ArticlePDF Available

Efficient perovskite light-emitting diodes featuring nanometre-sized crystallites

Authors:

Abstract and Figures

Organic–inorganic hybrid perovskite materials are emerging as highly attractive semiconductors for use in optoelectronics. In addition to their use in photovoltaics, perovskites are promising for realizing light-emitting diodes (LEDs) due to their high colour purity, low non-radiative recombination rates and tunable bandgap. Here, we report highly efficient perovskite LEDs enabled through the formation of self-assembled, nanometre-sized crystallites. Large-group ammonium halides added to the perovskite precursor solution act as a surfactant that dramatically constrains the growth of 3D perovskite grains during film forming, producing crystallites with dimensions as small as 10 nm and film roughness of less than 1 nm. Coating these nanometre-sized perovskite grains with longer-chain organic cations yields highly efficient emitters, resulting in LEDs that operate with external quantum efficiencies of 10.4% for the methylammonium lead iodide system and 9.3% for the methylammonium lead bromide system, with significantly improved shelf and operational stability.
Content may be subject to copyright.
Efcient perovskite light-emitting diodes featuring
nanometre-sized crystallites
Zhengguo Xiao1, Ross A. Kerner1, Lianfeng Zhao1,NhuL.Tran
2,KyungMinLee
1, Tae-Wook Koh1,
Gregory D. Scholes2and Barry P. Rand1,3*
Organicinorganic hybrid perovskite materials are emerging as highly attractive semiconductors for use in optoelectronics.
In addition to their use in photovoltaics, perovskites are promising for realizing light-emitting diodes (LEDs) due to their
high colour purity, low non-radiative recombination rates and tunable bandgap. Here, we report highly efcient perovskite
LEDs enabled through the formation of self-assembled, nanometre-sized crystallites. Large-group ammonium halides
added to the perovskite precursor solution act as a surfactant that dramatically constrains the growth of 3D perovskite
grains during lm forming, producing crystallites with dimensions as small as 10 nm and lm roughness of less than 1 nm.
Coating these nanometre-sized perovskite grains with longer-chain organic cations yields highly efcient emitters,
resulting in LEDs that operate with external quantum efciencies of 10.4% for the methylammonium lead iodide system
and 9.3% for the methylammonium lead bromide system, with signicantly improved shelf and operational stability.
Hybrid organicinorganic perovskites are emerging as a new
generation of low-cost, solution-processed semiconducting
materials with favourable optoelectronic properties such as
strong absorption coefcients, tunable bandgap1,2, large and
balanced electron and hole mobilities, long carrier diffusion
lengths, small exciton binding energy3,4 and unique defect proper-
ties with only shallow point defects formed5. These characteristics
have allowed for consistent and rapid progress in solar cell ef-
ciency, which has risen dramatically from 3.8% to over 22% in
less than ve years due to efforts on perovskite lm morphology
optimization, interface engineering, and so on617.
Owing to their facile solution processing, high colour purity and
tunable bandgap1822, hybrid perovskites are also promising for
LEDs. Nevertheless, the small exciton binding energy in 3D perovs-
kites (methylammonium lead iodide, MAPbI
3
, and methylammo-
nium lead bromide, MAPbBr
3
) result in small electronhole
capture rates for radiative recombination18. Therefore, ultrathin per-
ovskite layers and/or small perovskite grain size have been employed
to spatially conne electrons and holes to promote bimolecular
radiative recombination1820. Recently, an external quantum ef-
ciency (EQE) of 8.5% for MAPbBr
3
-based LEDs has been demon-
strated through the formation of MAPbBr
3
nanograins with an
average size of 100 nm, and reduction of metallic lead impurities19.
However, due to the rapid crystallization speed of 3D perovskites,
grain sizes are generally hundreds of nanometres in size from
either one-step or two-step solution processes13,23,24, with a resulting
large surface roughness.
Here, we report a solution process to form highly uniform and
ultra-at perovskite lms with nanometre-sized grains that allow
us to demonstrate highly efcient LEDs. The addition of long-
chain ammonium halides (for example, n-butylammonium
halides (BAX, X = I, Br)) in the 3D perovskite precursor solution
impedes the growth of 3D perovskite grains and dramatically
decreases lm roughness to 1 nm. The nanometre-sized grains
feature reduced dimensionality, starting a transition from 3D to
layered (so-called RuddlesdenPopper) perovskite structures,
manifesting in stronger and blue-shifted photoluminescence (PL)
and electroluminescence (EL) compared with the emission from
the 3D perovskite. Notably, the EQE of iodide perovskite (I-perovs-
kite, BAI:MAPbI
3
) LEDs increased from 1.0% to 10.4% due to the
incorporation of BAI. Similarly, the EQE of bromide perovskite
(Br-perovskite, BABr:MAPbBr
3
) LEDs increased from 0.03% to
9.3% following the incorporation of BABr. The power efciency
(PE) and current efciency (CE) reached 13.0 lm W
1
and
17.1 cd A
1
, respectively, for Br-perovskite LEDs. Furthermore,
the shelf stability of unencapsulated I- and Br-perovskite LEDs in
N
2
was dramatically improved after adding long-chain halides.
Specically, LEDs without BAX degrade to 6070% of their initial
value within 2 days whereas LEDs with BAX (X = I, Br) show no
degradation after storage in the glove box for more than 8 months.
Perovskite lm formation and characterization
To deposit uniform pin-hole-free perovskite lms with small grain
size, toluene was dropped on the spinning lm at an appropriate
time to extract the processing solvent (dimethylformamide,
DMF), thus halting the morphological evolution and maintaining
small crystallite size25,26. To ensure consistency with the device
structure, the optical, morphological and structural properties of
I- and Br-perovskite lms were probed on poly[N,N-bis(4-butyl-
phenyl)-N,N-bis(phenyl)-benzidine] (poly-TPD) and poly(N-
vinylcarbazole) (PVK), respectively. The perovskite thickness is
approximately 80 nm for I-perovskite lms and 70 nm for Br-per-
ovskite lms. Atomic force microscope (AFM) images of MAPbI
3
(Fig. 1a) and MAPbBr
3
(Fig. 1g) lms without annealing show
that both are very uniform. The root mean square (r.m.s.) roughness
is 4.9 nm for the MAPbI
3
lm and 3.4 nm for the MAPbBr
3
lm,
among the most at and uniform perovskite lms reported.
Long-chain ammonium ions cannot ll the corner of PbX
4
(X = I, Br, Cl) octahedral layers and therefore induce the formation
of layered perovskites27. Notably, when BAX (X = I, Br) is added into
the MAPbX
3
precursor solution, the growth of 3D perovskite grains
during the lm formation process is dramatically impeded.
1Department of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA. 2Department of Chemistry, Princeton University, Princeton,
New Jersey 08544, USA. 3Andlinger Center for Energy and the Environment, Princeton University, Princeton, New Jersey 08544, USA.
*e-mail: brand@princeton.edu
ARTICLES
PUBLISHED ONLINE: 16 JANUARY 2017 | DOI: 10.1038/NPHOTON.2016.269
NATURE PHOTONICS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturephotonics 1
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Figure 1bf and hl show AFM images of I- and Br-perovskite lms
prepared from different molar ratios of BAI:MAPbI
3
and BABr:
MAPbBr
3
in the precursor solution, respectively. As shown in
Fig. 1, when the molar ratio of BAX:MAPbX
3
(X = I, Br) increases
to 10:100, lms feature smaller grain size and therefore reduced
roughness. As the BAX concentration increases, grain size continues
to decrease and lms become smoother, decreasing to 0.6 nm r.m.s.
for I-perovskite lms and 1.0 nm r.m.s. for Br-perovskite lms, for a
molar ratio of BAX:MAPbX
3
(X = I, Br) of 40:100. In fact, the mor-
phologies of these lms are similar to solution-processed polymers,
where grain sizes are too small to be resolved by the AFM tip. When
the BAX:MAPbX
3
molar ratio is increased to larger than 60:100,
larger grains start to form with the roughness of the lm
simultaneously increasing, due to the formation of layered
perovskite phases.
X-ray diffraction (XRD) was conducted to determine the compo-
sition and crystallinity of the lms. As shown in Fig. 1m,n, no dif-
fraction peaks corresponding to layered perovskite grains are
observed when the BAX molar ratio is smaller than 40:100 for
I-perovskite and 20:100 for Br-perovskite lms, indicating that the
as-formed lms are composed of grains that are nearly indistin-
guishable from the 3D perovskite crystal structure, with BA
cations self-assembled at the grain surfaces. It should also be
noted that the full-width at half-maximum (FWHM) of the diffrac-
tion peaks after adding BAX increases, indicating that crystallite size
decreases. The crystallite size of MAPbI
3
calculated from the XRD
I-perovskite Br-perovskite
r.m.s. = 4.9 nm
4.2 nm
0.7 nm
0.6 nm
1.7 nm
r.m.s. = 3.4 nm
2.3 nm
2.4 nm
1.0 nm
3.3 nm
0
40 nm
0
25 nm
5 nm
0
0
15 nm
0
20 nm
0
40 nm
0
80 nm
0
25 nm
0:100
10:100
20:100
40:100
60:100
0:100
10:100
20:100
40:100
60:100
6.5 nm 9.3 nm100:100 100:100
a
b
c
d
e
g
h
i
j
k
fl
mn
6
5
4
Normalized intensity (a.u.)
3
2
1
0
10 15 20
2θ (°)
25
0:100
10:100
20:100
40:100
60:100
Br-perovskite **
*
*
*
*
*
30 35
6
5
4
Normalized intensity (a.u.)
3
2
1
0
10 15 20
2θ (°)
25
0:100
10:100
20:100
40:100
60:100
100:100 100:100
I-perovskite *
*
*
*
30 35
Figure 1 | Morphology and XRD study of I- and Br-perovskite lms. af, AFM images of the I-perovskite lms with different molar ratio of BAI:MAPbI
3
in
the precursor solution. gl, AFM images of the Br-perovskite lms with different molar ratio of BABr:MAPbBr
3
in the precursor solution. The molar ratios
and r.m.s. roughness of the lms are marked on the left and right sides of the images, respectively. All images use a zaxis range of 040 nm to best allow
for direct comparison, whereas the topographic colour-map is changed according to the colour scale bars. The scan area of all images is 2 µm×2µm.
m,n, Normalized XRD data of the I- and Br-perovskite lms with different molar ratio of BAI:MAPbI
3
(m) and BABr:MAPbBr
3
(n). The peaks marked with
* indicate the peaks corresponding to layered perovskite.
ARTICLES NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.269
NATURE PHOTONICS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturephotonics2
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
data applying a Scherrer analysis (Supplementary Information) is 16
and 13 nm, respectively, when the BAI:MAPbI
3
ratio is increased
from 10:100 to 20:100, which is approximately 25 and 20 PbI
4
layers (considering the lattice constants of a=b= 8.86 Å, c=
12.65 Å for the tetragonal structure)6. Similarly, the crystallite size
of the MAPbBr
3
is calculated to be 16 and 11 nm for BABr:
MAPbBr
3
loadings of 10:100 to 20:100, which is approximately 27
and 19 PbBr
4
layers (considering the lattice constants of cubic
MAPbBr
3
:a=b=c= 5.9 Å). In agreement with the AFM and
XRD results, the small grain size of the perovskite lm after incor-
poration of BAX is conrmed by scanning electron microscopy
(SEM) as shown in Supplementary Fig. 1. The crystallite of
MAPbX
3
with BAX self-assembled at the surface can also be con-
sidered within the unit cell of layered BA
2
MA
n1
Pb
n
X
3n+1
RuddlesdenPopper perovskite phases where nequals the number
of PbX
4
octahedral layers in the crystallite. The discrepancy
between the nvalues calculated above and the nvalues expected
from the BAX and MAX molar ratios in the precursor solution
may result from the different reaction speed between MAX and
BAX with PbX
2
(X = I, Br). Herein, we use the term 3D perovskite
crystallitesin the discussion of the abovementioned lms for sim-
plicity, even though we recognize that these particles possess
reduced dimensionality from n=. For loadings beyond these
threshold values, BAX is no longer playing the role of surfactant
400 500 600 700 800
0.0
0.2
0.4
0.6
0.8
1.0 I-perovskite
Absorbance
Wavelength (nm)
0:100
10:100
20:100
40:100
60:100
100:100
300 400 500
0.0
0.2
0.4
0.6
0.8 Br-perovskite
Absorbance
Wavelength (nm)
0:100
10:100
20:100
40:100
60:100
100:100
600 700 800 900
0.0
0.2
0.4
0.6
0.8
1.0
I-perovskite
Normalized PL intensity (a.u.)
Wavelength (nm)
Time delay (ns) Time delay (ns)
0:100
10:100
20:100
40:100
60:100
100:100
450 500 550 600
0.0
0.2
0.4
0.6
0.8
1.0
Br-perovskite
Normalized PL intensity (a.u.)
Wavelength (nm)
0:100
10:100
20:100
40:100
60:100
100:100
ab
cd
f
gh
e
MAPbI
3
Wavelength (nm)
700
800
750
650
MAPbBr
3
Wavelength (nm)
460
540
580
500
600
BAI:MAPbI
3 = 20:100
Wavelength (nm)
600
700
800
750
650
BABr:MAPbBr
3 = 20:100
Wavelength (nm)
02468100246810
Time delay (ns) Time delay (ns)
02468100246810
460
540
580
500
Figure 2 | Absorption and PL of I- and Br-perovskite lms. a,c, Absorption and PL spectra of I-perovskite lms with different molar ratios of BAI:MAPbI
3
in the precursor solution. b,d, Absorption and PL spectra of Br-perovskite lms with different molar ratios of BABr:MAPbBr
3
in the precursor solution.
eh, Time-resolved PL spectra of MAPbI
3
(e), MAPbBr
3
(f), BAI:MAPbI
3
(20:100) (g) and BABr:MAPbBr
3
(20:100) (h)lms.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.269 ARTICLES
NATURE PHOTONICS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturephotonics 3
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
and crystallite terminator, but rather participates in crystallization.
When this transition occurs, grain size begins to increase again
and lms become rougher. For instance, when the molar ratio is
further increased to 60:100 for BAI:MAPbI
3
or 40:100 for BABr:
MAPbBr
3
, layered perovskite diffraction peaks start to appear.
Ultimately, when the BAX ratio reaches 100:100, lms become
dominated by layered perovskite grains, which resemble the
BA
2
MA
n1
Pb
n
X
3n+1
perovskite, with n= 2 (Supplementary Fig. 2).
The absorption and PL spectra of I- and Br-perovskite lms with
different molar ratios of BAX:MAPbX
3
(X = I, Br) in the precursor
solution are shown in Fig. 2ad. With increased BAX loading, the
absorption edge blue-shifts, and excitonic absorption peaks corre-
sponding to layered perovskite grains emerge when the BAX ratio
is increased to 60:100 for I-perovskite or 40:100 for Br-perovskite,
in line with our observations from XRD as discussed above. In
turn, the PL peaks of BAX:MAPbX
3
lms gradually blue-shift with
increasing BAX content from 0:100 to 40:100. The PL quantum
yield (QY) concomitantly increases from 0.2 to 0.9, 1.9 and 6.6%
for I-perovskite, and from 0.2 to 0.5, 7.0 and 40.1% for Br-perovskite
lms with BAX molar ratio increasing from 0:100 to 10:100, 20:100
and 40:100, respectively. The time-resolved PL measurement showed
that the average decay time is increased from 8.8 ns (14.3 ns) to 11.4
ns (26.2 ns) with the BAI (BABr) molar ratio increasing from 0:100
to 20:100 (Fig. 2eh and Supplementary Table 1). The increased and
blue-shifted PL emission is due to the reduced 3D perovskite
crystallite dimensionality to n< 30 surrounded by BA cations, indu-
cing a disorder effect when the lattice periodicity is interrupted28.
When the BAX (X = I, Br) ratio is further increased to 60:100,
layered perovskite grains with excitonic absorption form, inducing
signicant blue-shifts of absorption and PL29, as shown in Fig. 2b,d.
The blue-shift accompanying the reduction in crystallite size is
also suggested by the evolution of the absorption and PL spectra
due to thermal annealing (Supplementary Fig. 3). Thermal anneal-
ing drives grain growth and increases grain size as conrmed by
both AFM (Supplementary Fig. 4) and XRD (Supplementary
Fig. 5) for both I- and Br-perovskite lms. The FWHM of the
(110) and (220) diffraction peaks for I-perovskite and (100) and
(200) diffraction peaks for Br-perovskite reduce after thermal
annealing for various molar ratios of BAX:MAPbX
3
in the
precursor solution.
Hybrid perovskite LEDs
The LED structures are: ITO (150 nm)/HTL (30 nm)/I-perovskite
(80 nm) (or Br-perovskite, 70 nm)/TPBi (60 nm)/LiF (1.2 nm)/Al
(100 nm) (ITO, indium tin oxide; HTL, hole-transporting layer;
TPBi, 2,2,2-(1,3,5-benzinetriyl)-tris(1-phenyl-1-H-benzimida-
zole)). This structure is shown schematically in Fig. 3a, along with
energy diagrams in Fig. 3b,c. Supplementary Fig. 6 shows cross-sec-
tional SEM images of I-perovskite devices, which conrm lm
thicknesses and the small grain size after BAI incorporation. The
Poly
TPD
ITO
I-perovskite
2.7
5.6
2.0
5.4
3.8
4.7
6.2
eV
eV
Br-perovskite
5.9
3.6
PVK
ITO
6.2
2.7
4.7
5.8
2.0
ITO
HTL
Al
LiF
V
BA cation
MAPbX3
a
bc
TPBi
LiF/Al
TPBi
LiF/Al
Perovskite
TPBi
Figure 3 | Struct ure and energy diagram of perovskite LEDs. a, Device structure of the perovskite LEDs. A schematic of a nanometre-sized grain with BA
cations decorating its surface is shown on the right. b,c, Energy diagram of I-perovskite (b) and Br-perovskite (c) LEDs. The energy levels of the 3D
perovskite (MAPbI
3
and MAPbBr
3
) and organic semiconductor layers are taken from literature37.
ARTICLES NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.269
NATURE PHOTONICS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturephotonics4
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
HTL utilized for I-perovskite-based LEDs was poly-TPD with a
highest occupied molecular orbital (HOMO) of 5.6 eV whereas
PVK with a deeper HOMO level of 5.8 eV was used for Br-perovs-
kite-based LEDs owing to better energy level alignment for hole
injection and electron blocking. The TPBi layer serves as an elec-
tron-injecting and hole-blocking layer for both I-perovskite and
Br-perovskite LEDs. It should be noted that the energy levels of
the perovskites may change as a result of the BAX incorporation.
However, this change is expected to be small (approximately 60
meV for I-perovskite and 80 meV for Br-perovskite calculated
from the PL peak shift) such that it has a minimal effect on
charge injection and blocking.
Figure 4a,d show the normalized EL spectra of I- and Br-perovs-
kite LEDs, respectively. Similar to the PL, EL is blue-shifted with
increasing molar ratio of BAX:MAPbX
3
(X = I, Br). The small
blue-shift of EL peak position compared with PL (solid lines) for
some devices may result from a distribution of crystallite sizes and
corresponding nvalues, with the smaller crystallites the more
easily electrically populated30. The EL of I-perovskite LEDs with
molar ratio of 100:100 and Br-perovskite LEDs with molar ratio
of 60:100 and 100:100 are too weak to be measured, possibly the
result of the increasing population of layered perovskite phases
that have been reported to have strong electronphonon coupling
that inhibits efcient EL31. The current densityvoltageluminance
(JVL) curves of I- and Br-perovskite LEDs and corresponding
EQE curves are shown in Fig. 4b,c and Fig. 4e,f, respectively. The
angular spectra and intensity proles are shown in Supplementary
Fig. 7. The EQE of LEDs without the addition of BAX are low, possi-
bly a result of larger grain size and consequently reduced electron
hole capture rate. With increasing molar ratio of BAX:MAPbX
3
to
20:100 in the precursor, the EQE of the device dramatically
increases. The highest EQEs reach 10.4% for 20:100 I-perovskite
and 9.3% for 20:100 Br-perovskite LEDs. Notably, the EQE of the
perovskite LEDs are relatively at over the range of the measure-
ment, in contrast to recent work showing that EQE can increase
substantially with current density20. We believe this to be primarily
due to the low leakage currents in our devices, at approximately
10
6
mA cm
2
, indicating very few parasitic leakage pathways exist
between electrodes. Also, the fact that the EQE versus current
density slope reduces with the addition of BAX indicates that,
even at low current density, non-radiative recombination is slower
than radiative bimolecular recombination, in agreement with the
increased QY and longer PL lifetimes of lms with BAX. It
should be noted that the EQE has a higher value than PLQY for
some devices, an aspect due to the nearly two-orders-of-magnitude
higher charge density associated to a current of 1020 mA cm
2
comparedwith the charge density generated by the optical generation
of the PLQY measurement. The PE and CE curves of both I- and
Br-perovskite LEDs are shown in Supplementary Fig. 8. The peak
PE and CE for Br-perovskite LEDs reach 13.0 lm W
1
at 1,200 cd m
2
and 17.1 cd A
1
at 2,900 cd m
2
, respectively, while PE and CE of
the I-perovskite devices are relatively low due to the near-infrared
light emission (Fig. 4a). While EQE reaches a maximum within
the range of 5.06.5 V, corresponding to 1550 mA cm
2
, there is
not a strong roll-off at higher voltages, although devices consistently
600 700 800
0.0
0.2
0.4
0.6
0.8
1.0
0:100
10:100
20:100
40:100
60:100
I-perovskite
Normalized EL/PL intensity (a.u.)
Wavelength (nm)
10
1
10
0
10
1
I-perovskite
EQE (%)
10 20 30 40 50
450 500 550
0.0
0.2
0.4
0.6
0.8
1.0 Br-perovskite
0:100
10:100
20:100
40:100
Normalized EL/PL intensity (a.u.)
Wavelength (nm)
10
2
10
1
10
0
10
1
Br-perovskite
EQE (%)
10 20 30 40 50
a c
d
b
ef
1356
10−7
10−5
10−3
10−1
101
103
0:100
10:100
20:100
40:100
60:100 0:100
10:100
20:100
40:100
60:100
I-perovskite
024
0
20
40
60
80
Luminance (cd m
2
)
Voltage (V)
0123456
10−7
10
−5
10
−3
10
−1
10
1
10
3
0:100
10:100
20:100
40:100
0:100
10:100
20:100
40:100
Br-perovskite
Current density (mA cm
2
)Current density (mA cm
2
)
Current density (mA cm
2
)
Current density (mA cm
2
)
Voltage (V)
2
4
6
8
Luminance (×1,000 cd m
2
)
Figure 4 | Perovskite LED performance characterization. ac, I-perovskite LEDs with different BAI:MAPbI
3
molar ratios, showing EL (open circles) and PL
(solid lines) spectra (a), JVLcurves (b) and EQE curves (c). It should be noted that BAI:MAPbI
3
(60:100) LED has higher luminance than 20:100 due to
the EL spectrum that has shifted closer to peak eye response. df, Br-perovskite LEDs with different BABr:MAPbBr
3
molar ratios, showing EL (open circles)
and PL (solid lines) spectra (d), JVLcurves (e) and EQE curves (f). The scanning rate for all devices is 0.1 V s
1
.
Table 1 | Perovskite LED performance metrics.
BAX:MAPbX
3
Peak PL/
EL
position
(nm)
PE (lm W
1
)
(average/
max.)
EQE (%)
(average/
max.)
CE (cd A
1
)
(average/
max.)
BAI:MAPbI
3
0:100 763/763 0.01/0.02 0.8/1.0 0.02/0.03
10:100 757/756 0.05/0.06 4.3/5.2 0.05/0.06
20:100 750/748 0.08/0.10 9.7/10.4 0.08/0.09
40:100 737/732 0.02/0.03 1.8/2.4 0.02/0.03
60:100 708/695 0.13/0.18 1.7/2.1 0.17/0.22
BABr:MAPbBr
3
0:100 526/522 0.03/0.04 0.02/0.03 0.06/0.09
10:100 520/519 3.5/4.3 1.8/2.4 4.2/4.9
20:100 516/513 11.2/13.0 8.8/9.3 15.8/17.1
40:100 509/503 3.5/4.0 4.2/4.5 3.9/4.5
The device statistics are based on ten devices of each composition from two batches.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.269 ARTICLES
NATURE PHOTONICS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturephotonics 5
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
fail at voltages of 810 V (0.21Acm
2
) due to Joule heating
(Supplementary Fig. 9). For either the I- or Br-perovskite LEDs,
when the BAX ratio increases beyond 40:100, the device perform-
ance decreases. Detailed performance metrics of both I- and
Br-perovskite devices are provided in Table 1.
It has recently been reported that excess MABr in MAPbBr
3
can
reduce the content of metallic lead and increase the EQE of
MAPbBr
3
-based LEDs19. As a comparison, we fabricated both
I- and Br-perovskite LEDs with MAX:MAPbX
3
(X = I, Br) molar
ratio of 5:100 and 20:100 (MAX:PbX
2
molar ratio of 105:100 and
120:100), with EQE curves shown in Supplementary Fig. 10.
While excess MAX is effective in increasing EQE for both I- and
Br-perovskite LEDs (to 3.6 and 2.1%, respectively), we nd that
the incorporation of bulky ammonium halide ligands provide
greater efciency gains because they are effective in allowing for
smaller grain size, and thin lms with low roughness. We note
that excess MAX can also effectively decrease crystallite size by
forming many stacking faults within a grain32. A high density of
stacking faults can serve the same crystal termination effect as the
long-chain alkylammonium halides used herein, although less ef-
ciently. As further evidence of the versatility of this approach,
another type of large-group cation, phenylethylammonium iodide
−1 0 1 2 3 4 5
10
−8
10
−6
10
−4
10
−2
10
0
10
2
10
−6
10
−4
10
−2
10
0
10
2
I-perovskite
Current density (mA cm
−2
)
Voltage (V)
0:100
Forward scan
Reverse scan
20:100
Forward scan
Reverse scan
10 1 2 3 4 5
10
4
10
3
10
2
10
1
10
0
10
1
10
2
I-perovskite
0:100
Forward scan
Reverse scan
20:100
Forward scan
Reverse scan
Luminance (cd m
−2
)
Voltage (V)
345
10
1
10
0
10
1
I-perovskite
0:100Forward scan
Reverse scan
20:100
Forward scan
Reverse scan
EQE (%)
Voltage (V)
Br-perovskite
0:100
Forward scan
Reverse scan
20:100
Forward scan
Reverse scan
Current density (mA cm
−2
)
Voltage (V)
10 1 2 3 4 5 6
10 1 2 3 4 5 6
10
1
10
0
10
1
10
2
10
3
Br-perovskite
0:100
Forward scan
Reverse scan
20:100
Forward scan
Reverse scan
Luminance (cd m
−2
)
Voltage (V)
3.5 4.0 4.5 5.0 5.5
10
3
10
2
10
1
10
0
10
1
Br-perovskite
0:100
Forward scan
Reverse scan
20:100
Forward scan
Reverse scan
EQE (%)
Voltage (V)
0.6
0102030405060708090
0.8
1.0
1.2 I-perovskite
Normalized EQE (a.u.)
0:100
20:100
0.4
0.2
0.6
0.8
1.0
1.2
1.4 Br-perovskite
Normalized EQE (a.u.)
0:100
20:100
abc
def
gh
ij
0.1 1 10 100
10
4
10
3
10
2
10
1
10
0
10
1
Br-perovskite
0:100
20:100
EQE (%)
Time (d)
0.1 1 10 100
10
2
10
1
10
0
10
1
I-perovskite
0:100
20:100
EQE (%)
Time (d)
Time (s)
0102030405060708090
Time (s)
Figure 5 | Hysteresis and steady-state output of perovskite LEDs. ac, Current density (a), luminance (b)andEQE(c) of I-perovskite LEDs with BAI:MAPbI
3
molar ratios of 0:100 and 20:100. df, Current density (d), luminance (e)andEQE(f) of Br-perovskite LEDs with BABr:MAPbBr
3
molar ratios of 0:100 and
20:100. The scanning rate is 0.1 V s
1
and scanning range is 0.5 V to +5.5 V for all devices. g,h, Steady-state EQE measurements of I-perovskite (g)and
Br-perovskite (h) LEDs at a constant voltage of 5.0 V. The EQE of the LEDs are normalized to their initial value. i,j, Shelf stability of unencapsulated
I-perovskite (i) and Br-perovskite (j) LEDs as a function of storage time in N
2
.
ARTICLES NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.269
NATURE PHOTONICS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturephotonics6
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
(PEAI), was incorporated with PEAI:MAPbI
3
molar ratio of 20:100.
As shown in Supplementary Fig. 10, the highest EQE reaches 9.6%
after the addition of PEAI in the MAPbI
3
precursor solution.
Recently, efcient I-perovskite LEDs using stoichiometric
PEA
2
MA
n1
Pb
n
X
3n+1
solutions with EQE up to 8.8% for n=5
have been reported33. The non-stoichiometric solution in our
work resulted in higher EQE that might be due to the passivation
effect of the extra organic cations.
Hysteresis and stability of hybrid perovskite LEDs
The photocurrent hysteresis of perovskite solar cells is thought to be
due to ion migration and associated traps3436, an aspect that com-
plicates accurate efciency measurement. Hysteresis has also been
previously reported to exist in the JVLresponse of perovskite
LEDs18, and can lead to inaccurate performance characterization
due to its inuence on the steady-state EL emission and angular
prole measurements. And because ion migration primarily
occurs along grain boundaries34, we have investigated the inuence
of BAX additives on LED hysteresis. By measuring the JVL
characteristics with various scanning directions and rates, as shown
in Fig. 5af, it can be seen that all devices show some level of hysteresis
for both Jand L. Nevertheless, the devices with BAX:MAPbX
3
(X = I, Br) molar ratio of 20:100 exhibit reduced hysteresis.
Supplementary Fig. 11 shows the scanning-rate-dependent EQE for
both I- and Br-perovskite LEDs, demonstrating that LEDs without
BAX are more scanning-rate dependent. We believe that the long-
chain BA cations that self-assemble at the crystallite surface either
impede ion motion or prevent ions from within crystallites from
participating in ion migration processes. In essence, the BA cation,
being larger than MA, decreases the cationsthermodynamicactivity
at the surface (that is, by reducing the equilibrium vapour pressure
at the surface). To ensure that devices are able to be reliably character-
ized, the steady-state EQE was measured at a constant voltage of 5.0 V.
As shown in Fig. 5g,h, the EQE of the device without BAX showed a
strong poling effect, while the devices with BAX rapidly stabilize. The
initial increase of the EQE in lms without BAX might result from
trap lling such that injected charges have a higher radiative recombi-
nation rate, while the decrease of the EQE after a few seconds might be
due to ion-migration-induced degradation of the perovskite layer.
Stability is of critical importance for perovskite LEDs, and thus
the shelf stability of unencapsulated LEDs stored in N
2
were evalu-
ated over time. As shown in Fig. 5i,j, the EQE of the control devices
without BAX in the precursor solution decreases to 6070% of their
initial value after 2 days, whereas the EQE of an I-perovskite LED
with BAI:MAPbI
3
molar ratio of 20:100 is stable without degra-
dation after storage for more than 8 months. Similarly, the
Br-perovskite LEDs also show no degradation after storage for
4 months. The improved stability of the perovskite LEDs with
BAX can be attributed to the ultra-smooth, pinhole-free and
compact perovskite lms as well as the bulky nature of the long
alkyl chains that stabilize the crystallite surfaces. The stabilized
nanocrystallite perovskite lms also enable improved operational
stability, shown in Supplementary Fig. 12, an aspect that is critical
for display and lighting applications. One possible reason for the
operational instability is that ion migration induces degradation
and sites for non-radiative recombination in the perovskite layer
as the LEDs work at high voltages. This is conrmed by the
decreased PL intensity of the LED after operating at 5 V for 5 min
(Supplementary Fig. 13). Another possible reason is thermal degra-
dation caused by Joule heating. We thus propose three approaches
to continue to improve the operational stability of perovskite LEDs:
(1) use more bulky organic ligands to stabilize the surface and
impede ion migration; (2) decrease turn-on voltage through further
interface engineering such that devices operate at lower voltages;
and (3) at least partially replace MA cations with formamidinium
ammonium or Cs to increase thermal stability.
In conclusion, we have shown that the addition of long-chain
ammonium halides in the precursor solution can dramatically
impede the grain growth of 3D perovskite crystallites and decrease
lm roughness. The nanometre-sized 3D perovskite grains induce
enhanced and blue-shifted PL and EL emission, with the long-
chain ammonium cations stabilizing the grain surface. As a result,
perovskite LEDs made from BAX-incorporated layers showed
greatly improved device performance, shelf stability and operational
stability. It is also worth investigating whether the incorporation of
long-chain ammonium halide ligands within perovskite lms can
benet other devices such as photodetectors, solar cells and lasers.
Methods
Methods and any associated references are available in the online
version of the paper.
Received 11 April 2016; accepted 6 December 2016;
published online 16 January 2017
References
1. Stoumpos, C. C., Malliakas, C. D. & Kanatzidis, M. G. Semiconducting tin
and lead iodide perovskites with organic cations: phase transitions, high
mobilities, and near-infrared photoluminescent properties. Inorg. Chem. 52,
90199038 (2013).
2. Xing, G. et al. Low-temperature solution-processed wavelength-tunable
perovskites for lasing. Nat. Mater. 13, 476480 (2014).
3. Miyata, A. et al. Direct measurement of the exciton binding energy and effective
masses for charge carriers in organicinorganic tri-halide perovskites. Nat. Phys.
11, 582587 (2015).
4. Noriko, O.-Y., Takasuke, M. & Hiroshi, S. Dielectric study of CH
3
NH
3
PbX
3
(X = Cl, Br, I). J. Phys. Chem. Solids 53, 935939 (1992).
5. Yin, W.-J., Shi, T. & Yan, Y. Unusual defect physics in CH
3
NH
3
PbI
3
perovskite
solar cell absorber. Appl. Phys. Lett. 104, 063903 (2014).
6. Kojima, A., Teshima, K., Shirai, Y. & Miyasaka, T. Organometal halide
perovskites as visible-light sensitizers for photovoltaic cells. J. Am. Chem. Soc.
131, 60506051 (2009).
7. Burschka, J. et al. Sequential deposition as a route to high-performance
perovskite-sensitized solar cells. Nature 499, 316319 (2013).
8. Liu, M., Johnston, M. B. & Snaith, H. J. Efcient planar heterojunction
perovskite solar cells by vapour deposition. Nature 501, 395398 (2013).
9. Lee, M. M., Teuscher, J., Miyasaka, T., Murakami, T. N. & Snaith, H. J. Efcient
hybrid solar cells based on meso-superstructured organometal halide
perovskites. Science 338, 643647 (2012).
10. Mei, A. et al. A hole-conductorfree, fully printable mesoscopic perovskite solar
cell with high stability. Science 345, 295298 (2014).
11. Yang, W. S. et al. High-performance photovoltaic perovskite layers fabricated
through intramolecular exchange. Science 348, 12341237 (2015).
12. Zhou, H. et al. Interface engineering of highly efcient perovskite solar cells.
Science 345, 542546 (2014).
13. Xiao, Z. et al.Efcient, high yield perovskite photovoltaic devices grown by
interdiffusion of solution-processed precursor stacking layers. Energy Environ.
Sci. 7, 26192623 (2014).
14. Liu, D. & Kelly, T. L. Perovskite solar cells with a planar heterojunction structure
prepared using room-temperature solution processing techniques. Nat. Photon.
8, 133138 (2014).
15. Jeon, N. J. et al. Solvent engineering for high-performance inorganicorganic
hybrid perovskite solar cells. Nat. Mater. 13, 897903 (2014).
16. Im, J.-H., Jang, I.-H., Pellet, N., Grätzel, M. & Park, N.-G. Growth of
CH
3
NH
3
PbI
3
cuboids with controlled size for high-efciency perovskite solar
cells. Nat. Nanotech. 9, 927932 (2014).
17. Kim, H.-S. et al. Lead iodide perovskite sensitized all-solid-state submicron
thin lm mesoscopic solar cell with efciency exceeding 9%. Sci. Rep. 2,
591 (2012).
18. Tan, Z.-K. et al. Bright light-emitting diodes based on organometal halide
perovskite. Nat. Nanotech. 9, 687692 (2014).
19. Cho, H. et al. Overcoming the electroluminescence efciency limitations of
perovskite light-emitting diodes. Science 350, 12221225 (2015).
20. Li, G. et al.Efcient light-emitting diodes based on nanocrystalline perovskite in
a dielectric polymer matrix. Nano Lett. 15, 26402644 (2015).
21. Sadhanala, A. et al. Blue-green color tunable solution processable organolead
chloridebromide mixed halide perovskites for optoelectronic applications.
Nano Lett. 15, 60956101 (2015).
22. Gong, X. et al. Highly efcient quantum dot near-infrared light-emitting diodes.
Nat. Photon. 10, 253257 (2016).
23. Wang, Q. et al. Large ll-factor bilayer iodine perovskite solar cells fabricated by
a low-temperature solution-process. Energy Environ. Sci. 7, 23592365 (2014).
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.269 ARTICLES
NATURE PHOTONICS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturephotonics 7
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
24. Eperon, G. E., Burlakov, V. M., Docampo, P., Goriely, A. & Snaith, H. J.
Morphological control for high performance, solution-processed planar
heterojunction perovskite solar cells. Adv. Funct. Mater. 24, 151157 (2014).
25. Xiao, M. et al. A fast deposition-crystallization procedure for highly efcient lead
iodide perovskite thin-lm solar cells. Angew. Chem. 126, 1005610061 (2014).
26. Kerner, R. A., Zhao, L., Xiao, Z. & Rand, B. P. Ultrasmooth metal halide perovskite
thin lms via solgel processing. J. Mater. Chem. A 4, 83088315 (2016).
27. Mitzi, D., Wang, S., Feild, C., Chess, C. & Guloy, A. Conducting layered organic-
inorganic halides containing <110>-oriented perovskite sheets. Science 267,
14731476 (1995).
28. Scholes, G. D. & Rumbles, G. Excitons in nanoscale systems. Nat. Mater. 5,
683696 (2006).
29. Mitzi, D. B., Chondroudis, K. & Kagan, C. R. Organic-inorganic electronics. IBM
J. Res. Dev. 45, 2945 (2001).
30. Cheng, K.-Y., Anthony, R., Kortshagen, U. R. & Holmes, R. J. High-efciency
silicon nanocrystal light-emitting devices. Nano Lett. 11, 19521956 (2011).
31. Gauthron, K. et al. Optical spectroscopy of two-dimensional layered
(C
6
H
5
C
2
H
4
-NH
3
)
2
-PbI
4
perovskite. Opt. Express 18, 59125919 (2010).
32. Song, Z. et al. Impact of processing temperature and composition on the
formation of methylammonium lead iodide perovskites. Chem. Mater. 27,
46124619 (2015).
33. Yuan, M. et al. Perovskite energy funnels for efcient light-emitting diodes. Nat.
Nanotech. 11, 872877 (2016).
34. Xiao, Z. et al. Giant switchable photovoltaic effect in organometal trihalide
perovskite devices. Nat. Mater. 14, 193198 (2015).
35. Shao, Y., Xiao, Z., Bi, C., Yuan, Y. & Huang, J. Origin and elimination of
photocurrent hysteresis by fullerene passivation in CH
3
NH
3
PbI
3
planar
heterojunction solar cells. Nat. Commun. 5, 5784 (2014).
36. van Reenen, S., Kemerink, M. & Snaith, H. J. Modeling anomalous hysteresis in
perovskite solar cells. J. Phys. Chem. Lett. 6, 38083814 (2015).
37. Schulz, P. et al. Interface energetics in organo-metal halide perovskite-based
photovoltaic cells. Energy Environ. Sci. 7, 13771381 (2014).
Acknowledgements
We acknowledge a DARPA Young Faculty Award (award no. D15AP00093) for research
funding. B.P.R. acknowledges the support of a DuPont Young Professor Awardfor research
funding. R.A.K. acknowledges support from the National Science Foundation Graduate
Research Fellowship under grant no. DGE 1148900.
Author contributions
T.-W.K. and Z.X. designed the device structure. Z.X. performed the AFM, PL and
absorption measurements, and fabricated the LEDs. R.A.K. developed the surfacted
perovskite processing protocol, synthesized the precursors and helped to calculate grain
size. L.Z. conducted the XRD and SEM measurements. N.L.T. and G.D.S conducted the
TRPL and QY measurements. K.M.L. assisted with LED characterization.B.P.R. supervised
the work. Z.X. and B.P.R. wrote the manuscript. All authors read and commented on
the manuscript.
Additional information
Supplementary information is available in the online version of the paper.
Reprints and
permissions information is available online at www.nature.com/reprints. Correspondence and
requests for materials should be addressed to B.P.R.
Competing nancial interests
The authors declare no competing nancial interests.
ARTICLES NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.269
NATURE PHOTONICS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturephotonics8
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Methods
Materials synthesis. MAI, BAI and PEAI were synthesized by mixing
methylamine, butylamine and phenylethylamine (Sigma Aldrich) with equimolar
amounts of aqueous HI (Sigma Aldrich, stabilized) at 0 °C with constant stirring
under N
2
. MABr and BABr were synthesized with analogous procedures. The
organoammonium halides (MAI, BAI, PEAI, MABr and BABr) were washed
with an ethanol:ether mixture and rotovaped several times to remove the
HI stabilizer.
Perovskite lm deposition and characterization. Poly-TPD (or PVK) was
dissolved in chlorobenzene at a concentration of 6 mg ml
1
. Poly-TPD (or PVK) was
spin-coated on ITO at 1,000 r.p.m. for 60 s, followed by thermal annealing at 150°C
(or 120 °C for PVK) for 20 min. The poly-TPD layer was treated by O
2
plasma for 2 s
to improve wetting. The MAPbI
3
or MAPbBr
3
precursors were dissolved in DMF
(Sigma Aldrich, 99.8% anhydrous) at concentrations of 0.43 M and 0.3 M,
respectively. Then, different molar amounts of BAI and BABr were added into the
MAPbI
3
and MAPbBr
3
solutions, respectively, before spin-coating. The spin-coating
rate for the I-perovskite was 6,000 r.p.m., and toluene was dropped on the spinning
substrate at 5 s. The spin speed for the Br-perovskite was 4,500 r.p.m., and the
toluene was dropped on the spinning substrate at 4 s.
The AFM measurements were conducted in the tapping mode in a N
2
lled
glove box (Veeco di Innova). The SEM measurements were conducted with an FEI
XHR (extreme high resolution) SEM (Verios 460) using immersion mode. XRD
measurements were performed with a Bruker D8 Discover X-ray diffractometer with
Bragg-Brentano parallel beam geometry, a diffracted beam monochromator and a
conventional Cu target X-ray tube set to 40 kV and 40 mA.Absorption spectra were
measured using a Cary 5000 UV-Vis-NIR system (Agilent). The PL spectra were
measured using an FLS980 spectrometer (Edinburgh Instruments) with an
excitation wavelength of 380 nm.
PLQY measurement and calculation. The QY of the I- and Br-perovskite lms was
measured using a PTI QuantaMaster 400 Steady State Fluorometer with Petite
Integrating Sphere. The excitation wavelengths for I- and Br-perovskite are 450 and
400 nm, respectively, with excitation intensities of 2.2 and 2.1 mW cm
2
, respectively.
Time-resolved PL measurement. Time-resolved PL measurements were taken using
a Horiba DeltaFlex time-correlated single-photon counting system. The samples
were excited by a pulsed laser diode (DeltaDiode-Horiba) with a centre wavelength
of 406 nm, an excitation intensity of 4mWcm
2
and a repetition rate that is less
than the reciprocal of the measurement range. The time resolution was determined
to be 100 ps from the instrument response function. All decay traces were taken at
5 nm increment and with constant run time of 10 s. The emission bandpass was
16 nm for all samples except for BABr:MAPbBr
3
(20:100 and 40:100) samples
where it was set to 1 nm to protect the detector from oversaturation.
Device fabrication and characterization. All perovskite lms were dried at 70 °C for
5 min following spin-coating to fully dry the lm and ensure the full reaction of the
precursors without affecting the morphology and grain size substantially. TPBi, LiF
and Al layers were sequentially thermally deposited on top of the perovskite lm to
thicknesses of 60, 1.2 and 100 nm, respectively. The device area was 0.1 cm
2
.
LEDs were measured in N
2
using a homemade motorized goniometer set-up
consisting of a Keithley 2400 sourcemeter unit, a calibrated Si photodiode (FDS-100-
CAL, Thorlabs), a picoammeter (4140B, Agilent) and a calibrated ber optic
spectrophotometer (UVN-SR, StellarNet).
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.269 ARTICLES
NATURE PHOTONICS |www.nature.com/naturephotonics
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
... Consequently, several strategies have been developed to enhance the film quality using spin-coating. For instance, Rand et al. [167] investigated the effect of adding ammonium halides to perovskites by comparing the ratio of the components ( Figure 12a). They found that the surfactant restricted the grain growth, resulting in reduced roughness and a notable 10.4% EQE of LED device. ...
... In contrast to 3D perovskites, the charge-carrier recombination mechanism changes when the perovskite crystal size approaches or is smaller than the exciton Bohr radius, leading to the formation of excitons and emergence of radiative monomolecular recombination [333]. In addition, the amount of the bulky ammonium halides should be carefully controlled (approximately 20 mol% to lead halides); higher loading of the halides could lead to the formation of quasi-2D perovskites [167]. The use of non-ammonium molecules or polymers can be another effective way to constrain the perovskite grain size to tens of nanometers without forming quasi-2D perovskites [334]. ...
... It was believed that nanoscale perovskite grains can help to enhance the spatial confinement of the injected charge carriers and improve the radiative bimolecular recombination rate of 3D perovskites while the dense and uniform morphology is to reduce the leakage current in PeLEDs. Various methods, such as modifying film growth interfaces [335], engineering precursor stoichiometry [321] as well as introducing polymer [334] and organic ammonium halide [167] matrixes, have been developed for the fabrication of dense and uniform perovskite emissive layers, enabling PeLEDs with improved EQEs. ...
Article
Metal-halide perovskites have garnered considerable research attention as highly efficient light emitters in recent years due to their outstanding optoelectronic properties with remarkable tunability and excellent solution processabilities. Substantial advancements have been achieved in the development of novel halide perovskites, and the exploitations of these materials in light-emitting devices. This review comprehensively outlines recent breakthroughs in metal-halide perovskites, encompassing the rational design of perovskite materials with tunable light emission properties, the controllable growth of single crystal for a deeper understanding of their structure–property relationships, as well as the fundamental insights into the photophysics and carrier dynamics in perovskite systems. Additionally, it provides an overview of recent applications of perovskite materials in high-performance light-emitting diodes (LEDs) and lasers.
... Hybrid organic-inorganic perovskites have garnered significant attention in recent years owing to their remarkable properties, positioning them as one of the most intriguing optoelectronic materials [9,[12][13][14][15][16][17][18][19][20] for potential device applications [21]. Remarkable performance across various applications spans from light-emitting diodes (LEDs) and lasers to highly efficient solar cells [22]. These devices leverage the distinctive structure of metal halide perovskites, with MAPbI 3 serving as a prime example. ...
Article
Full-text available
Entropy transfer in metal halide perovskites, characterized by significant lattice anharmonicity and low stiffness, underlies the remarkable properties observed in their optoelectronic applications, ranging from solar cells to lasers. The conventional view of this transfer involves stochastic processes occurring within a thermal bath of phonons, where the lattice arrangement and energy flow from higher- to lower-frequency modes. Here, we unveil a comprehensive chronological sequence detailing a conceptually distinct coherent transfer of entropy in a prototypical perovskite CH 3 NH 3 Pbl 3 . The terahertz periodic modulation imposes vibrational coherence into electronic states, leading to the emergence of mixed (vibronic) quantum beat between approximately 3 and 0.3 THz. We highlight a well-structured bidirectional time-frequency transfer of these diverse phonon modes, each developing at different times and transitioning from high to low frequencies from 3 to 0.3 THz, before reversing direction and ascending to around 0.8 THz. First-principles molecular dynamics simulations disentangle a complex web of coherent-phononic coupling pathways and identify the salient roles of the initial modes in shaping entropy evolution at later stages. Capitalizing on coherent entropy transfer and dynamic anharmonicity presents a compelling opportunity to exceed the fundamental thermodynamic (Shockley-Queisser) limit of photoconversion efficiency and to pioneer novel optoelectronic functionalities. Published by the American Physical Society 2024
... By contrast, the models show a high probability for the formation of a head or double-end connection, suggesting that either can exist on the perovskite surface. As has been widely reported 32,33 , the head connection mode breaks the originally continuous three-dimensional CsPbI 3 structure and leads to reduced-dimensional perovskites. ...
Article
Full-text available
Light-emitting diodes (LEDs) based on metal halide perovskites (PeLEDs) with high colour quality and facile solution processing are promising candidates for full-colour and high-definition displays1–4. Despite the great success achieved in green PeLEDs with lead bromide perovskites⁵, it is still challenging to realize pure-red (620–650 nm) LEDs using iodine-based counterparts, as they are constrained by the low intrinsic bandgap⁶. Here we report efficient and colour-stable PeLEDs across the entire pure-red region, with a peak external quantum efficiency reaching 28.7% at 638 nm, enabled by incorporating a double-end anchored ligand molecule into pure-iodine perovskites. We demonstrate that a key function of the organic intercalating cation is to stabilize the lead iodine octahedron through coordination with exposed lead ions and enhanced hydrogen bonding with iodine. The molecule synergistically facilitates spectral modulation, promotes charge transfer between perovskite quantum wells and reduces iodine migration under electrical bias. We realize continuously tunable emission wavelengths for iodine-based perovskite films with suppressed energy loss due to the decrease in bond energy of lead iodine in ionic perovskites as the bandgap increases. Importantly, the resultant devices show outstanding spectral stability and a half-lifetime of more than 7,600 min at an initial luminance of 100 cd m⁻².
... Metal halide perovskites have garnered significant interest from a diverse array of researchers and industrial businesses worldwide in recent years. Solar cells, catalysts, light-emitting diodes (LEDs), lasers, X-ray detectors, photodetectors, and field-effect transistors are only a few of the technological and commercial uses for them [1][2][3][4][5][6][7][8][9][10] . The development of LEDs and luminous bodies with light pumping in the form of luminescent materials has also made halide perovskites popular and advised [8][9][10] . ...
Article
Full-text available
In this work, the structural stability of cubic (α), tetragonal (β), and orthorhombic (γ) phases of perovskite CsSnI 3 and their phase transitions have been studied by density functional theory (DFT) calculations, and the results obtained have been compared with the characteristics of nonperovskite orthorhombic (δ) modification of CsSnI 3 compounds. The relaxed structures of the CsSnI 3 phases were produced and their geometrical properties were assessed using the strictly constrained normalized potential (SCAN) functional. According to the results, the energetic hierarchy of CsSnI 3 polymorphs is E β >E α >E γ >E δ. The phonon and thermodynamic characteristics as well as the temperatures of phase transitions of CsSnI 3 have been estimated using the Phonopy tool based on SCAN relaxed structures. The nature of the change in the total energy of the four phases of CsSnI 3 from VASP package calculations justifies the trends of free energy, entropy, enthalpy, and heat capacity. In contrast, the β-phase, which has the highest energy among the perovskite phases, is extremely unstable. It was discovered that the tetragonal phase becomes stable at 450K and transitions to the cubic phase at lowering temperatures. CsSnI 3 undergoes a phase transition between γ-and β-phases at 300-320K. At temperatures below 320K, a black-yellow transformation of CsSnI 3 occurs, in which the black perovskite transforms into a yellow non-perovskite conformation. It was found that temperature phase transitions occur between two orthorhombic phases of CsSnI 3 at 360K, although direct transitions of the α⟷γ and γ⟷δ types have not yet been reported in the literature, with the exception of γ→δ transitions under the influence of moisture.
Article
Full-text available
Cesium-based lead-free double perovskite materials (Cs2AgBiBr6) have garnered significant attention in the X-ray detection field due to their environment friendly characteristics. However, their substantial ion migration properties lead to large dark currents and detection limits in Cs2AgBiBr6-based X-ray detectors, restricting the detection performance of the device. In terms of process technology, ultrasonic spraying is more suitable than a spin-coating method for fabricating large-area, micron-scale perovskite thick films, with higher cost-effectiveness, which is crucial for X-ray detection. This work introduces a BA⁺ (BA⁺ = CH3CH2CH2CH2NH3⁺, n-butyl) source into the precursor solution and employs ultrasonic spraying to fabricate quasi-two-dimensional structured polycrystalline (BA)2Cs9Ag5Bi5Br31 perovskite thick films, developing a low-cost, eco-friendly X-ray detector with low dark current density and low detection limit. Characterization results reveal that the ion migration activation energy of (BA)2Cs9Ag5Bi5Br31 reaches 419 meV, approximately 17% higher than that of traditional three-dimensional perovskites, effectively suppressing perovskite ion migration and subsequently reducing the dark current. The (BA)2Cs9Ag5Bi5Br31-based X-ray detectors exhibit high resistivity (about 1.75 × 10¹⁰ Ω cm), low dark current density (66 nA cm⁻²), minimal dark current drift (0.016 pA cm⁻¹ s⁻¹ V⁻¹), and detection limit (138 nGyair s⁻¹), holding considerable promise for applications in low-noise, low-dose X-ray detection.
Article
Full-text available
Perovskites exhibit appealing optical and electrical properties, making them attractive candidates for efficient luminescent materials with low‐cost and straightforward fabrication processes. Their versatility is highlighted by the ability to deposit perovskite thin films on various substrates, including silicon, glass, sapphire, and flexible substrates, enabling potential monolithic integration on silicon for applications such as photonic integrated circuits, high‐speed communication. Extensive studies on perovskite light‐emitting diodes have shown external quantum efficiencies exceeding 20% across a wide spectral range from deep blue to near‐infrared, with chirality. Additionally, perovskite‐based lasing action has been achieved under pulsed optical excitation and continuous‐wave operation, as well as in functional diode structures. However, realizing electrically driven perovskite laser diodes for practical applications requires the injection of intense current densities. This review provides a comprehensive overview of the historical progress in perovskite lasers and light‐emitting didoes, along with important design considerations essential for their development.
Article
This review summaries the research progress of perovskite materials in X-ray detectors and provides ideas for the development of more environmentally friendly and higher-performance X-ray detectors.
Article
Halide perovskites have demonstrated great potential in x-ray detectors, due to their high x-ray attenuation coefficient, large bulk resistance, ultralong carrier diffusion length, and adjustable bandgap. Moreover, their abundant raw materials and simple processing combined with excellent compatibility with integrated circuits make them ideal for cost-efficient and high-efficiency real-world imaging applications. Herein, we comprehensively reviewed advances and progress in x-ray detection devices based on halide perovskites. We expound on the fundamental mechanisms of interactions between x rays and matter as background and indicate different parameters for different types of x-ray detectors, which guides the basic requirements on how to select and design suitable materials for active layers. After emphasizing the superb properties of halide perovskites through the shortcomings of commercial materials, we evaluate the latest advancements and ongoing progress in halide perovskites with different dimensions and structures for both direct and indirect x-ray detectors, and discuss the effect of dimensional varieties on the device performance. We also highlight current challenges in the area of perovskite x-ray detectors and propose corresponding solutions to optimize halide perovskites and optimize x-ray detectors for next-generation imaging applications.
Article
Full-text available
Organic-inorganic hybrid perovskites are emerging low-cost emitters with very high color purity, but their low luminescent efficiency is a critical drawback. We boosted the current efficiency (CE) of perovskite light-emitting diodes with a simple bilayer structure to 42.9 candela per ampere, similar to the CE of phosphorescent organic light-emitting diodes, with two modifications: We prevented the formation of metallic lead (Pb) atoms that cause strong exciton quenching through a small increase in methylammonium bromide (MABr) molar proportion, and we spatially confined the exciton in uniform MAPbBr3 nanograins (average diameter = 99.7 nanometers) formed by a nanocrystal pinning process and concomitant reduction of exciton diffusion length to 67 nanometers. These changes caused substantial increases in steady-state photoluminescence intensity and efficiency of MAPbBr3 nanograin layers.
Article
Full-text available
Solution processed organo-lead halide perovskites are produced with sharp, colour-pure electroluminescence that can be tuned from blue to green region of visible spectrum (425nm to 570 nm). This was accomplished by controlling the halide composition of CH3NH3Pb(BrxCl1-x)3 [0 ≤ x ≤ 1] perovskites. The bandgap and lattice parameters change monotonically with composition. The films possess remarkably sharp band edges and a clean bandgap, with a single optically active phase. These chloride-bromide perovskites can potentially be used in optoelectronic devices like solar cells and light emitting diodes (LEDs). Here we demonstrate high colour-purity, tunable LEDs with narrow emission full width at half maxima (FWHM) and low turn on voltages using thin-films of these perovskite materials, including a blue CH3NH3PbCl3 perovskite LED with a narrow emission FWHM of 5 nm.
Article
Full-text available
Solar cells based on the organic-inorganic tri-halide perovskite family of materials have shown remarkable progress recently, offering the prospect of low-cost solar energy from devices that are very simple to process. Fundamental to understanding the operation of these devices is the exciton binding energy, which has proved both difficult to measure directly and controversial. We demonstrate that by using very high magnetic fields it is possible to make an accurate and direct spectroscopic measurement of the exciton binding energy, which we find to be only 16 meV at low temperatures, over three times smaller than has been previously assumed. In the room temperature phase we show that the binding energy falls to even smaller values of only a few millielectronvolts, which explains their excellent device performance due to spontaneous free carrier generation following light absorption. Additionally, we determine the excitonic reduced effective mass to be 0.104me (where me is the electron mass), significantly smaller than previously estimated experimentally but in good agreement with recent calculations. Our work provides crucial information about the photophysics of these materials, which will in turn allow improved optoelectronic device operation and better understanding of their electronic properties.
Article
Organometal halide perovskites exhibit large bulk crystal domain sizes, rare traps, excellent mobilities and carriers that are free at room temperature-properties that support their excellent performance in charge-separating devices. In devices that rely on the forward injection of electrons and holes, such as light-emitting diodes (LEDs), excellent mobilities contribute to the efficient capture of non-equilibrium charge carriers by rare non-radiative centres. Moreover, the lack of bound excitons weakens the competition of desired radiative (over undesired non-radiative) recombination. Here we report a perovskite mixed material comprising a series of differently quantum-size-tuned grains that funnels photoexcitations to the lowest-bandgap light-emitter in the mixture. The materials function as charge carrier concentrators, ensuring that radiative recombination successfully outcompetes trapping and hence non-radiative recombination. We use the new material to build devices that exhibit an external quantum efficiency (EQE) of 8.8% and a radiance of 80 W sr(-1) m(-2). These represent the brightest and most efficient solution-processed near-infrared LEDs to date.
Article
We demonstrate that lead halide perovskite thin film formation displays the characteristics of a sol–gel process. By performing a solvent exchange at different times, the stages of the sol–gel process are elucidated and their sensitivity to processing conditions are examined. For CH3NH3PbI3, the reaction and aging kinetics are found to be extremely rapid, complete within 10 s, and a competing formation of a highly crystalline PbI2:N,N-dimethylformamide (DMF) complex introduces additional complications relative to a well-behaved sol–gel process. Perovskite formation from strongly polar solvents can be described, in most cases, by the sol–gel processing of PbI2 regarding other solution components as additives. Understanding the details of additive and environmental influences on the sol–gel properties allow us to exploit fundamental concepts of sol–gel engineering to direct solvent and surfactant choices to control particle size and demonstrate multiple widely employed lead halide perovskite films (CH3NH3PbI3, CH3NH3PbBr3, and CsPbBr3) with unprecedented surface roughness of less than 2 nm.
Article
Colloidal quantum dots (CQDs) are emerging as promising materials for constructing infrared sources in view of their tunable luminescence, high quantum efficiency and compatibility with solution processing. However, CQD films available today suffer from a compromise between luminescence efficiency and charge transport, and this leads to unacceptably high power consumption. Here, we overcome this issue by embedding CQDs in a high-mobility hybrid perovskite matrix. The new composite enhances radiative recombination in the dots by preventing transport-assisted trapping losses; yet does so without increasing the turn-on voltage. Through compositional engineering of the mixed halide matrix, we achieve a record electroluminescence power conversion efficiency of 4.9%. This surpasses the performance of previously reported CQD near-infrared devices two-fold, indicating great potential for this hybrid QD-in-perovskite approach.
Article
Organic-inorganic lead halide perovskites are distinct from most other semiconductors because they exhibit characteristics of both electronic and ionic motion. Accurate understanding of the optoelectronic impact of such properties is important to fully optimize devices and be aware of any limitations of perovskite solar cells and broader optoelectronic devices. Here we use a numerical drift-diffusion model to describe device operation of perovskite solar cells. To achieve hysteresis in the modeled current-voltage characteristics, we must include both ion migration and electronic charge traps, serving as recombination centers. Trapped electronic charges recombine with oppositely charged free electronic carriers, of which the density depends on the bias-dependent ion distribution in the perovskite. Our results therefore show that reduction of either the density of mobile ionic species or carrier trapping at the perovskite interface will remove the adverse hysteresis in perovskite solar cells. This gives a clear target for ongoing research effort and unifies previously conflicting experimental observations and theories.
Article
A delicate control of the stoichiometry, crystallographic phase, and grain structure of the photoactive material is typically required to fabricate high-performance photovoltaic (PV) devices. Organo-metal halide perovskite materials, however, exhibit a large degree of tolerance in synthesis, and can be fabricated into high efficiency devices by a variety of different vacuum and solution-based processes, with a wide range of precursor ratios. This suggests that the phase field for the desired material is wider than expected, or that high device efficiency may be achieved with a range of phases. Here, we investigate the structural and optical properties of the materials formed when a range of compositions of methylammonium iodide (MAI) and lead iodide (PbI2) were reacted at temperatures from 40 to 190 °C. The reactions were performed according to a commonly employed synthetic approach for high efficiency PV devices, and the data was analyzed to construct a pseudo-binary, temperature-dependent, phase-composition processing diagram. Escape of MAI vapor at the highest temperatures (150 - 190 °C) enabled a PbI2 phase to persist to very high MAI concentrations, and the processing diagram was not representative of phase equilibrium in this range. Data from reactions performed with a fixed vapor pressure of MAI allowed the high temperature portion of the diagram to be corrected and a near-equilibrium phase diagram to be proposed. The perovskite phase field is wider than previously thought under both processing conditions and extended by the existence of stacked perovskite sheet phases. Several aspects of the diagrams clarify why the organo-halide perovskite materials are compatible with solution processing.
Article
The band gap of formamidinium lead iodide (FAPbI3) perovskites allows broader absorption of the solar spectrum compared to conventional methylammonium lead iodide (MAPbI3). The optoelectronic properties of perovskite films are closely related to the film-quality, so depositing dense and uniform films is crucial for fabricating high-performance perovskite solar cells (PSCs). We report an approach for depositing high-quality FAPbI3 films, involving FAPbI3 crystallization by the direct intramolecular exchange of dimethylsulfoxide (DMSO) molecules intercalated in PbI2 with formamidinium iodide. This process produces FAPbI3 films with (111)-preferred crystallographic orientation, large-grained dense microstructures, and flat surfaces without residual PbI2. Using films prepared by this technique, FAPbI3-based PSCs with maximum power conversion efficiency of over 20% were fabricated. Copyright © 2015, American Association for the Advancement of Science.