°C (3.4 nm), (b) 100°C (4.5 nm), (c) 130°C (5 nm), (d) 150°C (6.8 nm), (e)
cubic-phase stability and attractive optical proper- 170°C (8 nm), (f) 180°C (9 nm), and (g) 185°C (12.5 nm).The numbers in parentheses are the average size
ties for a wide range of applications (11, 18–22). from TEM. (B) Normalized photoluminescence spectra and photographs under UV illumination of the QDs
Experiments on size- and shape-dependent op- from (A). (C) High-resolution TEM of CsPbI3 QDs synthesized at 180°C. (D) XRD patterns of QDs syn-
tical properties (11, 23–25), surface chemistry (26), thesized at (from bottom to top) 60°, 100°, 170°, 180°, and 185°C, confirming that they crystallize in the cubic
and other photophysics (27) are being explored for phase of CsPbI3.
CsPbBr3 QDs. However, previous studies were un-
able to achieve a-CsPbI3 QDs that were stable
enough for extensive characterization or to be
used in PV cells.
We present an improved synthetic route and
purification approach of CsPbI3 QDs. Once puri-
fied, the QDs retain the cubic phase for months
in ambient air and even at cryogenic temperatures.
A method for perovskite QD film assembly is de-
scribed that allows for efficient dot-to-dot elec-
tronic transport while retaining the phase stability
of the individual QDs. The PV cells produced from
this approach have the highest power conversion
efficiency (PCE) and stabilized power output (SPO)
of any all-inorganic perovskite absorber, produce
1.23 V at open circuit (among the best of any
perovskite PV cells), and also function as light-
emitting diodes (LEDs), emitting visible red light
with low turn-on voltage.
The tunability of the band gap via size control
due to quantum confinement is shown in Fig. 1.
The series of CsPbI3 QDs, with varied size (band Fig. 2. Phase stability of CsPbI3 QDs. (A) Powder XRD patterns and (B) UV-visible absorption spectra,
gap), were synthesized with the addition of Cs- normalized at 370 nm, of CsPbI3 QDs synthesized at 170°C and stored in ambient conditions for a period of
oleate to a flask containing PbI2 precursor, as first 60 days. (Inset) The slight blue shift that is seen in the excitonic peak with extended storage. (C) Rietveld
described by Protesescu et al. (11)—here, using refinement fitting of CsPbI3 QD XRD pattern, revealing pure cubic-phase CsPbI3.

SCIENCE sciencemag.org 7 OCTOBER 2016 • VOL 354 ISSUE 6308 93


R ES E A RC H | R E PO R TS

ligand (a hard acid) in the case of CsPbI3, compared


Solution with that of CsPbBr3 (30, 32). Therefore, the iso-

Transmission (a.u.)
Film
lation of CsPbI3 QDs is more difficult than that of
Abs./PL (norm.)

100 °C CsPbBr3 QDs because of the loss of ligand during


extraction, causing agglomeration and conversion
150 °C to the orthorhombic phase. Thus, we found that
MeOAc Washed
MeOAc, which isolates the QDs without full re-
180 °C As-Cast Film moval of the surface species, is critical to the phase-
stable devices described below.
550 600 650 700 750 4000 3000 2000 1000 The high-resolution transmission electron mi-
Wavelength (nm) Wavenumber (cm-1) crograph (TEM) of the sample synthesized at 180°C
(Fig. 1C) shows an interplanar distance of 0.62 nm,
Fig. 3. CsPbI3 QD films. (A) UV-visible absorption (solid lines) and PL spectra (dashed lines) of CsPbI3
which is consistent with the (100) plane of cubic
QDs in solution (blue) and as-cast films (black) for QDs synthesized at 100°, 150°, and 180°C. (B) FTIR
phase CsPbI3 (24, 31, 33). In Fig. 2, A and B, powder
spectra showing the IR transmission of a CsPbI3 QD film as cast (black) and after treating with MeOAc (red).
x-ray diffraction (XRD) patterns and UV-visible ab-
sorption spectra confirm the absence of diffraction
injection temperatures between 60° and 185°C to photograph of the QDs in hexane. Upon ultra- peaks or the high-energy (~3 eV) sharp absorption
control the size (28). This produces QDs solubi- violet (UV) excitation, emission was in the orange characteristic of orthorhombic phase formation
lized by noncrystalline iodide and oleylammonium (600 nm) to red (680 nm) color range, corres- (31), even after 60 days of storage in ambient con-

Downloaded from http://science.sciencemag.org/ on October 6, 2016


surface ligands (26). Unpurified QDs transform to ponding to a band gap between 2.07 and 1.82 eV ditions. Additionally, the QDs remained in the
the orthorhombic phase within several days (fig. (photographs showing PL from dried QD powders cubic phase even after the solution was cooled to
S1) (28), as in previous reports (29, 30). However, are shown in fig. S2) (28). The full width at half- 77 K, further demonstrating the expanded tempe-
we developed a process to purify the QDs by using maximum of the PL for the smallest QDs was rature stability of the cubic phase.
methyl acetate (MeOAc), an antisolvent that re- 83 meV and increased slightly for the larger sizes, Rietveld refinement of the XRD patterns (Fig.
moves excess unreacted precursors without in- whereas the PL quantum yield varied from 21 to 2C) (28) allowed us to quantify the contribution
ducing agglomeration. Using this extraction 55% for different sizes (fig. S3) (28). from cubic and orthorhombic phases. No detect-
procedure, we found that the QDs are stable in In contrast to the instability of the cubic phase able orthorhombic phase was found. Addition-
the cubic phase for months with ambient storage. of bulk CsPbI3 at room temperature, QDs have ally, lattice parameters of three different size CsPbI3
The excitonic peak of CsPbI3 shifted between been reported to retain the cubic phase because QD samples were estimated (Table 1). The lattice
585 and 670 nm, corresponding to QD sizes be- of the large contribution of surface energy (Fig. parameter values showed a size dependence and
tween 3 and 12.5 nm, respectively. The correspond- 1D) (11, 31). The softer basic nature of I– as com- were lower than the previously measured exper-
ing normalized photoluminescence (PL) spectra pared with Br– results in weaker acid-base interac- imental value (6.2894 Å at 634 K) of bulk cubic
of the samples are shown in Fig. 1B, along with a tions between the halide and the oleylammonium CsPbI3 (33). Our measurements were performed

Fig. 4. CsPbI3 optoelectronic devices. (A) Schematic (with TEM image of QDs) and (B) SEM cross-section of the CsPbI3 PV cell. (C) Current density–voltage
curves of a device measured in air over the course of 15 days. The black diamond represents the stabilized power output of the device at 0.92 V, as shown in fig. S9.
(D) External quantum efficiency (black, left ordinate) and integrated current density (blue, right ordinate) of the device. (E) EL spectra of CsPbI3 PV cell (CsPbI3 QDs
synthesized at 170°C) under forward bias. (Inset) A photograph of the luminescent device. (F) PL (dashed lines) and EL (solid lines) spectra of completed devices
fabricated by using CsPbI3 QDs synthesized at 170° and 180°C, demonstrating size quantization effects in the completed devices.

94 7 OCTOBER 2016 • VOL 354 ISSUE 6308 sciencemag.org SCIENCE


RE S EAR CH | R E P O R T S

REFERENCES AND NOTES


Table 1. Results of the Rietveld refinement. a, lattice parameter; Rwp, weighted-profile R factor. 1. A. Kojima, K. Teshima, Y. Shirai, T. Miyasaka, J. Am. Chem.
Soc. 131, 6050–6051 (2009).
2. D. P. Nenon et al., Energy Environ. Sci. 9, 2072–2082 (2016).
3. S. Sharma, N. Weiden, A. Weiss, Z. Phys. Chem. 175, 63–80 (1992).
QD size (TEM) QD size (Rietveld) a (Å) Rwp
4. R. E. Beal et al., J. Phys. Chem. Lett. 7, 746–751 (2016).
8 nm 9 ± 1 nm 6.231 ± 0.002 3.42 5. A. P. Alivisatos, Science 271, 933–937 (1996).
.....................................................................................................................................................................................................................
6. S. H. Tolbert, A. P. Alivisatos, Science 265, 373–376 (1994).
9 nm 10 ± 1 nm 6.220 ± 0.002 6.50
..................................................................................................................................................................................................................... 7. F. Wang et al., Nature 463, 1061–1065 (2010).
15.5 nm 17 ± 2 nm 6.189 ± 0.002 7.79
.....................................................................................................................................................................................................................
8. C. C. Yang, S. Li, J. Phys. Chem. C 112, 16400–16404 (2008).
9. S. Schlag, H. F. Eicke, Solid State Commun. 91, 883–887 (1994).
10. L. C. Schmidt et al., J. Am. Chem. Soc. 136, 850–853 (2014).
11. L. Protesescu et al., Nano Lett. 15, 3692–3696 (2015).
12. F. Zhang et al., ACS Nano 9, 4533–4542 (2015).
13. D. Zhang, S. W. Eaton, Y. Yu, L. Dou, P. Yang, J. Am. Chem.
at 297 K, whereas high temperatures are required device architecture is shown in Fig. 4A, and a
Soc. 137, 9230–9233 (2015).
to characterize bulk cubic CsPbI3. A similar in- scanning electron micrograph (SEM) cross-section 14. H. Huang et al., ACS Appl. Mater. Interfaces 7, 28128–28133 (2015).
crease in lattice parameter with decreasing particle image of the reported device with 9 nm QDs is 15. Y. Hassan et al., Adv. Mater. 28, 566–573 (2016).
size has been reported in other systems and at- shown in Fig. 4B. The reverse-scan current density- 16. S. Sun, D. Yuan, Y. Xu, A. Wang, Z. Deng, ACS Nano 10,
tributed to electrostatic relaxation with decreasing voltage (JV) curves showed an open-circuit voltage 3648–3657 (2016).
17. T. C. Jellicoe et al., J. Am. Chem. Soc. 138, 2941–2944 (2016).
crystal size (34). (VOC) of 1.23 V, and 10.77% PCE for a 0.10 cm2 cell 18. Y. Wang et al., Adv. Mater. 27, 7101–7108 (2015).
In order to use these highly phase-stable made and tested completely in ambient conditions 19. S. Yakunin et al., Nat. Commun. 6, 8056 (2015).

Downloaded from http://science.sciencemag.org/ on October 6, 2016


a-CsPbI3 QDs in optoelectronic devices, we devel- (relative humidity ~15 to 25%) (Fig. 4C). The hys- 20. Y. Xu et al., J. Am. Chem. Soc. 138, 3761–3768 (2016).
oped a method to cast electronically conductive teresis along with SPO of a device scanned at 21. A. Swarnkar et al., Angew. Chem. Int. Ed. Engl. 54,
15424–15428 (2015).
QD films. The QDs were first spin-cast from octane various sweep rates is shown in fig. S7 (28). Fur- 22. F. Hu et al., ACS Nano 9, 12410–12416 (2015).
then dipped in a saturated MeOAc solution of thermore, the PCE improved from its initial value 23. Q. A. Akkerman et al., J. Am. Chem. Soc. 138, 1010–1016 (2016).
either Pb(OAc)2 or Pb(NO3)2 (neat MeOAc was over the course of 60 days storage in dry but am- 24. N. S. Makarov et al., Nano Lett. 16, 2349–2362 (2016).
used as a control). This process was repeated bient conditions (fig. S8) (28). In fig. S9 (28), we 25. Y. Bekenstein, B. A. Koscher, S. W. Eaton, P. Yang,
A. P. Alivisatos, J. Am. Chem. Soc. 137, 16008–16011 (2015).
multiple times—typically, three to five—to produce show the SPO of the cell by measuring the current 26. J. De Roo et al., ACS Nano 10, 2071–2081 (2016).
QD films with thicknesses between 100 and 400 nm. density while the device is biased at 0.92 V. In Fig. 27. Y. S. Park, S. Guo, N. S. Makarov, V. I. Klimov, ACS Nano 9,
The optical absorption and PL spectra (Fig. 3A, 4D, the spectral response of the PV cell is shown, 10386–10393 (2015).
for three samples with indicated reaction tem- indicating a band gap of 1.75 eV for this film. We 28. Materials and methods are available as supplementary
materials on Science Online.
perature) show that in each case, the film absorb- compare QD devices to thin-film CsPbX3 perov- 29. C. C. Lin, A. Meijerink, R.-S. Liu, J. Phys. Chem. Lett. 7,
ance and PL was red-shifted ~20 nm from that skite solar cells following literature reports, which 495–503 (2016).
of the QDs in solution, whereas the tunable emis- have thus far reported at 9.8% PCE and SPO as 30. Q. A. Akkerman et al., J. Am. Chem. Soc. 137, 10276–10281
sion properties of the films indicate that quantum high as 6.5% (4, 31, 36). The QD devices show (2015).
31. G. E. Eperon et al., J. Mater. Chem. A 3, 19688–19695 (2015).
confinement is preserved. Fourier-transform infra- improved JV-scan efficiency, operational stability, 32. R. G. Pearson, J. Am. Chem. Soc. 85, 3533–3539 (1963).
red (FTIR) spectra show the removal of organic and tolerance to higher relative humidity levels 33. D. M. Trots, S. V. Myagkota, J. Phys. Chem. Solids 69,
ligands from the film with exposure to neat MeOAc (figs. S10 and S11 and table S3) (28). The VOC is 2520–2526 (2008).
(Fig. 3B), given the near absence of C–H modes remarkably higher than that of other QD solar 34. S. Tsunekawa, K. Ishikawa, Z. Li, Y. Kawazoe, A. Kasuya, Phys.
Rev. Lett. 85, 3440–3443 (2000).
near 3000 cm−1 or below ~2000 cm−1 belonging to cells (typically <0.7 V) and among the highest VOC 35. S. Dastidar et al., Nano Lett. 16, 3563–3570 (2016).
oleylammonium, oleate, or octadecene. We there- in all perovskite PV cells for band gap values 36. R. J. Sutton et al., Adv. Energy Mater. 6, 1502458 (2016).
fore attribute the preserved phase stability of the below 2 eV (fig. S12, stabilized VOC) (28). We 37. G. R. Yettapu et al., Nano Lett. 16, 4838–4848 (2016).

QDs in the films to the size of the crystals (given the have not optimized the device architecture or AC KNOWLED GME NTS
quantum confined optical properties) independent the QD film-treatment scheme. We found that We thank J. van de Lagemaat, W. Tumas, H. Choi, M. Beard, and
of the surface species. However, we found that pro- dip-coating spin-cast films in neat MeOAc and J. Berry for helpful discussions and B. To for SEM imaging. We
longed annealing at temperatures >200°C causes MeOAc saturated with Pb(OAc)2 or Pb(NO3)2 all acknowledge support from the Center for Advanced Solar
further grain growth and thus induces a phase work reasonably well (JV-scanned PCE > 9%) in Photophysics, an Energy Frontier Research Center funded by the
U.S. Department of Energy, Office of Science, Office of Basic
transition to the orthorhombic phase (fig. S4 and PV devices. Large diffusion lengths and mobility Energy Sciences for quantum dot coupling and solar cell
table S1) (28). Additional strategies to preserve values have been measured in CsPbBr3 QDs by structures. Device durability and structural phase characterization
the phase in sintered QD films are being explored means of terahertz spectroscopy (37); however, a was performed within the hybrid perovskite solar cell program of
(35). We have observed cubic-phase CsPbI3 with better understanding of the electronic coupling the National Center for Photovoltaics funded by the U.S.
Department of Energy, Office of Energy Efficiency and Renewable
edge length up to 50 nm using the solution-phase is critical to maximizing long-range transport in Energy, Solar Energy Technologies Office under contract DE-AC36-
synthesis described here. QD perovskite films. 08GO28308DOE. The original conception and QD synthesis was
We also probed the interaction of Pb2+ salts Given the PL properties of these perovskite QDs, performed under the Laboratory Directed Research and
with QDs in solution and on films by monitoring we explored their use as LEDs. The PV devices Development program at NREL. A.S. acknowledges the Bhaskara
Advanced Solar Energy fellowship funded by the Department of
the fluorescence (fig. S5) (28). Titration of a small produced bright visible electroluminescence (EL) Science and Technology, government of India, and Indo-U.S.
amount of Pb(OAc)2 dissolved in MeOAc to the when biased above VOC (Fig. 4E, inset). The EL had Science and Technology Forum (IUSSTF). E.M.S. acknowledges a
QD solution showed an enhancement in PL, sug- a low turn-on voltage near the band gap of the NASA Space Technology Research Fellowship. D.T.M. acknowledges
gesting improved surface passivation. The sur- CsPbI3, with increasing intensity at larger applied the NREL Director’s Fellowship. All data in the paper and
supplementary materials are available. An application has been made
face treatments increase the PL lifetime over that biases (Fig. 4E). These spectra provide direct evi- for a provisional patent (U.S. patent application no. 62/343,251).
of neat QD films, which highlights the importance dence that quantum confinement is retained in
of surface chemistry in this QD system (fig. S6 and the complete devices, which is critical to retain- SUPPLEMENTARY MATERIALS
table S2) (28). Titrations with only MeOAc caused ing the improved cubic-phase stability, as seen by www.sciencemag.org/content/354/6308/92/suppl/DC1
fast PL quenching. Similarly, dip-coating of the the shift in both the EL and PL spectra of devices Materials and Methods
Supplementary Text
QD film in a saturated solution of Pb(OAc)2 in with different-size QDs (Fig. 4F). The synthesis of Figs. S1 to S12
MeOAc resulted in a PL enhancement of ~350% normally unstable material phases stabilized Tables S1 to S3
compared with dip-coating in MeOAc alone. through colloidal QD synthesis provides another References (38, 39)
We fabricated PV cells with CsPbI3 QD films mechanism for material design for PVs, LEDs, 1 June 2016; accepted 7 September 2016
as the photoactive material. A schematic of the and other applications. 10.1126/science.aag2700

SCIENCE sciencemag.org 7 OCTOBER 2016 • VOL 354 ISSUE 6308 95


Quantum dot−induced phase stabilization of α-CsPbI3
perovskite for high-efficiency photovoltaics
Abhishek Swarnkar, Ashley R. Marshall, Erin M. Sanehira, Boris D.
Chernomordik, David T. Moore, Jeffrey A. Christians, Tamoghna
Chakrabarti and Joseph M. Luther (October 6, 2016)
Science 354 (6308), 92-95. [doi: 10.1126/science.aag2700]

Editor's Summary

Downloaded from http://science.sciencemag.org/ on October 6, 2016


Maintaining a stable phase
For solar cell applications, all-inorganic perovskite phases could be more stable than those
containing organic cations. But the band gaps of the former, which determine the electrical conductivity
of these materials, are not well matched to the solar spectrum. The cubic structure of CsPbI 3 is an
exception, but it is stable in bulk only at high temperatures. Swarnkar et al. show that surfactant-coated
α-CsPbI3 quantum dots are stable at ambient conditions and have tunable band gaps in the visible
range. Thin films of these materials can be made by spin coating with an antisolvent technique to
minimize surfactant loss. When used in solar cells, these films have efficiencies exceeding 10%, making
them promising for light harvesting or for LEDs.
Science, this issue p. 92

This copy is for your personal, non-commercial use only.

Article Tools Visit the online version of this article to access the personalization and
article tools:
http://science.sciencemag.org/content/354/6308/92

Permissions Obtain information about reproducing this article:


http://www.sciencemag.org/about/permissions.dtl

Science (print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week
in December, by the American Association for the Advancement of Science, 1200 New York
Avenue NW, Washington, DC 20005. Copyright 2016 by the American Association for the
Advancement of Science; all rights reserved. The title Science is a registered trademark of AAAS.

You might also like