2014-Green-The Emergence of Perovskite Solar Cells

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/280388277

The emergence of perovskite solar cells

Article  in  Nature Photonics · July 2014


DOI: 10.1038/NPHOTON.2014.134

CITATIONS READS

4,936 25,199

3 authors, including:

Martin Green Anita Ho-Baillie


UNSW Sydney The University of Sydney
942 PUBLICATIONS   76,110 CITATIONS    215 PUBLICATIONS   18,777 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Development of light-trapping schemes in evaporated laser-crystallised silicon thin-film solar cells on glass superstrates View project

Si-based Solar Cell related stuff View project

All content following this page was uploaded by Martin Green on 26 July 2015.

The user has requested enhancement of the downloaded file.


REVIEW ARTICLE
PUBLISHED ONLINE: XX JULY 2014 | DOI: 10.1038/NPHOTON.2014.134

The emergence of perovskite solar cells


Martin A. Green1*, Anita Ho-Baillie1 and Henry J. Snaith2
The past two years have seen the unprecedentedly rapid emergence of a new class of solar cell based on mixed organic–inor-
ganic halide perovskites. Although the first efficient solid-state perovskite cells were reported only in mid-2012, extremely
rapid progress was made during 2013 with energy conversion efficiencies reaching a confirmed 16.2% at the end of the year.
This increased to a confirmed efficiency of 17.9% in early 2014, with unconfirmed values as high as 19.3% claimed. Moreover,
a broad range of different fabrication approaches and device concepts is represented among the highest performing devices —
this diversity suggests that performance is still far from fully optimized. This Review briefly outlines notable achievements to
date, describes the unique attributes of these perovskites leading to their rapid emergence and discusses challenges facing the
successful development and commercialization of perovskite solar cells.

R
ecent substantial reductions in the manufacturing costs of ethylammonium (CH3CH2NH3+) (RA  =  0.23  nm)9 and forma-
mainstream silicon solar cell technology assure the future midinium (NH2CH=NH2+) (RA is estimated to lie in the range
large-scale use of photovoltaics, with a recent forecast antici- 0.19–0.22  nm) also give good results10–13. The anion X is a halo-
pating photovoltaics will contribute nearly a third of new electric- gen, generally iodine (RX = 0.220 nm), although Br and Cl are also
ity generation capacity worldwide between now and 20301. As in commonly used (RX = 0.196 nm and 0.181 nm), usually in a mixed
microelectronics, silicon has a combination of strengths that has halide material. For efficient cells, cation B has universally been Pb
made it difficult to displace as the favoured photovoltaic material. (RB = 0.119 nm); Sn (RB = 0.110 nm) forms similar compounds with
Opportunities exist for technologies that promise either signifi- lower, theoretically more ideal bandgaps14, but generally lower sta-
cantly higher energy conversion efficiencies or significantly lower bility (attributed to the ease of oxidation of Sn to SnI4 in the iodide
processing costs. A new generation of mixed organic–inorganic hal- perovskite; relativistic effects in the Pb counterpart are thought to
ide perovskites offers tantalizing prospects on both fronts2–6. provide greater protection against oxidation14). The archetypal com-
Some key attributes of these perovskites include ease of fabrica- pound is thus methylammonium lead triiodide (CH3NH3PbI3), with
tion, strong solar absorption and low non-radiative carrier recom- mixed halides CH3NH3PbI3−xClx and CH3NH3PbI3−xBrx also being
bination rates for such simply prepared materials, plus the ability to important. Calculated and estimated t and μ factors for a range of
capitalize on over 20 years of development of related dye-sensitized these perovskites are shown in Fig. 1b.
and organic photovoltaic cells. A reasonably high carrier mobil-
ity is an important property for some cell architectures, as is the Notable achievements to date
range of properties accessible by forming mixed compounds within Interest in organic–inorganic halide perovskites dates back over
a compatible materials system. One negative aspect of perovskites a century 15, although present results benefit from recent investi-
is the fact that lead has been a major constituent of all highly per- gations of related thin-film transistors and light-emitting diodes
forming perovskite cells to date, raising toxicity issues during device (LEDs) by Mitzi and co-workers16,17. Although these LEDs doubt-
fabrication, deployment and disposal. Also, they generally undergo less exhibited photovoltaic properties and their use in solar
degradation (sometimes quite rapid) on exposure to moisture and cells was anticipated17, this was not studied at the time, partly
ultraviolet radiation. because of Pb toxicity and the fact that more benign Sn materials
Perovskites are materials described by the formula ABX3, where were not considered to be sufficiently robust (D. Mitzi, personal
X is an anion and A and B are cations of different sizes (A  being communication).
larger than B). The crystal structure of perovskites is depicted in Miyasaka and co-workers18–20 were apparently the first to report
Fig. 1a. Their crystallographic stability and probable structure can photovoltaic results for perovskites; they were attracted by the self-
be deduced by considering a tolerance factor t and an octahedral organization potential of perovskite in the nanoporous TiO2 layer of
factor μ (ref. 7); here, t is defined as the ratio of the distance A-X to dye-sensitized cells. In 2006, they reported CH3NH3PbBr3 cells with
the distance B-X in an idealized solid-sphere model (t = (RA + RX)/ an efficiency of 2.2% (ref.  18); by replacing bromine with iodine
{√2(RB + RX)}, where RA, RB and RX are the ionic radii of the corre- they were able to increase the efficiency to 3.8% in 200919 (although
sponding ions) and μ is defined as the ratio RB/RX. For halide perovs- all devices were unstable). An organic electrolyte containing lithium
kites (X = F, Cl, Br, I)7, generally 0.81 < t < 1.11 and 0.44 < μ < 0.90. halide and the corresponding halogen formed the hole-transporting
If t lies in the narrower range 0.89–1.0, the cubic structure of Fig. 1a medium (HTM), allowing positive contact. Attempts to replace this
is likely, with lower t values giving less symmetric tetragonal or HTM with a solid-state HTM were not particularly successful20.
orthorhombic structures. Despite these constraints, transitions Subsequently, Park and colleagues21 employed similar structures
between such structures on heating are common for any given per- in which perovskite was shown to be deposited as sparsely spaced
ovskite, with the high-temperature phase generally being cubic. hemispherical nanoparticles that were approximately 2.5  nm in
For the organic–inorganic halide perovskites of present inter- diameter. By applying TiO2 surface treatment prior to deposition,
est, the larger cation A is organic; it is generally methylammo- they realized an efficiency of 6.5% in 201121. Perovskite nanoparti-
nium (CH3NH3+) with RA  =  0.18  nm (ref.  8), although related cles exhibited better absorption than standard N719 dye sensitizers,

1
Australian Centre for Advanced Photovoltaics (ACAP), School of Photovoltaic and Renewable Energy Engineering, University of New South Wales, Sydney
2052, Australia. 2University of Oxford, Clarendon Laboratory, Parks Road, Oxford OX1 3PU, UK. *e-mail: [email protected]

506 NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics


© 2014 Macmillan Publishers Limited. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.134 REVIEW ARTICLE
a b 0.90
0.80

0.70

0.60

Octahedral factor, μ
A
0.50

B 0.40

X 0.30 MAPbl3 MAPbBr3 MAPbCl3


MASnl3 MASnBr3 MASnCl3
0.20
EAPbl3 EAPbBr3 EAPbCl3
0.10 EASnl3 EASnBr3 EASnCl3
0.00
0.75 0.80 0.85 0.90 0.95 1.00 1.05
Tolerance factor, t

Figure 1 | Perovskite crystal structure and associated tolerance and octahedral factors. a, Cubic perovskite crystal structure. For photovoltaically
interesting perovskites, the large cation A is usually the methylammonium ion (CH3NH3), the small cation B is Pb and the anion X is a halogen ion
(usually I, but both Cl and Br are also of interest). For CH3NH3PbI3, the cubic phase forms only at temperatures above 330 K due to a low t factor (0.83).
b, Calculated t and μ factors for 12 halide perovskites. The corresponding formamidinium (NH2CH=NH2) based halides are expected to have intermediate
values between those of the methylammonium (MA) and ethylammonium (EA; CH3CH2NH3) compounds shown.

but they dissolved in the electrolyte, resulting in a rapid degradation higher t factor due to the smaller ionic radius of Br (molecules in
of performance21. position A exclude full cubic symmetry 29).
This stimulated the replacement of problematic electrolytes by a Further progress was reported at the European Materials
solid-state HTM22,23. Park, Grätzel and colleagues22 introduced a spiro- Research Symposium in May 2013, with two groups reporting
MeOTAD (2,2’,7,7’-tetrakis(N,N-di-p-methoxyphenylamine)-9,9’- efficiencies above 15%. Grätzel’s group used TiO2 scaffolding and
spirobifluorene) HTM, which was developed for organic LEDs24 two-step iodide deposition, which improved the morphology 30.
but was also found to be effective in solid-state dye cells25. When They also reported the first independent measurements of effi-
dissolved in an organic solvent, spiro-MeOTAD penetrates nanopo- ciency by an independent accredited test centre, which confirmed
rous TiO2, leaving only solute molecules after solvent evaporation. an efficiency of 14.1%. (Independent measurements were found
Spiro-MeOTAD not only improved the stability, as expected, it also to be an essential quality-control measure with other photovoltaic
boosted the reported efficiency to 9.7% (ref. 22). The cell structure technologies to prevent inflated results infiltrating the literature31;
is encompassed by the more general device of Fig. 2a if the optional inexperience or overenthusiasm often result in ‘in-house’ data being
continuous perovskite layer is removed, leaving only scaffolding overestimates.) Snaith’s group reported similar results using vastly
infiltrated by perovskite (and subsequently HTM). different planar cells that did not have scaffolding. The simpler
Almost simultaneously (mid-2012), Snaith and co-workers23 structure allowed CH3NH3PbI3−xClx deposition by two-source ther-
also reported success with spiro-MeOTAD along with four addi- mal evaporation, again giving a better morphology 32 and a reported
tional developments that split the field wide open. One of these efficiency of 15.4%.
developments was the use of the mixed-halide CH3NH3PbI3−xClx, No efficiency improvements were announced at the 2013 Fall
which exhibited better stability and carrier transport than its Materials Research Society (MRS) Meeting, although a special per-
pure iodide equivalent 23,26. A second involved going beyond ear- ovskite session attracted considerable attention. At the end of 2013,
lier nanoparticle structures by coating nanoporous TiO2 surfaces Seok’s group achieved an independently confirmed efficiency of
with a thin perovskite layer and thereby forming extremely thin 16.2% by using the mixed-halide CH3NH3PbI3−xBrx (10–15% Br)
absorber (ETA) cells. A third was replacing conducting nanopo- and a poly-triarylamine HTM (S.  I. Seok, personal communica-
rous TiO2 by a similar but non-conducting Al2O3 network. This tion). Both optional layers shown in Fig. 2a were included, with the
improved the open-circuit voltage (Voc), boosting the reported thickness ratio of perovskite-infiltrated TiO2 scaffolding relative to
efficiency to 10.9%; it also demonstrated that perovskites have a the continuous perovskite layer being the key to the improved effi-
broader potential than just being used as sensitizers, as they are ciency (S. I. Seok, personal communication). This was increased to
able to transport both electrons and holes between cell terminals. a confirmed efficiency of 17.9% in early 2014 (S. I. Seok, personal
The fourth development exploited such ambipolar transport by communication). An unconfirmed efficiency of 19.3% was reported
demonstrating simple planar cells with the scaffolding (Fig.  2a) at the 2014 Spring MRS Meeting 6 (some uncertainty surrounds this
completely eliminated. reported efficiency due to the absence of supporting information).
A jump to a reported efficiency of 12.0% came from the com- Significantly, the latter three results were obtained using device
bined efforts of Seok, Grätzel and colleagues using both optional structures that span all the possibilities inherent in Fig.  2a: one
layers shown in Fig. 2a, including a solid perovskite capping layer or both optional layers, three different mixed-halide perovskites
overlying the scaffolding 27 (nanoporous TiO2 infiltrated by per- and two different HTMs were used (although some ambiguities
ovskite). Of the HTMs they investigated (which included spiro- remain30,33). Fabrication simplicity combined with similarities with
MeOTAD), poly-triarylamine proved to be the best. Seok’s group dye-sensitized and organic photovoltaics has resulted in a rapid
further improved the performance to realize a reported efficiency increase in the number of researchers working in this field. Before
of 12.3% using similar structures and mixed-halide CH3NH3PbI3− 2013, only seven journal papers had been published that discussed
xBrx perovskites . A low Br content (<10%) gave the best initial photovoltaic devices based on halide perovskites, whereas by the end
28

efficiency due to a lower bandgap, but higher Br contents (>20%) of 2013, relevant publications were appearing at the rate of seven per
provided a better high-humidity stability. This was correlated with month. This spurt further increased the diversity of approaches that
a tetragonal to pseudo-cubic structural transition arising from a gave creditable performances.

NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics 507


© 2014 Macmillan Publishers Limited. All rights reserved.
REVIEW ARTICLE NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.134

Electron
Light energy Electrons
−2.1
−3.9
−4.0
Glass −4.4
Scaffold/
perovskite FTO −5.1
(optional)
Compact TiO2

Perovskite Perovskite
(optional) −5.2
−5.5
HTM

Au −7.2

Hole
energy
Au or Ag −8.1 Holes

Figure 2 | Perovskite cell structure and associated vacuum energy levels. a, General organic–inorganic halide solar cell, which includes two optional layers
that are not essential for high performance; an energy conversion efficiency of over 15% has been reported for devices that have both optional layers, with
only the scaffold layer infiltrated by the perovskite (and then by the HTM) and without scaffolding; the structure then corresponds to a simple planar thin-
film cell. b, Vacuum energy levels (in eV) for corresponding materials (CH3NH3PbI3 perovskite, conducting TiO2 scaffold).

Key recent results include demonstrations of an additional depo- the challenges in collecting photogenerated carriers. Absorption
sition process involving PbI2 deposition from solution with in situ measurements (Fig.  3a) are consistent with calculations show-
conversion to perovskite by a vapour-phase CH3NH3I reaction34. ing direct-bandgap properties for perovskites of interest 14,45. Two
Promising results have also been reported with perovskites based on strong, spin–orbit split, excitonic absorption thresholds are appar-
organic cations other than CH3NH3+. Cations with larger ionic radii, ent, as in direct-bandgap III–V semiconductors46. However, reverse
specifically ethylammonium (CH3CH2NH3+)10 and formamidinium ordering of band-edge states (specifically a p-like conduction
(NH2CH=NH2+)11–13, increase the t factor and push structures towards band)47 results in splitting in the conduction band, rather than in
the symmetrical cubic phase. Different proportions of organic cati- the valence band. Interestingly, reverse band-edge ordering also
ons, inorganic cations (Pb, Sn) and halide anions (I, Br, Cl) can be gives a bandgap that increases with increasing temperature for any
incorporated in mixed perovskites, allowing their properties to be given phase48,49, which is the opposite trend to that of tetrahedrally
fine-tuned; however, the use of mixed halides has dominated. coordinated semiconductors.
Recent work has also demonstrated the use of new HTMs and The strong excitonic absorption edge also means there is no
electron transport media (ETMs). Effective ETMs have been reported basis for the common practice of determining bandgaps using Tauc
in which the standard fluorine-doped tin oxide (FTO)/compact plots50. The absorption edge is determined by a broadened exciton
TiO2 combination is replaced by indium tin oxide as a transparent impulse response, as is the case for direct III–V semiconductors,
conducting oxide combined with a thin (25 nm) ZnO-nanoparticle with the unbroadened response described by Elliott’s theory 46,51.
layer 35; this gave a reported efficiency of 15.7% for planar cells on The relatively high exciton binding energy compared to those of
glass. Low-temperature processing also gave a creditable perfor- III–V semiconductors with a similar bandgap (37–50 meV has been
mance on flexible polyethylene terephthalate. Inorganic HTMs such reported for iodide in the low-temperature phase52 and 35–75 meV
as CuI36 and CuSCN37 also give reasonable results, as do organic- for the mixed chloride at room temperature49), not only lowers the
photovoltaic-derived organics for both ETMs and HTMs, specifi- absorption threshold, but also increases the strength of the above-
cally a (6,6)-phenyl C61-butyric acid methyl ester ETM combined bandgap absorption that generates unbound electron–hole pairs.
with a poly(2,3-dihydrothieno-1,4-dioxin)-poly(styrenesulphonate) Correspondingly, above-bandgap absorption is comparable to or
HTM38–40 (efficiencies up to 12%, refs 38,39). stronger than that in many direct-bandgap III–V semiconductors,
Such diversification increases potential applications. Flexible such as GaAs, although it is lower than that of some inorganic chal-
cells5 require low processing temperatures of less than 150 °C, rather cogenides (see Fig. 3a).
than 500  °C, which is typical for compact TiO2. Graphene nano- For devices with a continuous perovskite layer or an insulating
flakes in the normally compact TiO2 layer allow such processing 41. scaffold (Fig.  2a), transport across the perovskite is important for
An efficiency of 15.6% has been reported for a structure that includes device operation. The exciton binding energy is still sufficiently low
all optional layers and Al2O3 scaffolding (0.6% graphene/TiO2 by for photogeneration of both excitons and unbound electron–hole
weight)41, and an efficiency of 15.9% has been achieved by com- pairs, with thermal dissociation of excitons into free carriers (and
bining small TiO2 nanoparticles with a titanium diisopropoxide reassociation into excitons) expected. The respective photocurrent
bis(acetylacetonate) binder 42. In other work, the non-uniformity contributions ideally depend on the coupling between these popula-
produced by solution deposition is exploited to produce neutral- tions, as characterized by the intrinsic exciton dissociation time —
colour semi-transparent cells43. Instead of continuous perovskite, the time that would prevail if excitons were not mobile or could not
small invisible islands are grown, which should be less expensive be ionized at heterojunctions.
than laser approaches for fabricating semi-transparent amorphous- At the low-coupling extreme (dissociation times large compared
silicon cells44. to recombination times), the two populations will be independent
and relative concentrations will be determined by the spectral com-
Enabling attributes position of illumination. For sunlight, free electron–hole pair gen-
Strong optical absorption is the key to the outstanding performance eration is expected to be completely dominant, as primary exciton
of these perovskite cells, reducing both the required thickness and generation will occur only at wavelengths near band edges (at least

508 NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics


© 2014 Macmillan Publishers Limited. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.134 REVIEW ARTICLE
a 107 b 70
GaAs Real
106 Si 60
Absorption coefficient (cm−1)

105 InP

Relative permittivity
50
CIGS
104 40
CIS
103 CdTe 30

102 CdS 20
CH3NH3Pbl3
101 10
CH3NH3Pbl3−xClx Imaginary
0 0
200 400 600 800 1,000 1,200 1,400 5 6 7 8 9 10 11 12 13 14 15
Wavelength (nm) Log (frequency, Hz)

Figure 3 | Absorption coefficients and relative permittivity. a, Absorption coefficient of CH3NH3PbI3 (ref. 40) (these values are about 50% higher
those obtained by Xing et al.74) and CH3NH3PBI3−xClx (ref. 96) compared to other solar cell materials (various sources). b, Real and imaginary parts of the
CH3NH3PbI3 dielectric constant at 300 K as a function of frequency. Dipolar and ionic components successively disappear as the frequency increases.
Low-frequency values (circles) are from ref. 64, mid-frequency (90 GHz) values (squares) are from ref. 65 and optical-frequency value (triangle) is
from ref. 66. The solid lines are Debye relaxation fits to mid-range values (ref. 65). The dashed lines show possible transitional values in the far infrared
(qualitative only).

for continuous perovskite layers). To contribute to the photocurrent In conventional polycrystalline semiconductors, low non-
in this low-coupling limit, excitons would need to dissociate at an radiative recombination generally requires large grain sizes, a low
interface (that with the ETM or the HTM or both), as in organic grain-boundary activity and a low density of intragranular defects.
photovoltaic devices. The present perovskites give narrow X-ray diffraction peaks that
At the opposite extreme of high coupling (rapid dissociation), are consistent with both near micrometre grain sizes and a reason-
free electron–hole and exciton concentrations will equilibrate with ably low intergranular defect density 58. Large grain sizes might be
relative concentrations theoretically determined by a mass action expected from the high ‘effective’ homologous temperatures corre-
law rather than by the illumination (this law will take the form sponding to relatively low processing temperatures due to the low
deduced by Combescot 53,54, when corrected to account for excited imputed perovskite melting points (most decompose before melt-
exciton states and scattering states associated with electron–hole ing 29). For metals, a 0.2 increase in the homologous temperature
attraction55). Excitons are expected to flow in the same direction as increases the grain size by a factor of ten59. To reduce detrimental
electrons or holes (Fig. 2b), depending on which has the larger elec- grain boundary activity, specific processing steps are required for
trochemical gradient 56. Because electron and hole currents increase polycrystalline Si, CdTe and copper–indium–gallium diselenide
in opposite directions (Fig.  2b), these gradients will probably be (CIGS) cells. These perovskites do not seem to require such process-
equal at one point in the device, where the total flux is free carriers. ing steps60. This may be because of their high effective homologous
No additional exciton dissociation features will be required in this temperatures, the tendency for the inorganic planes to align parallel
case. Transport could then be treated as free carrier with transport to substrates61 (which reduces one degree of misorientations), crys-
and recombination parameters determined by a weighted combi- tallographic flexibility (allowing more graceful accommodation of
nation of free carrier and excitonic values56. For a relatively narrow misorientations than in more rigid materials), or a combination of
coupling range lying between the low and high extremes, interme- all these factors.
diate behaviour is expected54. As well as strengthening absorption, A recent study 62 investigated the activity of intrinsic intergranular
excitons provide an additional pool of carriers that may assist carrier defects in CH3NH3PbI3 using density functional theory. Two types of
transport. Benefits are expected if the lifetime of exciton diffusion or intrinsic defects were studied: neutral Schottky defects (equal num-
recombination exceeds that of free carriers54. However, despite rela- ber of positive and negative vacancies) and Frenkel defects (equal
tively high exciton binding energies and uncertainties arising from number of vacancies and interstitials of the same ion). Schottky
attempting to separate correlated electron–hole pairs into bound and defects, such as PbI2 and CH3NH3I vacancies, were found not to
scattering states55, it seems as if exciton concentrations will be low generate defect states having energies within the perovskite band-
compared to both free-carrier concentrations except in extreme cir- gap; this was attributed to the ionic bonding of the perovskite. This
cumstances, a conclusion supported by a recently published work49. implies Schottky defects are unlikely to be effective as non-radiative
Another striking attribute of these perovskites is their low non- recombination centres. Elemental defects, such as Pb, I and CH3NH3
radiative recombination rates compared to other thin-film poly- vacancies associated with Frenkel defects, were found to form shal-
crystalline semiconductors. This property manifests itself in the low levels near band edges, again reducing the effectiveness as
relatively small difference between Voc of experimental cells and non-radiative recombination centres. However, these conclusions
their effective bandgap potential3 (Eg/q) or alternatively by their may need to be moderated by the well-known limitations of den-
high external radiative efficiency, a parameter deducible from Voc sity functional theory studies14,45. Another recent study 63 that used a
and the spectral response57. The best perovskite cells have relatively similar approach reached similar conclusions. Dominant defects in
low values for the difference Eg/q − Voc (about 450 meV, ref. 3) and the iodide are deduced to be p-type Pb vacancies and n-type meth-
relatively high calculated external radiative efficiencies (0.058%; ylammonium interstitials, with growth conditions determining the
I. Al Mansouri, personal communication). This makes perovskites final doping polarity. This study also considered anti-site substitu-
particularly interesting for high Eg cells in tandem cell stacks3, where tions, which, along with Pb interstitials, formed states near the mid-
the high Voc values of such cells give rise to substantial efficiency dle of the bandgap. However, formation energies were high, meaning
advantages. The diverse range of fabrication methods that have been that these defects are expected to form in low concentrations at the
used, which include low-temperature approaches, further enhances relatively low material deposition temperature involved. Although
the prospects of these perovskites as a tandem cell component. experimental confirmation is required, these results may help explain

NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics 509


© 2014 Macmillan Publishers Limited. All rights reserved.
REVIEW ARTICLE NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.134

a b Electron beam c
A χP
Perovskite
χP
TiO2
2 ФHTM
χT χF
7 1 4 6

5 hf
Au Perovskite FTO
HTM TiO2 Glass HTM FTO

EBIC signal
Perovskite
3
HTM
Perovskite 1 µm

Scan position TiO2 FTO

Figure 4 | Electron-transfer processes in nanoparticle and bulk cells together with a bulk energy-band diagram. a, Schematic of electron-transfer
processes in a perovskite nanoparticle or ETA device71. The thick green and thin red arrows respectively indicate the processes desirable for energy
conversion and those associated with losses. hf, photon energy. b, Schematic of EBIC experiment 72. A scanned electron beam generates a cloud of
carriers, creating a position-dependent current in a short-circuiting load. c, Energy-band diagram deduced from the vacuum energy levels shown in Fig. 2b.
χP, χT and χF represent the electron affinities of the perovskite, TiO2 and FTO layers, respectively, and ΦHTM represents the work function of the HTM layer.
(The barrier at the TiO2/FTO interface has been a source of discussion in the dye-cell literature97,98, because it impedes carrier collection in the direction
shown. Some work suggests states generated at the interface under white-light illumination cause a pinning effect, reducing its significance98. Other work
suggests a failure of the Anderson rule, which is not uncommon, and that the barrier has the opposite direction at thermal equilibrium99.)

why high-performance perovskite cells can be produced by a diverse points on the surfaces of titania nanoparticles26 to conventional pla-
range of deposition approaches and a wide variety of cell structures. nar thin-film operation32.
Although the dielectric properties of perovskites have not yet In the dye-sensitized and ETA configurations, perovskites are
been shown to be important for good performance, they are so not required to have good carrier transport, as the interfacial prop-
extreme compared to those of conventional semiconductors that erties mainly determine the performance (Fig.  4a depicts the key
it is a possibility. Figure  3b shows a composite of reported data processes involved71). The desirable processes involve photoex-
obtained at 300  K for CH3NH3PbI3. At low frequencies, the die- citation in perovskite (1), electron transfer to titania (2) and hole
lectric constant is large; it has been reported to be 60.9 over the transfer to the HTM (3) (or, equivalently, electron transfer from the
20 Hz – 1 MHz range64, which is consistent with the low-frequency HTM to the perovskite). Undesirable processes are recombination
value of 60.2 obtained from fits at higher frequencies65. This low- of photogenerated species (4), back charge transfer at the interfaces
frequency value is appropriate in determining steady-state prop- of TiO2 and the HTM with the perovskite (5,6) and between TiO2
erties, with static fields consequently varying slowly with position and the HTM (7) (this may occur if perovskite is absent in some
— five times slower than in silicon under similar electrostatic dis- areas — for example, when nanoparticles or voids are present). For
turbances. The high value of the dielectric constant results from a high performance, processes (4)–(7) must operate on much slower
combination of dipolar, ionic and electronic contributions. As the timescales than charge generation and extraction (1)–(3). Time-
excitation frequency increases, the permanent dipole associated resolved transient techniques have proved useful for studying the
with the organic cation can no longer respond and a new plateau is related kinetics72.
reached with a dielectric constant of 29.7 (ref. 65). Finally, at infra- For thicker perovskite layers, photogenerated carrier transport
red frequencies, the ionic component drops out, leaving only the becomes important. Initial data suggested that the mixed halide
electronic response. The dielectric constant drops to 6.5 at optical CH3NH3PbI3−xClx has a distinct advantage over the pure iodide —
frequencies66, lower than that of inorganic semiconductors with a its carrier diffusion length (>1 μm; ref. 73) is about ten times greater
similar bandgap. The dashed lines in Fig. 3b indicate possible tran- than that (about 100 nm) of the pure iodide73,74. Subsequent work
sition values of the constants. with the iodide prepared by the vapour-phase CH3NH3I conversion
The other possibility is that CH3NH3PbI3 may possess paraelec- of PbI2 resulted in good carrier collection in 350-nm-thick iodide
tric or even ferroelectric properties at room temperature and above, films, suggesting that the diffusion lengths exceed this thickness34;
impacting device performance67,68. The main evidences for this is this is supported by recent electron-beam measurements60.
the hysteresis observed in resistivity measurements, which is not Recently, collection across a film cross-section has been directly
seen in related compounds prepared similarly, and the remnant probed using the electron beam induced current (EBIC) technique
polarization, which is apparent as a finite voltage output at zero cur- (Fig. 4b)60,72. In this technique, the electron beam generates a cloud
rent 29. The crystallographic point group (4mm) of both the room- of excited carriers (these are expected to be largely free electron–hole
temperature and high-temperature iodide phases is consistent with pairs, some of which may combine to form excitons if they have short
ferroelectric behaviour 29. Hysteresis has also been reported for association times). Because the resolution is limited by the size of this
high-efficiency devices69,70, although many other explanations are cloud, approximately 1,500-nm-thick perovskite layers were grown,
also possible for this besides ferroelectric effects69. Best performance which are appreciably thicker than those in optimal planar cells.
has been realized when hysteresis is minimized (S. I. Seok, personal The typical line scan shown in Fig. 4b has two interesting features.
communication). Carrier collection along the boundaries of micro- The first is the double peak in the EBIC signal within the perovs-
scopic ferroelectric domains has also been suggested68, although kite. This indicates the cell’s ambipolar response with carriers gener-
this presently lacks experimental support. ated both near the spiro-MeOTAD HTM and near the ETM (TiO2
plus FTO) being collected. The second interesting feature is the dip
Device operation between the two peaks, which arises because an appreciable fraction
High efficiencies have been realized using a diverse range of oper- of the generated carriers recombine on their way to their preferred
ating modes, spanning from generation and collection at sparse contact, due to the sample thickness exceeding the optimal thickness.

510 NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics


© 2014 Macmillan Publishers Limited. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.134 REVIEW ARTICLE
These features can be understood by constructing the expected obtained is roughly consistent with that shown in Fig.  4c. TiO2
thermal-equilibrium energy-band diagram of the device from the and the perovskite were deduced to have similar electron affini-
vacuum energy levels shown in Fig. 2b. At thermal equilibrium, a ties, as shown. Undoped spiro-MeOTAD was used in the study 80
common Fermi level prevails throughout the device. A first-order with the alignments deduced consistent with those in Fig.  4c for
estimate of how the energies within the different materials align the doped material.
when brought together is possible by the combined use of the
Anderson rule for semiconductor heterojunctions75 and the related Commercialization challenges
Schottky–Mott rule for metal–semiconductor interfaces. Both are A series of unsuccessful attempts to commercialize new solar-cell
known to have severe limitations, but they have the advantage over technology in recent years vividly demonstrates the non-trivial chal-
more sophisticated approaches76 of being universal in application. lenges of commercialization. A key prerequisite for commercializa-
In both rules, the vacuum reference level is assumed to be continu- tion is a compelling market advantage over incumbent technologies.
ous across material interfaces. For semiconductors, this means that For perovskites, such an advantage might be low processing costs,
the relative alignment of conduction band edges across interfaces high conversion efficiencies through tandem device structures or
depends on the difference of the electron affinities of the two mate- unique products, such as the above-mentioned flexible or partially
rials (the energy from the conduction band edge to the vacuum transparent modules. The present reliance on Pb as a key perovskite
level). For metal–semiconductor interfaces, the conduction band component may militate against the adoption of such products in
edge will lie above the metal Fermi level by an energy equal to the consumer or building integrated applications. The present relative
metal work function minus the semiconductor electron affinity. lack of robustness may present an additional barrier for mainstream
Applying these two rules and treating the HTM as a metal gives bulk power applications.
the energy-band diagram shown in Fig.  4c for the case when the In terms of cost, the closest commercial technology that perovs-
perovskite layer is reasonably thick. The associated potential varia- kites must compete with is CdTe, the photovoltaic thin-film tech-
tions across the perovskite are due to the low work function of the nology with the lowest production cost. Fabrication is by simple
compact TiO2 layer (doped at about 1018–1019 cm−3, ref.  77) com- vapour-phase deposition onto FTO-coated glass with a ‘glass in’ to
pared to the HTM layer. Background perovskite doping levels are ‘module out’ time of 2.5  h (ref.  81). Semiconductor costs are low,
low (they are commonly estimated to be ~1014–1016 cm−3; resistivity with Te currently costing <$115 kg−1. Other material costs (such as
and Seebeck coefficient measurements29 suggest even lower values, those of FTO-coated glass sheets, junction boxes, tabbing, encapsu-
but they are probably dependent on preparation conditions). This lants, sealants and tempered glass backsheets) account for most of
results in the build up of holes near the HTM and electrons near the the manufacturing cost, which was reported to be US$0.54 W−1 in
ETM (their concentrations equal the product of the effective density the final quarter of 2013 or US$72 m−2 at an average module con-
of states in the respective band and the negative exponential of the version efficiency of 13.4% (ref. 82). There is considered to be lim-
difference between the Fermi level and this band edge normalized ited potential for reducing the cost per unit area, which is primarily
by the thermal energy, kT, where k is the Boltzmann constant and determined by the above-mentioned material costs; US$68  m−2 is
T is the absolute temperature). The slope of either band edge gives the target for the end of 201783. This reduction in conjunction with
the local electric field, which is strongest near the HTM and the a predicted increase in the average module efficiency to 17.2% over
ETM. For thick devices, these high-field regions may be independ- this timeframe84 is anticipated to reduce the manufacturing cost to
ent of each other (as shown in Fig. 4c), whereas they merge for thin US$0.38-0.41 W−1 by 2017, which is slightly higher than the pro-
devices, eventually creating an essentially uniform field across the jected manufacturing costs for mainstream Si-based modules pro-
device. With increasing device voltage, the high-field regions tend duced by leading manufacturers85.
to decouple. Although perovskite modules could arguably be processed more
The EBIC outputs are low short-circuit currents, which is con- simply than CdTe modules, this is unlikely to greatly affect the cost
sistent with the use of small-signal calculations in which a thermal- per unit area, because their overall cost is mainly determined by
equilibrium equivalent circuit is used to model carrier drift and the costs of the mostly common materials. In fact, perovskite mod-
diffusion78. The low response when the beam is centred at the HTM/ ules may require more expensive encapsulation to overcome their
perovskite interface is probably because of a high electron recombi- present relatively low robustness. Consequently, perovskite mod-
nation velocity at this interface, corresponding to the back charge- ules would need to at least match the projected efficiencies of CdTe
transfer process (6) in Fig. 4a. This velocity needs to be lower than modules to gain a competitive advantage over them. This may just
the drift velocity of electrons to prevent the formation of a dead layer be possible with single cells, for which both technologies appear
near this interface78. The response at the TiO2–perovskite interface capable of achieving efficiencies of over 20%. The opportunity for
is much higher, suggesting that there is no dead layer near this inter- perovskite modules could lie with tandem cells, where a clear effi-
face and that carrier generation in the TiO2 itself might contribute ciency advantage over CdTe appears feasible without greatly affect-
to the EBIC signal. The higher peak value on the HTM side of the ing the manufacturing cost per unit area but significantly decreasing
device may correspond to a more extended high-field region on this the cost per watt.
side as a result of the band alignment imposed at the HTM interface. The present perovskites share a common disadvantage with
Recent work79,80 has allowed the accuracy of the Anderson and CdTe, namely reliance on an environmentally hazardous heavy
Schottky–Mott rules used for deducing Fig.  4c to be assessed. metal. Apart from in Japan and for specific applications (such as
Lindblad et  al.79 found that the alignment between the occupied residential rooftops), the toxicity of CdTe appears to have had
valence bands of iodide perovskite and nanoporous TiO2 was 2.1 eV minimal impact on market acceptance to date. Photovoltaic mod-
using hard-X-ray photoelectron spectroscopy. Using ultraviolet ules are presently exempted from legislation such as the European
photoemission spectroscopy and inverse photoemission spectros- Restriction on Hazardous Substances (RoHS)86, at least for large,
copy to respectively interrogate occupied and unoccupied states, stationary, professionally installed systems. Without this exemp-
Schulz et  al.80 self-consistently deduced the work functions, elec- tion, CdTe modules would have no possibility of meeting the RoHS
tron affinities and valence-band maxima. Moreover, all measure- requirement of having less than 0.01% Cd by weight in any homo-
ments were conducted in the same ultrahigh vacuum for compact geneous layer (defined as a layer of “uniform composition through-
uncoated TiO2 samples, TiO2 samples coated with iodide, bro- out” or one “that cannot be … separated into different materials by
mide and mixed chloride perovskites, and for the latter samples mechanical actions such as unscrewing, cutting, crushing, grinding
overlaid with different thicknesses of the HTM. The final picture and abrasive processes”86). Modules based on present perovskites

NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics 511


© 2014 Macmillan Publishers Limited. All rights reserved.
REVIEW ARTICLE NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.134
fare better, as the RoHS limit on Pb in a homogeneous layer is ten technology that uses both perovskite and existing technologies; this
times higher than that for Cd, but they also appear to have no pros- may allow market introduction as a new premium product.
pects for compliance if they contain the optional continuous layer
shown in Fig. 2a. If the scaffold layer were accepted as a homogenous Received 10 February 2014; accepted 12 May 2014; published
layer, the Pb content would be diluted in both nanoparticle-sensi- online 27 June 2014
tized and ETA devices. (If 18 2.5-nm-diameter hemispherical iodide
perovskite nanoparticles coat a 20-nm-diameter TiO2 nanosphere21, References
the scaffolding layer has a porosity of 60% and the pores are filled by 1. Turner, G. Global Renewable Energy Market Outlook 2013. Bloomberg New
Energy Finance https://www.bnef.com/insightdownload/7526/pdf
spiro-MeOTAD, the Pb content will be 0.4–0.5% by weight, which
(11 April 2014).
exceeds the RoHS limit, although not appreciably). Cd and Pb com- 2. Park, N.-G. Organometal perovskite light absorbers toward a 20% efficiency
pounds have different solubilities5, which does not affect compli- low-cost solid-state mesoscopic solar cell. J. Phys. Chem. Lett. 4,
ance with RoHS criteria, but it may affect end-of-life disposal. 2423–2429 (2013).
The low robustness of present perovskite technology to moist 3. Snaith, H. J. Perovskites: the emergence of a new era for low-cost, high-
air and water vapour means that out of all the present commer- efficiency solar cells. J. Phys. Chem. Lett. 4, 3623–3630 (2013).
4. Kim, H.-S., Im, S. H. & Park, N.-G. Organolead halide perovskite: new
cial photovoltaic technologies it is most closely related to CIGS.
horizons in solar cell research. J. Phys. Chem. C 118, 5615–5625 (2014).
Unencapsulated CIGS cells generally degrade during damp heat 5. Hodes, G. & Cahen, D. Photovoltaics: perovskite cells roll forward. Nature
testing, although this can be controlled to acceptable levels by Photon. 8, 87–88 (2014).
employing suitable encapsulation. A recent study investigating the 6. Service, R. F. Perovskite solar cells keep on surging. Science 344, 458 (2014).
feasibility of encapsulating CIGS cells in flexible modules concluded 7. Li, C. et al. Formability of ABX3 (X = F, Cl, Br, I) halide perovskites. Acta
that ‘breathable’ designs are not viable87,88. Layers that effectively Crystallogr. B 64, 702–707 (2008).
8. McKinnon, N. K., Reeves, D. C. & Akabas, M. H. 5-HT3 receptor ion size
restrict water ingress are required together with internal encapsulant
selectivity is a property of the transmembrane channel, not the cytoplasmic
layers with a high moisture solubility to keep their relative moisture vestibule portals. J. Gen. Physiol. 138, 453–466 (2011).
saturation low. Double glass layers are used in present commercial 9. Cohen, B. N., Labarca, C., Davidson, N. & Lester, H. A. Mutations in M2 alter
CIGS modules together with edge sealants, both of which effectively the selectivity of the mouse nicotinic acetylcholine receptor for organic and
prevent moisture penetration89,90. Alternative approaches, such as alkali metal cations. J. Gen. Physiol. 100, 373–400 (1992).
integrating moisture barriers like Al2O3 into the device structure, 10. Im, J.-H., Chung, J., Kim, S.-J. & Park, N.-G. Synthesis, structure, and
photovoltaic property of a nanocrystalline 2H perovskite-type novel sensitizer
might eventually eliminate the need to use such restricted designs
(CH3CH2NH3)Pbl3. Nanoscale Res. Lett. 7, 353 (2012).
for perovskite91 and CIGS92 cells. Degradation under ultraviolet 11. Koh, T. M. et al. Formamidinium-containing metal-halide: an alternative
exposure, which is reportedly more severe for devices with TiO2 material for near-IR absorption perovskite solar cells. J. Phys. Chem. C
scaffolding, may similarly be addressed by protective encapsulation http://dx.doi.org/10.1021/jp411112k (13 December 2013).
features (in this case, ultraviolet filtering) or device design93. 12. Eperon, G. E. et al. Formamidinium lead trihalide: a broadly tunable
perovskite for efficient planar heterojunction solar cells. Energy Environ. Sci. 7,
982–988 (2014).
Summary and future prospects 13. Pang, S. et al. NH2CH=NH2Pbl3: An alternative organolead iodide perovskite
The next few years promise to be exciting ones for research and sensitizer for mesoscopic solar cells. Chem. Mater. 26, 1485–1491 (2014).
development of organic–inorganic halide perovskite solar cells. 14. Umari, P., Mosconi, E. & De Angelis, F. Relativistic GW calculations on
On-going efficiency improvements are expected, as well as a rap- CH3NH3PbI3 and CH3NH3SnI3 perovskites for solar cell applications. Sci. Rep.
idly growing understanding of their material properties and optimal 4, 4467 (2014).
cell designs. 15. Topsöe, H. Krystallographisch-chemische untersuchungen homologer
verbindungen. Zeitschrift für Kristallographie 8, 246–296 (1884).
Advantages over existing photovoltaic technologies include
16. Mitzi, D. B., Wang, S., Feild, C. A., Chess, C. A. & Guloy, A. M. Conducting
material properties that simplify the manufacture of high-perfor- layered organic–inorganic halides containing <110>-oriented perovskite
mance devices. The diversity in demonstrated approaches may give sheets. Science 267, 1473–1476 (1995).
rise to low processing costs and simple implementation of attractive 17. Mitzi, D. B., Chondroudis, K. & Kagan, C. R. Organic-inorganic electronics.
products, such as flexible, transparent or all-perovskite tandem cell IBM J. Res. Dev. 45, 29–45 (2001).
modules. This diversity may also allow perovskite cells to be directly 18. Kojima, A., Teshima, K., Miyasaka, T. & Shirai, Y. Novel photoelectrochemical
cell with mesoscopic electrodes sensitized by lead-halide compounds (2). in
integrated with other cell technologies to form high-performance
Proc. 210th ECS Meeting (ECS, 2006).
tandem cells; Si and CIGS modules appear particularly promising 19. Kojima, A., Teshima, K., Shirai, Y. & Miyasaka, T. Organometal halide
in this respect 3,6. perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc.
In the present market, the toxicity of Pb is not a major impedi- 131, 6050–6051 (2009).
ment to large-scale, professional applications, as is evidenced by the 20. Kojima, A., Teshima, K., Shirai, Y. & Miyasaka, T. Novel photoelectrochemical
fact that CdTe cells have already gained a reasonable market share. cell with mesoscopic electrodes sensitized by lead-halide compounds (11). in
Proc. 214th ECS Meeting (ECS, 2014).
Cd or Pb is also present in some CIGS and silicon modules at the
21. Im, J.-H., Lee, C.-R., Lee, J.-W., Park, S.-W. & Park, N.-G. 6.5% efficient
same general level as that likely in perovskite modules94. The danger perovskite quantum-dot-sensitized solar cell. Nanoscale 3, 4088–4093 (2011).
is that technology relying on toxic materials may be increasingly 22. Kim, H.-S. et al. Lead iodide perovskite sensitized all-solid-state submicron
marginalized as legislation becomes increasingly more pervasive thin film mesoscopic solar cell with efficiency exceeding 9%. Sci. Rep. 2,
and restrictive. Elimination of Pb seems the only sure solution, 591 (2012).
with replacement by Sn a possibility 95. Alternatively, research into 23. Lee, M. M., Teuscher, J., Miyasaka, T., Murakami, T. N. & Snaith, H. J.
Efficient hybrid solar cells based on meso-superstructured organometal halide
the present perovskites might allow more precise determination of
perovskites. Science 338, 643–647 (2012).
the features that have resulted in such rapid progress, encouraging 24. Salbeck, J., Yu, N., Bauer, J., Weissörtel, F. & Bestgen, H. Low molecular
identification and investigation of non-toxic material systems with organic glasses for blue electroluminescence. Synthetic Metals 91,
similar properties. 209–215 (1997).
The recent surge in interest makes it likely there will be multi- 25. Bach, U. et al. Solid-state dye-sensitized mesoporous TiO2 solar cells with
ple attempts to commercialize perovskite photovoltaic products in high photon-to-electron conversion efficiencies. Nature 395, 583–585 (1998).
26. Stranks, S. D. et al. Electron–hole diffusion lengths exceeding 1 micrometer in
coming years. Perovskites have the advantage that they provide mul-
an organometal trihalide perovskite absorber. Science 342, 341–344 (2013).
tiple paths to commercialization. As well as the traditional approach 27. Heo, J. H. et al. Efficient inorganic–organic hybrid heterojunction solar cells
of challenging established manufacturers, there is also the opportu- containing perovskite compound and polymeric hole conductors. Nature
nity to work with these to develop high-performance tandem cell Photon. 7, 486–491 (2013).

512 NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics


© 2014 Macmillan Publishers Limited. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.134 REVIEW ARTICLE
28. Noh, J. H, Im, S. H., Heo, J. H., Mandal, T. N. & Seok, S. I. Chemical 59. Edalati, K. & Horita, Z. Significance of homologous temperature in softening
management for colorful, efficient, and stable inorganic-organic hybrid behavior and grain size of pure metals processed by high-pressure torsion.
nanostructured solar cells. Nano Lett. 13, 1764–1769 (2013). Mater. Sci. Eng. A 528, 7514–7523 (2011).
29. Stoumpos, C. C., Malliakas, C. D. & Kanatzidis, M. G. Semiconducting tin 60. Edri, E. et al. Why lead methylammonium tri-iodide perovskite-based solar
and lead iodide perovskites with organic cations: phase transitions, high cells require a mesoporous electron transporting scaffold (but not necessarily a
mobilities, and near-infrared photoluminescent properties. Inorg. Chem. 52, hole conductor). Nano Lett. 14, 1000–1004 (2014).
9019–9038 (2013). 61. Liang, K., Mitzi D. B. & Prikas, M. T. Synthesis and characterization of
30. Burschka, J. et al. Sequential deposition as a route to high-performance organic–inorganic perovskite thin films prepared using a versatile two-step
perovskite-sensitized solar cells. Nature 499, 316–319 (2013). dipping technique. Chem. Mater. 10, 403–411 (1998).
31. Green, M. A., Emery, K., Hishikawa, Y., Warta, W. & Dunlop, E. D. Solar cell 62. Kim, J., Lee, S.-H., Lee, J. H. & Hong, K.-H. The role of intrinsic defects in
efficiency tables (version 43). Prog. Photovolt. 22, 1–9 (2014). methylammonium lead iodide perovskite. J. Phys. Chem. Lett. 5,
32. Liu, M., Johnston, M. B. & Snaith, H. J. Efficient planar heterojunction 1312–1317 (2014).
perovskite solar cells by vapour deposition. Nature 501, 395–398 (2013). 63. Yin, W.-J., Shi, T. & Yan, Y. Unusual defect physics in CH3NH3PbI3 perovskite
33. Burschka, J. High performance solid-state mesoscopic solar cells. PhD thesis, solar cell absorber. Appl. Phys. Lett. 104, 063903 (2014).
École Polytechnique Fédérale de Lausanne 107 (2013). 64. Onoda-Yamamuro, N., Matsuo, T. & Suga, H. Dielectric study of CH3NH3PbX3
34. Chen, Q. et al. Planar heterojunction perovskite solar cells via vapor-assisted (X = Cl, Br, I). J. Phys. Chem. Solids 53, 935–939 (1992).
solution process. J. Am. Chem. Soc. 136, 622–625 (2014). 65. Poglitsch, A. & Weber, D. Dynamic disorder in
35. Liu, D. & Kelly, T. L. Perovskite solar cells with a planar heterojunction methylammoniumtrihalogenoplumbates (II) observed by millimeter-wave
structure prepared using room-temperature solution processing techniques. spectroscopy. J. Chem. Phys. 87, 6373 (1987).
Nature Photon. 8, 133–138 (2014). 66. Hirasawa, M., Ishihara, T., Goto, T., Uchida, K. & Miura, N.
36. Christians, J. A., Fung, R. C. M. & Kamat, P. V. An inorganic hole conductor Magnetoabsorption of the lowest exciton in perovskite-type compound
for organo-lead halide perovskite solar cells. Improved hole conductivity with (CH3NH3)PbI3. Physica B 201, 427–430 (1994).
copper iodide. J. Am. Chem. Soc. 136, 758–764 (2014). 67. Yuan, Y., Xiao, Z., Yang, B. & Huang, J. Arising applications of ferroelectric
37. Ito, S. Pb perovskite solar cells using inorganic hole conductor of CuSCN. materials in photovoltaic devices. J. Mater. Chem. A 2, 6027–6041 (2014).
Paper Y-RS-14 in 2013 MRS Fall Meeting & Exhibit (MRS, 2013). 68. Frost, J. M. et al. Atomistic origins of high-performance in hybrid halide
38. Docampo, P., Ball, J. M., Darwich, M., Eperon, G. E. & Snaith, H. J. Efficient perovskite solar cells. Nano Lett. 14, 2584–2590 (2014).
organometal trihalide perovskite planar-heterojunction solar cells on flexible 69. Snaith, H. J. et al. Anomalous hysteresis in perovskite solar cells. J. Phys. Chem.
polymer substrates. Nature Commun. 4, 2761 (2013). Lett. 5, 1511–1515 (2014).
39. Malinkiewicz, O. et al. Perovskite solar cells employing organic charge- 70. Hoke, E. T., Unger, E. L., Vandewal, K. & McGehee, M. D. Charge
transport layers. Nature Photon. 8, 128–132 (2014). recombination and transport in hybrid perovskite solar cells: why do
40. Sun, S. et al. The origin of high efficiency in low-temperature solution- perovskite solar cells have large Voc? in Proc. MRS Fall Meeting and
processable bilayer organometal halide hybrid solar cells. Energy Environ. Sci. Exhibit (2013).
7, 399–407 (2014). 71. Marchioro, A. et al. Unravelling the mechanism of photoinduced charge
41. Wang, J. T.-W. et al. Low-temperature processed electron collection layers of transfer processes in lead iodide perovskite solar cells. Nature Photon. 8,
graphene/TiO2 nanocomposites in thin film perovskite solar cells. Nano Lett. 250–255 (2014).
14, 724–730 (2014). 72. Edri, E. et al. Elucidating the charge carrier separation mechanism in
42. Wojciechowski, K., Saliba, M., Leijtens, T., Abate, A. & Snaith, H. J. Sub- CH3NH3PbI3-xClx perovskite solar cells. Nature Commun. 5, 3461 (2014).
150 °C processed meso-superstructured perovskite solar cells with enhanced 73. Stranks, S. D. et al. Electron–hole diffusion lengths exceeding
efficiency. Energy Environ. Sci. 7, 1142–1147 (2014). 1 micrometer in an organometal trihalide perovskite absorber. Science 342,
43. Eperon, G. E., Burlakov, V. M., Goriely, A. & Snaith, H. J. Neutral color 341–344 (2013).
semitransparent microstructured perovskite solar cells. ACS Nano 8, 74. Xing, G. et al. Long-range balanced electron- and hole-transport lengths in
591–598 (2014). organic–inorganic CH3NH3Pbl3. Science 342, 344–347 (2013).
44. Wang, J. et al. Performance improvement of amorphous silicon see-through 75. Anderson, R. L. Germanium-gallium arsenide heterojunction. IBM J. Res. Dev.
solar modules with high transparency by the multi-line ns-laser scribing 4, 283–287 (1960).
technique. Opt. Las. Eng. 51, 1206–1212 (2013). 76. Peressi, M., Baldereschi, A. & Baroni, S. in Characterization of Semiconductor
45. Even, J., Pedesseau, L., Jancu, J.-M. & Katan, C. Importance of spin-orbit Heterostructures and Nanostructures 2nd Edn (eds Agostini G. & Lamberti, C)
coupling in hybrid organic/inorganic perovskites for photovoltaic applications. Ch. 2 (Elsevier, 2008).
J. Phys. Chem. Lett. 4, 2999–3005 (2013). 77. Kavan, L. & Grätzel, M. Highly efficient semiconducting TiO2
46. Sell, D. D. & Lawaetz, P. New analysis of direct exciton transitions: application photoelectrodes prepared by aerosol pyrolysis. Electrochimica Acta 40,
to GaP. Phys. Rev. Lett. 26, 311–314 (1971). 643–652 (1995).
47. Even, J., Pedesseau, L., Dupertuis, M.-A., Jancu, J.-M. & Katan, C. Electronic 78. Green, M. A. The depletion layer collection efficiency for p-n junction,
model for self-assembled hybrid organic/perovskite semiconductors: reverse Schottky diode, and surface insulator solar cells. J. Appl. Phys. 47,
band edge electronic states ordering and spin-orbit coupling. Phys. Rev. B 86, 547–554 (1976).
205301 (2012). 79. Lindblad, R. et al. Electronic structure of TiO2/CH3NH3PbI3 perovskite solar
48. Ishihara, T. Optical properties of PbI-based perovskite structures. cell interfaces. J. Phys. Chem. Lett. 5, 648–653 (2014).
J. Luminescence 60, 269–274 (1994). 80. Schulz, P. et al. Interface energetics in organo-metal halide perovskite-based
49. D’Innocenzo, V. et al. Excitons versus free charges in organo-lead tri-halide photovoltaic cells. Energy Environ. Sci. 7, 1377–1381 (2014).
perovskites. Nature Commun. 5, 3586 (2014). 81. First Solar Sets New World Record for CdTe Solar Cell Efficiency. http://investor.
50. Tauc, J. Optical properties and electronic structure of amorphous Ge and Si. firstsolar.com/releasedetail.cfm?releaseid=743398 (26 February 2013).
Mater. Res. Bull. 3, 37–46 (1968). 82. Widmar, M. First Solar Q4’13 Earnings Call. http://files.shareholder.
51. Elliott, R. J. Intensity of optical absorption by excitons. Phys. Rev. 108, com/downloads/fslr/1347979521x0x728649/bddfd430-a025-43e4-
1384–1389 (1957). 8b91-c1915066b274/q413_earnings_call_presentation_final1.pdf
52. Tanaka, K. et al. Comparative study on the excitons in lead-halide-based (25 February 2014).
perovskite-type crystals CH3NH3PbBr3 CH3NH3PbI3. Solid State Commun. 83. De Jong, T. First Solar Manufacturing Update. http://files.shareholder.
127, 619–623 (2003). com/downloads/fslr/3084163747x0x652323/53f7a04f-fcf5-4729-8bb2-
53. Combescot, M. Thermodynamics of an electron-hole system in 0abf6033c046/4.%20fsanalystday_manufacturing.pdf (11 April 2014).
semiconductors. Phys. Status Solidi B 86, 349–358 (1978). 84. Garabedian, R. Technology Update. 2013 Analyst Meeting, First Solar.
54. Corkish, R., Chan, D. S.-P. & Green, M. A. Excitons in silicon diodes and solar http://files.shareholder.com/downloads/FSLR/3084163747x0x652328/
cells: a three-particle theory. J. Appl. Phys. 79, 195–203 (1996). d0af6554-e193-47e4-9dd0-59b02968272b/fsanalystday_technologyupdate.pdf
55. Geelhaar, F. Coulomb Correlation Effects in Silicon Devices (Series in (11 April 2014).
Microelectronics 147 Hartung-Gorre, 2004). 85. Rinaldi, N. Solar PV Module Costs to Fall to 36 Cents per Watt by 2017 http://
56. Green, M. A. Many-body theory applied to solar cells: excitonic and related carrier www.greentechmedia.com/articles/read/solar-pv-module-costs-to-fall-to-36-
correlation effects in Proc. 26th IEEE Photovoltaic Specialists Conference (1997). cents-per-watt (18 June 2013).
57. Green, M. A. Radiative efficiency of state-of-the-art photovoltaic cells. Prog. 86. Directive 2011/65/EU of the European Parliament and of the Council of 8 June
Photovolt. 20, 472–476 (2012). 2011 on the restriction of the use of certain hazardous substances in electrical
58. Ungár, T. The meaning of size obtained from broadened X-ray diffraction and electronic equipment (recast). http://eur-lex.europa.eu/legal-content/en/
peaks. Adv. Eng. Mater. 5, 323–329 (2003). TXT/?uri=celex:32011L0065 (8 June 2011).

NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics 513


© 2014 Macmillan Publishers Limited. All rights reserved.
REVIEW ARTICLE NATURE PHOTONICS DOI: 10.1038/NPHOTON.2014.134
87. Coyle, D. J. Life prediction for CIGS solar modules part 1: 96. Wehrenfennig, C., Liu, M., Snaith, H. J., Johnston, M. B. & Herz, L. M.
modeling moisture ingress and degradation. Prog. Photovolt. 21, Homogeneous emission line broadening in the organo lead halide perovskite
156–172 (2013). CH3NH3PbI3-xClx. J. Phys. Chem. Lett. 5, 1300–1306 (2014).
88. Coyle, D. J. et al. Life prediction for CIGS solar modules part 2: degradation 97. Rühle, S. & Dittrich, T. Investigation of the electric field in TiO2/FTO junctions
kinetics, accelerated testing, and encapsulant effects. Prog. Photovolt. 21, used in dye-sensitized solar cells by photocurrent transients. J. Phys. Chem. B
173–186 (2013). 109, 9522–9526 (2005).
89. Kempe, M. D., Dameron, A. A. & Reese, M. O. Evaluation of moisture ingress 98. Snaith, H. J. & Grätzel, M. The role of a “Schottky barrier” at an electron-
from the perimeter of photovoltaic modules. Prog. Photovolt. http://dx.doi. collection electrode in solid-state dye-sensitized solar cells. Adv. Mater. 18,
org/10.1002/pip.2374 (2013). 1910–1914 (2006).
90. Kempe, M. D., Panchagade, D., Reese, M. O. & Dameron, A. A. Modeling 99. Kron, G., Rau, U. & Werner, J. H. Influence of the built-in voltage on the fill
moisture ingress through polyisobutylene-based edge-seals. Prog. Photovolt. factor of dye-sensitized solar cells. J. Phys. Chem. B 107, 13258–13261 (2003).
http://dx.doi.org/10.1002/pip.2465 (2014).
91. Niu, G. et al. Study on the stability of CH3NH3PbI3 films and the effect of post- Acknowledgements
modification by aluminum oxide in all-solid-state hybrid solar cells. J. Mater. The Australian Centre for Advanced Photonics (ACAP) is supported by the Australian
Chem. A 2, 705–710 (2014). Government through the Australian Renewable Energy Agency (ARENA). Responsibility
92. Carcia, P. F., McLean, R.  S. & Hegedus, S. ALD Moisture barrier for Cu(InGa) for the views, information or advice expressed herein is not accepted by the Australian
Se2 solar cells. in Proc. 218th ECS Meeting (ECS, 2010). Government. H.J. S. is supported by the Engineering and Physical Sciences Research
93. Leijtens, T. et al. Overcoming ultraviolet light instability of sensitized TiO2 Council UK and the European Research Council.
with meso-superstructured organometal tri-halide perovskite solar cells.
Nature Commun. 4, 2885 (2013).
94. Werner, J. H., Zapf-Gottwick, R., Koch, M. & Fischer, K. Toxic Additional information
substances in photovoltaic modules. in Proc. 21st Int. Photovoltaic Sci. Eng. Reprints and permissions information is available at www.nature.com/reprints. Requests
for materials and correspondence should be addressed to M.A.G.
Conf. (2011).
95. Noel, N. K. et al. Lead-free organic-inorganic tin halide perovskites for
photovoltaic applications. Energy Environ. Sci. http://dx.doi.org/10.1039/ Competing financial interests
c4ee01076K (2014). The authors declare no financial interests.

514 NATURE PHOTONICS | VOL 8 | JULY 2014 | www.nature.com/naturephotonics


© 2014 Macmillan Publishers Limited. All rights reserved.
View publication stats

You might also like