You are on page 1of 6

ARTICLES

PUBLISHED ONLINE: 19 JUNE 2017 | DOI: 10.1038/NPHOTON.2017.94

High-performance direct conversion X-ray


detectors based on sintered hybrid lead
triiodide perovskite wafers
Shreetu Shrestha1,2, René Fischer3, Gebhard J. Matt1*, Patrick Feldner4, Thilo Michel5, Andres Osvet1,
Ievgen Levchuk1, Benoit Merle4, Saeedeh Golkar6, Haiwei Chen1, Sandro F. Tedde3, Oliver Schmidt3,
Rainer Hock7, Manfred Rührig3, Mathias Göken4, Wolfgang Heiss1, Gisela Anton5
and Christoph J. Brabec1,8

Lead halide perovskite semiconductors are in general known to have an inherently high X-ray absorption cross-section and
a significantly higher carrier mobility than any other low-temperature solution-processed semiconductor. So far, the
processing of several-hundred-micrometres-thick high-quality crystalline perovskite films over a large area has been
unresolved for efficient X-ray detection. In this Article, we present a mechanical sintering process to fabricate
polycrystalline methyl ammonium lead triiodide perovskite (MAPbI3) wafers with millimetre thickness and well-defined
crystallinity. Benchmarking of the MAPbI3 wafers against state-of-the-art CdTe detectors reveals competitive conversion
−1
efficiencies of 2,527 µC Gyair cm−2 under 70 kVp X-ray exposure. The high ambipolar mobility–lifetime product of
−4 2 −1
2 × 10 cm V is suggested to be responsible for this exceptionally high sensitivity. Our findings inform a new generation
of highly efficient and low-cost X-ray detectors based on perovskite wafers.

D
ue to their outstanding physical properties, hybrid organic– hundreds of micrometres for the X-ray energies commonly used
inorganic perovskites (HOIP), and most notably in medical applications. Although the widely reported solution-
CH3NH3(Pb,Sn)(I,Br)3 , have received extraordinary atten- process protocols for fabricating thin films of HOIPs are efficient
tion from the research community. Especially significant are the in the sub-micrometre regime, producing several-hundred-micro-
charge-transport properties of MAPbI3 , which has demonstrated metres-thick, high-quality films covering a large area is extremely
long minority-carrier lifetimes and diffusion lengths that are difficult. In this Article, we present a room-temperature sintering
comparable with single-crystalline covalent semiconductors1–5. process for MAPbI3 microcrystals to form thick (200 µm to 1 mm)
A new application for this material class is the sensitive detection rigid wafers of virtually any size.
of high-energy radiation such as X-ray6,7 and γ radiation8. The
search for an ideal X-ray-sensitive photoconductor (PC) is of MAPbI3 wafer sintering process
ongoing interest because most semiconductors with good charge- In contrast to covalent semiconductors such as Si and Cd(Zn)Te,
transport properties do not absorb high-energy radiation effectively. which require high-temperature crystallization processes, ionic crys-
The latter, however, is an intrinsic property of HOIPs due to the tals can be processed from solution at low temperatures. An easily
presence of heavy metal and halide ions. Current state-of-the-art overlooked property of ionic crystals is their plasticity, which
materials for direct X-ray detection9 include stabilized amorphous results in pressure-induced flow processes18. Although HOIPs
Se (a-Se)10, PbI2 (ref. 11), HgI2 (refs 12, 13), CdTe (ref. 14) and are not pure ionic crystals, this material class does exhibit
CdZnTe (ref. 15). a-Se detectors have been successfully commercia- pressure-induced agglomeration effects.
lized but are limited to mammography applications due to their low The MAPbI3 microcrystals presented here were synthesized by
absorption coefficient in the spectral regime higher than 50 keV. precipitation from a precursor solution at room temperature19 (see
PbI2 and HgI2 detectors face stability issues that hinder their wide- Methods). The precipitated microcrystals were irregular in shape,
spread application. Significant effort is being directed at Cd(Zn)Te with sizes ranging from 50 nm to 1 µm (Supplementary Fig. 1).
(ref. 16), but upscaling to larger wafers and charge carrier trapping By applying a pressure of 0.3 GPa to the microcrystals for 5 min
remain issues yet to be overcome. using a hydraulic press, a compact MAPbI3 layer (which will
Generally, for efficient X-ray absorption, the PC layer should be be referred to as a wafer) was formed. A series of wafers with
approximately three times larger than the attenuation length, which 1/2 inch diameter and thickness ranging from 200 µm to 1 mm
is the length at which 63% of photons are absorbed17. As discussed were prepared. It is remarkable that these free-standing MAPbI3
later in detail, HOIPs have attenuation lengths on the order of wafers have a mirror-like reflective surface with a root-mean-square

1
Friedrich-Alexander-University Erlangen-Nürnberg, Institute of Materials for Electronics and Energy Technology (i-MEET), Martensstrasse 7, 91058
Erlangen, Germany. 2 School of Advanced Optical Technologies (SAOT), Paul-Gordan-Strasse 6, 91052 Erlangen, Germany. 3 Siemens Healthineers,
Technology Center, Guenther-Scharowsky-Str. 1, 91058 Erlangen, 91058 Erlangen, Germany. 4 Materials Science and Engineering I, Friedrich-Alexander-
University Erlangen-Nürnberg, Martensstrasse 5, 91058 Erlangen, Germany. 5 Friedrich-Alexander-University Erlangen-Nürnberg, Centre for Astroparticle
Physics, Erwin-Rommel-Straße 1, 91058 Erlangen, Germany. 6 Institute of Particle Technology (LFG), Haberstrasse 9a, 91058 Erlangen, Germany.
7
Friedrich-Alexander-University Erlangen-Nürnberg, Lehrstuhl für Kristallografie und Strukturphysik, Staudtstrasse 3, 91058 Erlangen, Germany.
8
Bavarian Center for Applied Energy Research (ZAE Bayern), Am Weichselgarten 7, 91058 Erlangen, Germany. * e-mail: gebhard.matt@fau.de

436 NATURE PHOTONICS | VOL 11 | JULY 2017 | www.nature.com/naturephotonics

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2017.94 ARTICLES
a b

100 nm

200 μm

Figure 1 | Images of the sintered MAPbI3 wafer. a, Free-standing MAPbI3 wafer (1/2 inch × 1 mm). b, SEM cross-section of a sintered MAPbI3 wafer.
Inset: higher-magnification image. Circles indicate two interconnected microcrystals.

roughness of only ∼75 nm (Fig. 1a; for atomic force microscopy I4cm space group previously reported by Stoumpos et al. and
topography see Supplementary Fig. 3). Scanning electron microscopy Kawamura and co-authors21,29. The full-width at half-maximum
(SEM) of a wafer cross-section shows that the wafer is homogeneous (FWHM) of the diffraction peaks from the wafer and microcrystals
and dense (Fig. 1b). The density of the wafer is 3.76 g cm–3, which is are comparable. Also, the maximum of the photoluminescence
close to the density of single-crystalline MAPbI3 (4.15 g cm–3), as
calculated from lattice parameters20,21. a 1.2
It is evident from the SEM cross-section that the grain boundaries MAPbI3 single crystal
of the microcrystals remain well defined in the wafer. Furthermore, MAPbI3 wafer
1.0
grain boundary interconnections can be easily recognized (circles in
the inset of Fig. 1b).
Hardness (GPa)

0.8

Structural and mechanical investigations 0.6


Nanoindentation has emerged as the method for evaluating the
hardness and Young’s modulus of small material volumes and 0.4
thin films22,23. Measurements were performed with a Berkovich
tip, which was pressed 2.5 µm into the MAPbI3 wafer surface 0.2
using a constant strain rate of 0.05 s−1. Single-crystal MAPbI3
synthesized in house served as a reference for this measurement. 0.0
To provide a measurement with continuous stiffness, a sinusoidal 0.0 0.5 1.0 1.5 2.0 2.5
oscillation was superimposed to the load to evaluate the hardness Indentation depth (μm)
and Young’s modulus as a function of indentation depth24.
b 150
110
(counts s−1)

220

Hardness is defined as the ratio of the load on the tip to the MAPbI3 single crystal
Intensity

100
004

projected tip contact area.


202
211

The steep decrease in hardness at shallow indentation depths 150


shown in Fig. 2a is typical for such measurements and is attributed 0
to uncertainties in the geometry of the tip apex25. At higher indenta- 10 20 30 40 50 60
tion depths, the hardness of the MAPbI3 wafer and that of the single 2θ (°)
crystal levels off at ∼0.36 GPa and ∼0.47 GPa, respectively. The latter
(counts s−1)

value is in good agreement with previously reported values26. The


Intensity

40 MAPbI3 microcrystals
mechanical investigations were complemented by strain rate jump 20
tests, which are a measure of the dynamic mechanical behaviour of 0
a material, that is, the dependence of its strength on the imposed
10 20 30 40 50 60
deformation rate27. With this technique, a strain rate sensitivity m
2θ (°)
value of 0.065 was estimated (Supplementary Figs 4 and 5). This
value is in the upper range of m values found for metallic nanocrys-
(counts s−1)
Intensity

100 MAPbI3 wafer


tals28 and indicates that time-dependent deformation mechanisms
such as diffusion, dislocation activity and plastic flow due to grain 50
boundary sliding are dominant. This viscoplasticity is interpreted as 0
the underlying mechanism for the mechanical rigidity of the wafers. 10 20 30 40 50 60
To investigate the structural properties we performed X-ray dif- 2θ (°)
fraction (XRD) measurements. Figure 2b clearly shows that the dif-
fraction peaks of the MAPbI3 wafer, microcrystals and single crystal Figure 2 | Mechanical and structural properties of the MAPbI3 wafer.
are at the same positions. Major diffraction peaks at angles of 14.0, a, Nanointendation measurements on a MAPbI3 wafer and single crystal.
23.4, 24.4, 28.0 and 28.3° correspond to (110), (211), (202), (004) b, X-ray diffraction spectra of MAPbI3: single crystal (top), microcrystals
and (220) lattice planes, and are consistent with the tetragonal (middle) and sintered wafer (bottom).

NATURE PHOTONICS | VOL 11 | JULY 2017 | www.nature.com/naturephotonics 437

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
ARTICLES NATURE PHOTONICS DOI: 10.1038/NPHOTON.2017.94

spectrum at 770 nm (Supplementary Fig. 6) is in good agreement a 1


0.57 V μm−1
with the literature30–33.
To conclude, the observed time-dependent flow process triggers
a pressure-induced agglomeration of the microcrystals due to defor-
mations on their grain boundaries, and photoluminescence, XRD

Photocurrent (a.u.)
and SEM investigations confirm that the crystallinity of the micro- 0.39 V μm−1
crystals is retained. These results are consistent with in situ XRD
measurements performed on MAPbBr3 and MAPbI3 crystals,
which indicate that amorphization is observed at a much higher
pressure of 2 GPa, and it is reversible up to 34 GPa (ref. 34) and 0.20 V μm−1
50 GPa, respectively33.
0.1

Charge-transport investigations
In polycrystalline materials, the macroscopic charge-carrier
1 10
mobility and the free charge-carrier lifetime are dominated by the
Time (μs)
nature of the grain boundaries; that is, boundaries will scatter car-
riers being transported from grain to grain, surface defects may b 0.70
lead to deep trap states, and space-charge layers will act as potential
barriers35. To investigate the charge-carrier transport properties of
the polycrystalline MAPbI3 wafers, the time-of-flight (TOF) tech-

Hole mobility (cm2 V−1 s−1)


nique was used. TOF is an elegant method frequently used for 0.65
organic36 and inorganic semiconductors37. Generally, a laser pulse
generates a carrier reservoir Q in a thin layer near the illuminated
surface, which is subsequently extracted by the applied electric 0.60
field. The laser pulse intensity was kept low to prevent a distortion
of the electric field (Q < CV, where C is the sample capacitance and
V is the bias voltage). At too high laser intensities the excess charge-
carrier concentration distorts the electric field and the temporal 0.55
response features an overshoot38,39. Consequently, the laser intensity
was reduced until the temporal response of the transient current
remained unaltered. The device stack consisted of glass/indium 0.24 0.32 0.40 0.48 0.56
tin oxide (ITO)/poly(methyl methacrylate) (PMMA)/MAPbI3 Electric field (V μm−1)
wafer/Au. The 0.5-µm-thick PMMA layer was used to suppress
current injection from the ITO electrode. The sample was optically Figure 3 | Charge transport investigations. a, Time-of-flight transients of
excited from the glass side using a nanosecond laser emitting at holes. b, Electric field dependency of hole mobility.
532 nm. The transient photocurrent response for holes, shown in
Fig. 3a, represents a characteristic and well-resolved TOF transient, respectively. A bottom ITO/glass substrate and a silver top electrode
in which the transit time ttr of the fastest charge carriers is defined as finish the device stack.
the intersection of asymptotes from the plateau and tail in a double The sample was irradiated with X-ray radiation of 70 kVp from
logarithmic I(t) plot40,41. The TOF transients show a clear trend the silver electrode side for 2 s. The most intense emission was at
towards non-dispersive transport at higher electric fields. 38 keV (see Supplementary Information, pages 7 and 8 for the
The mobility is calculated from ttr using μ = d/(Ettr), where d set-up with the X-ray source and the simulated X-ray photon-
is the sample thickness and E is the applied electric field. For density spectrum). A simulation using the XCOM database esti-
holes, the mobility is in the range of 0.53–0.70 cm2 V−1 s−1, with a mates an attenuation depth in the MAPbI3 wafer at 38 keV of
negative field dependency (that is, due to increased scattering at ∼125 µm (Supplementary Fig. 11). Consequently, the fraction of
higher E fields, Fig. 3b). Under the same experimental conditions incident X-ray photons that are attenuated approaches unity17
but with reversed bias polarity, a slightly lower mobility of (Supplementary equation (3)).
0.50–0.56 cm2 V−1 s−1 and a more dispersive transport for electrons The photocurrent response under constant reverse bias (the
(Supplementary Figs 7 and 8) is measured. electrodes are selectively e/h extracting) and increasing dose rates
This observed ambipolar transport is remarkable, as very low or is presented in Fig. 4b. The extracted charge shows a nearly linear
even missing charge transport of either electrons or holes is com- X-ray dose dependency up to 5 mGyair (Fig. 4d). The slope of a
monly observed in amorphous and polycrystalline systems11,42. linear fit (red line in Fig. 4d) is defined as the sensitivity of the
The field dependency of the charge transport indicates potential detector and we obtained a conservatively estimated sensitivity of
−1
barriers (that is, at the grain boundaries), while quenching of free ∼2,527 µC Gyair cm−2 at an electric field of 0.2 V µm–1. Figure 4c
charge-carrier lifetime and of the macroscopic mobility via deep shows the dependence of the extracted charge from the MAPbI3
traps is unlikely to be present. These findings are consistent with sample on the electric field at a constant dose rate of 6.72 mGyair s−1.
recent theoretical investigations that show that MAPbI3 is a A fit with the Hecht equation gives a tentative μτ product of
defect-tolerant semiconductor and that vacancy-type defects (that 2 × 10–4 cm2 V−1 (red line in Fig. 4c).
is, iodine vacancies) are resonant in the bands43. To compare and validate the X-ray response of the MAPbI3 wafer-
based detector, we measured the charge released from a 1-mm-thick
Direct X-ray to current conversion single-crystalline CdTe sensor layer on the photon-counting pixel
The device stack of the X-ray detector is shown in Fig. 4a. The detector ‘Timepix’ (see ref. 44 and the Supplementary Information,
electric contacts of the 1-mm-thick MAPbI3 wafer are formed by page 12, for a technical description). For comparison, the CdTe
poly(3,4-ethylenedioxythiophene) polystyrene sulfonate (PEDOT: sensor layer has the same thickness as the MAPbI3 wafer and was
PSS) and phenyl-C61-butyric acid methyl ester (PCBM) and ZnO biased with the same electric field. As is evident from Fig. 4d, the
as hole-selective and electron-selective contact and buffer layers, amount of extracted charge Q from the MAPbI3 sample is comparable

438 NATURE PHOTONICS | VOL 11 | JULY 2017 | www.nature.com/naturephotonics

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2017.94 ARTICLES
a b 25

Ag ZnO

M 20
PCB

Photocurrent (μA cm−2)


PbI 3
MA 15
:PSS
OT
PED
ITO
10
ss
Gla

5
0.048 5.86 mGyair s−1

20 40 60 80 100 120 140 160


Time (s)

c 25 d
50 CdTe 'Timepix' detector

MAPbI3 sample
20 40
Charge (μC cm−2)

Charge (μC cm−2)


30
15

20

10
10

μτ E d
Q = Q0 1 − exp −
5 d μτ E 0

0.00 0.05 0.10 0.15 0.20 0 2 4 6 8 10 12


Electric field (V μm−1) Dose (mGyair)

Figure 4 | Direct X-ray to current conversion. a, Device stack of the MAPbI3-wafer-based X-ray detector. The MAPbI3 wafer is 1 mm thick. b, Time-resolved
photocurrent at E = 0.2 V µm–1 with different dose rates. c, Extracted charge versus electric field. The dose rate is kept constant at 6.72 mGyair s–1.
d, Extracted charge at constant E = 0.2 V µm–1 for the MAPbI3 wafer-based sample and the CdTe ‘Timepix’ reference detector. All exposures are
2-s-long pulses from an X-ray source operated at 70 kV.

to the commercial CdTe ‘Timepix’ reference detector. For doses Carrier-selective electrodes featuring improved e/h selectivity
<5 mGyair , Q is almost identical to that of the CdTe reference (∼5% are currently an area of intense research. Improvements in this
higher for the CdTe sample). For doses >5 mGyair , Q deviates from direction are expected to further advance the performance of
the CdTe reference (due to enhanced recombination). perovskite-based X-ray converters.
Another essential parameter is the ionization energy W± , which A comparison of previously reported sensitivities, as shown in
is the amount of radiation energy absorbed to create a single free the Supplementary Information (Supplementary Fig. 14), reveals
electron and/or hole. Stated differently, it is the total adsorbed that the sensitivity value, the μτ product9 and the ionization
energy divided by the total number of extracted electrons. energy W± are on par with the best available technologies, such as
For the MAPbI3 wafer, an ionization energy of ∼5 eV at 0.2 V HgI2 and Cd(Zn)Te (refs 9, 13). In addition to its toxicity, the elec-
µm–1 electric field is calculated (see Supplementary equation (7)). tronic parameters of HgI2 are strongly dependent on the deposition
This value is in close agreement with the Que-Rowlands45 and process and can vary from sample to sample49. Conversely, crystal-
Klein rules46, which predict the lowest limit of W± according to line Cd(Zn)Te wafers offer close to ideal properties but require
W± = 2.2Eg + Ephonon ≤ 3.9 eV and W± = 3Eg = 4.65 eV, respectively, energy-demanding production processes such as zone melting, the
where Eg = 1.55 eV is the bandgap and Ephonon ≤ 0.5 eV is the Bridgman method, epitaxial growth, or the travelling heater
phonon energy. We conclude that the low W± is due to an excellent method (THM)50. Furthermore, wafer inhomogeneities due to Cd,
collection efficiency (CE) and a low geminate recombination rate. Te inclusions51 and upscaling to larger areas remain issues with
This finding is in agreement with recent investigations on these technologies16. Many of theses shortcomings are resolved by
MAPbI3 single crystals claiming a CE of 75% using 20–35 keV the presented sintering method for MAPbI3 microcrystals, which
soft X-ray irradiation with an electrode distance of 2.5 mm (ref. 47). allows the formation of rigid wafers of virtually any size.
Despite these excellent electronic properties, one important In conclusion, in this Article, we present a sintering process to
drawback of these devices is the high and rather unstable dark- manufacture rigid, several-hundred-micrometres-thick MAPbI3
current density of 6 µA cm–2 at 0.2 V µm–1 reverse bias (see baseline wafers. The wafer conserves the structural and optical properties of
drift in Fig. 4b). In MAPbI3 , the dark conductivity is dominated by the microcrystalline starting material. Ambipolar charge transport is
ionic conduction (free iodine migration48) and the current instabil- demonstrated with a mobility of 0.45–0.7 cm2 V−1 s−1. Under X-ray
ity at reverse bias is caused by the non-ideal e/h selective electrodes. exposure, a μτ product of ∼2 × 10–4 cm2 V−1 is measured with an

NATURE PHOTONICS | VOL 11 | JULY 2017 | www.nature.com/naturephotonics 439

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
ARTICLES NATURE PHOTONICS DOI: 10.1038/NPHOTON.2017.94

−1
excellent conversion sensitivity of 2,527 µC Gyair cm−2 at 0.2 V µm–1 30. Kim, Y. H. et al. Multicolored organic/inorganic hybrid perovskite light-emitting
electric field. diodes. Adv. Mater. 27, 1248–1254 (2015).
31. Yamada, Y., Nakamura, T., Endo, M., Wakamiya, A. & Kanemitsu, Y. Near-
band-edge optical responses of solution-processed organic–inorganic hybrid
Methods perovskite CH3NH3PbI3 on mesoporous TiO2 electrodes. Appl. Phys. Express 7,
Methods and any associated references are available in the online 032302 (2014).
version of the paper. 32. Fedeli, P. et al. Influence of the synthetic procedures on the structural and optical
properties of mixed-halide (Br, I) perovskite films. J. Phys. Chem. C 119,
Received 20 October 2016; accepted 12 May 2017; 21304–21313 (2015).
published online 19 June 2017 33. Jaffe, A. et al. High-pressure single-crystal structures of 3D lead-halide hybrid
perovskites and pressure effects on their electronic and optical properties. ACS
Cent. Sci. 2, 201–209 (2016).
References 34. Wang, Y. et al. Pressure-induced phase transformation, reversible amorphization
1. Saparov, B. & Mitzi, D. B. Organic–inorganic perovskites: structural versatility
and anomalous visible light response in organolead bromide perovskite. J. Am.
for functional materials design. Chem. Rev. 116, 4558–4596 (2016).
Chem. Soc. 137, 11144–11149 (2015).
2. Stranks, S. D. & Snaith, H. J. Metal-halide perovskites for photovoltaic and
35. Bube, R. H. Electronic transport in polycrystalline films. Ann. Rev. Mater. Sci. 5,
light-emitting devices. Nat. Nanotech. 10, 391–402 (2015).
201–224 (1975).
3. Stranks, S. D. et al. Electron–hole diffusion lengths exceeding 1 micrometer in an
36. Bässler, H. Charge transport in disordered organic photoconductors: a Monte
organometal trihalide perovskite absorber. Science 342, 341–344 (2013).
Carlo simulation study. Phys. Status Solidi B 175, 15–56 (1993).
4. Zakutayev, A. et al. Defect tolerant semiconductors for solar energy conversion.
37. Suzuki, K., Shorohov, M., Sawada, T. & Seto, S. Time-of-flight measurements on
J. Phys. Chem. Lett. 5, 1117–1125 (2014).
TiBr detectors. IEEE Trans. Nucl. Sci. 62, 433–436 (2015).
5. Dong, Q. et al. Electron–hole diffusion lengths >175 μm in solution-grown
38. Papadakis, A. C. Theory of transient space-charge perturbed currents in
CH3NH3PbI3 single crystals. Science 347, 967–970 (2015).
insulators. J. Phys. Chem. Solids 28, 641–647 (1967).
6. Yakunin, S. et al. Detection of X-ray photons by solution-processed lead halide
39. Suzuki, K., Ichinohe, Y., Sawada, T., Kazuaki Imai, K. & Seto, S. Time-of-flight
perovskites. Nat. Photon. 9, 444–449 (2015).
measurements on Schottky CdTe nuclear detectors. Phys. Status Solidi C 11,
7. Wei, H. et al. Sensitive X-ray detectors made of methylammonium lead
1337–1340 (2014).
tribromide perovskite single crystals. Nat. Photon. 10, 333–339 (2016).
40. Scharfe, M. E. Transient photoconductivity in vitreous As2Se3. Phys. Rev. B 2,
8. Yakunin, S. et al. Detection of gamma photons using solution-grown single
5025–5034 (1970).
crystals of hybrid lead halide perovskites. Nat. Photon. 10, 585–589 (2016).
41. Hirao, A., Tsukamoto, T. & Nishizawa, H. Analysis of nondispersive time-of-
9. Kasap, S. O., Kabir, M. Z. & Rowlands, J. A. Recent advances in X-ray
flight transients. Phys. Rev. B 59, 12991–12995 (1999).
photoconductors for direct conversion X-ray image detectors. Curr. Appl. Phys.
42. Pfister, G. & Scher, H. Time-dependent electrical transport in amorphous solids:
6, 288–292 (2006).
As2Se3. Phys. Rev. B 15, 2062–2083 (1977).
10. Kasap, S. O. X-ray sensitivity of photoconductors: application to stabilized a-Se.
43. Brandt, R. E., Stevanović, V., Ginley, D. S. & Buonassisi, T. Identifying defect-
J. Phys. D 33, 2853–2865 (2000).
tolerant semiconductors with high minority-carrier lifetimes: beyond hybrid
11. Street, R. A. et al. Electronic transport in polycrystalline PbI2 films. J. Appl. Phys.
lead halide perovskites. MRS Commun. 5, 265–275 (2015).
86, 2660–2667 (1999).
44. Filipenko, M., Gleixner, T., Anton, G., Durst, J. & Michel, T. Characterization of
12. Zentai, G., Schieber, M., Partain, L., Pavlyuchkova, R. & Proano, C. Large area
the energy resolution and the tracking capabilities of a hybrid pixel detector with
mercuric iodide and lead iodide X-ray detectors for medical and non-destructive
CdTe-sensor layer for a possible use in a neutrinoless double beta decay
industrial imaging. J. Cryst. Growth 275, e1327–e1331 (2005).
experiment. Eur. Phys. J. C 73, 2374 (2013).
13. Schieber, M. et al. Thick films of X-ray polycrystalline mercuric iodide detectors.
45. Que, W. & Rowlands, J. A. X-ray photogeneration in amorphous selenium:
J. Cryst. Growth 225, 118–123 (2001).
geminate versus columnar recombination. Phy. Rev. B 51, 10500–10507 (1995).
14. Richter, M. & Siffert, P. High resolution gamma ray spectroscopy with CdTe
46. Klein, C. A. Bandgap dependence and related features of radiation ionization
detector systems. Nucl. Instrum. Methods Phys. Res. Sect. A 322, 529–537 (1992).
energies in semiconductors. J. Appl. Phys. 39, 2029–2038 (1968).
15. Szeles, C. CdZnTe and CdTe materials for X-ray and gamma ray radiation
47. Náfrádi, B., Náfrádi, G., Forró, L. & Horváth, E. Methylammonium lead iodide
detector applications. Phys. Status Solidi B 241, 783–790 (2004).
for efficient X-ray energy conversion. J. Phys. Chem. C 119, 25204–25208 (2015).
16. Seller, P. et al. Pixellated Cd(Zn)Te high-energy X-ray instrument. J. Instrum. 6,
48. Besleaga, C. et al. Iodine migration and degradation of perovskite solar cells
C12009 (2011).
enhanced by metallic electrodes. J. Phys. Chem. Lett. 7, 5168–5175 (2016).
17. Kasap, S. et al. Amorphous and polycrystalline photoconductors for direct
49. Schieber, M. et al. Theoretical and experimental sensitivity to X-rays of single
conversion flat panel X-ray image sensors. Sensors 11, 5112–5157 (2011).
and polycrystalline HgI2 compared with different single-crystal detectors. Nucl.
18. Skrotzki, W., Frommeyer, O. & Haasen, P. Plasticity of polycrystalline ionic
Instrum. Methods Phys. Res. Sect. A 458, 41–46 (2001).
solids. Phys. Status Solidi A 66, 219–228 (1981).
50. Beer, A. C., Willardson, R. K. & Weber, E. (eds) Semiconductors for Room
19. Zhu, F. et al. Shape evolution and single particle luminescence of organometal
Temperature Nuclear Detector Applications Vol. 43 (Academic, 1995).
halide perovskite nanocrystals. ACS Nano 9, 2948–2959 (2015).
51. Belas, E. et al. Reduction of inclusions in (CdZn)Te and CdTe:In single crystals
20. Poglitsch, A. & Weber, D. Dynamic disorder in
by post-growth annealing. J. Electron. Mater. 37, 1212–1218 (2008).
methylammoniumtrihalogenoplumbates (II) observed by millimeter-wave
spectroscopy. J. Chem. Phys. 87, 6373–6378 (1987).
21. Stoumpos, C. C., Malliakas, C. D. & Kanatzidis, M. G. Semiconducting tin and Acknowledgements
lead iodide perovskites with organic cations: phase transitions, high The Cluster of Excellence Engineering of Advanced Materials (EAM) at the FAU
mobilities, and near-infrared photoluminescent properties. Inorg. Chem. 52, University Erlangen and the Gradko 1896 ‘in situ Microscopy’ (DFG) is acknowledged for
9019–9038 (2013). support. The authors thank C.O. Quiroz and M. Salvador for reading the manuscript.
22. Lucca, D. A., Herrmann, K. & Klopfstein, M. J. Nanoindentation: measuring
methods and applications. CIRP Ann. Manuf. Technol. 59, 803–819 (2010). Author contributions
23. Oliver, W. C. & Pharr, G. M. An improved technique for determining hardness S.S., R.F., G.J.M., T.M., O.S., M.R. and S.F.T. performed the electrical characterization under
and elastic modulus using load and displacement sensing indentation X-ray exposure at the Siemens Healthineers Technology Center. I.L. and S.S. synthesized
experiments. J. Mater. Res. 7, 1564–1583 (2011). the MAPbI3 single crystals and microcrystals. H.C. helped with processing of the X-ray
24. Hay, J., Agee, P. & Herbert, E. Continuous stiffness measurment during detectors. S.S., A.O. and G.M. performed TOF experiments. B.M. and P.F. carried out the
instrumented indentation testing. Exp. Techniques 34, 86–94 (2010). mechanical investigations. R.H. interpreted the XRD data. S.S. and S.G. performed the SEM
25. Nix, W. D. & Gao, H. Indentation size effects in crystalline materials: a law for investigation. G.J.M. wrote the manuscript. M.G., W.H., G.A. and C.J.B. initiated and
strain gradient plasticity. J. Mech. Phys. Solids 46, 411–425 (1998). supervised the work. All authors reviewed the manuscript.
26. Sun, S., Fang, Y., Kieslich, G., White, T. J. & Cheetham, A. K. Mechanical
properties of organic–inorganic halide perovskites, CH3NH3PbX3 (X = I, Br and
Cl), by nanoindentation. J. Mater. Chem. A 3, 18450–18455 (2015). Additional information
27. Maier, V. et al. Nanoindentation strain-rate jump tests for determining the local Supplementary information is available in the online version of the paper. Reprints
strain-rate sensitivity in nanocrystalline Ni and ultrafine-grained Al. J. Mater. and permissions information is available online at www.nature.com/reprints.
Rec. 26, 1421–1430 (2011). Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
28. Wei, Q., Cheng, S., Ramesh, K. T. & Ma, E. Effect of nanocrystalline and ultrafine published maps and institutional affiliations. Correspondence and requests for materials
grain sizes on the strain rate sensitivity and activation volume: fcc versus bcc should be addressed to G.J.M.
metals. Mater. Sci. Eng. A 381, 71–79 (2004).
29. Kawamura, Y., Mashiyama, H. & Hasebe, K. Structural study on cubic– Competing financial interests
tetragonal transition of CH3NH3PbI3. J. Phys. Soc. Jpn 71, 1694–1697 (2002). The authors declare no competing financial interests.

440 NATURE PHOTONICS | VOL 11 | JULY 2017 | www.nature.com/naturephotonics

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NATURE PHOTONICS DOI: 10.1038/NPHOTON.2017.94 ARTICLES
Methods emission was recorded with a Si charge-coupled device camera (Syncerity, Horiba
MAPbI3 single-crystal and microcrystal synthesis. The MAPbI3 microcrystals were Jobin-Yvon) attached to an iHR320 monochromator (Horiba Jobin-Yvon) with a
synthesized by a precipitation reaction as reported in ref. 19. A 1 M precursor resolution of 5 nm. The spectra were corrected for the spectral sensitivity of the
solution was prepared by dissolving an equimolar ratio of methylammonium system, determined with the help of a calibrated halogen lamp (Stellarnet).
iodide CH3NH3I (synthesized in house) and lead(II) iodide (Sigma Aldrich) in
γ-butyrolactone (GBL) at 60 °C and stirred for 3 h. Chloroform was Nanoindentation. Nanoindentation measurements were performed using a
added to the filtered precursor solution, which immediately precipitated Keysight G200 Nanoindenter. A Berkovich tip was pressed 2.5 µm into the sample
CH3NH3PbI3 microcrystals. The microcrystals were washed twice with chloroform surface, using a constant strain rate of 0.05 s–1. Nanoindentation was carried out in
in a centrifuge, and then dried in a vacuum oven at 40 °C overnight. The synthesis of continuous stiffness mode (CSM), where a small oscillation is superposed on the
the single crystals followed procedures in refs 52 and 53. In brief, the precursor was indentation load. A 2 nm sinusoidal oscillation was superposed to the loading signal.
heated at 100 °C for 3 h, which produced small MAPbI3 crystals. One small The hardness values were average over 25 measurements taken at least 100 µm apart.
crystal was used as a seed in 1 M precursor solution and allowed to grow overnight Strain-rate jumps tests were implemented during the nanoindentation loading
at 100 °C. segment. The corresponding strain rates ranged between 0.025 s–1 and 0.005 s–1.
Strain-rate sensitivity was determined as the slope of a linear fit in a double
Sintering process. A hydraulic press (Specac) with a 1/2-inch die was used for logarithmic plot of hardness versus strain rate.
sintering. A polished stainless-steel cylinder was placed in the bore of the cylinder
body followed by 250–500 mg of CH3NH3PbI3 powder. Next, a second polished TOF. PMMA (Sigma Aldrich) in chlorobenzene (100 mg ml–1) was spin-coated on
cylinder and a plunger were inserted into the cylinder body. A vacuum membrane cleaned ITO substrates to form a 500-nm-thick layer. MAPbI3 wafers were pressed
pump was connected to the die. After about 2 min of turning on the vacuum pump, into the PMMA/ITO substrates at ∼50 MPa for 2 min using the hydraulic press. The
a pressure of 0.3 GPa was applied to the plunger for about 5 min to form the Au top contacts (70 nm thick and 0.2 cm2 in area) were thermally evaporated to
MAPbI3 wafer. provide the charge-collecting electrode. As excitation, a nanosecond pulse laser
(532 nm, CryLaS) was used. The bias voltage was applied to the sample using a
Fabrication of the MAPbI3 wafer based photodetector. Glass substrates with function generator (Agilent 33500B) and a voltage amplifier (Falco Systems
sputtered ∼100-nm-thick ITO were cleaned in an ultrasonic bath with acetone and WMA 300). The transient current was measured as the voltage drop over a load
isopropanol for 10 min and then placed in an oxygen plasma oven for 2 min. resistor and recorded with a digital oscilloscope (Tektronix DPO 3034).
PEDOT:PSS (VP Al 4083 from H.C. Starck) was spin-coated onto the cleaned
substrate to form a 100-nm-thick thin film. MAPbI3 wafers were pressed onto the Photocurrent under X-ray exposure. A 70 kV X-ray source (Siemens MEGALIX
PEDOT:PSS/ITO substrates at about 15 MPa for 2 min using a hydraulic press. Cat Plus 125/40/90, 124 GW) with a tungsten anode was used. The X-ray spectrum
PCBM (20 mg ml–1) then ZnO were subsequently spin-coated onto the wafers. The was filtered with a 2.5-mm-thick Al plate. The dose rate was changed by changing
Ag top contacts (6.6 mm2 in area and 70–100 nm thick) were deposited by the X-ray tube current and was calibrated with a PTW Diados T11003-001896
thermal evaporation. dosimeter. A Keithley 2400 SMU measuring at 10 Hz was used to record the
photocurrent. See Supplementary Information for additional details.
SEM. SEM measurements were performed on MAPbI3 wafers and the
microcrystalline particles supported on standard adhesive carbon pads using an Data availability. The data that support the plots within this paper and other
ULTRA 55 Carl Zeiss AG SEM under 10 kV acceleration voltage. For the image of findings of this study are available from the corresponding author upon
the cross-section, the wafers were ‘broken’ with a hammer. reasonable request.

Atomic force microscopy measurements. An NT-MDT nano-educator AFM in References


contact mode was used. 52. Liu, Y. et al. Two-inch-sized perovskite CH3NH3PbX3 (X = Cl, Br, I) crystals:
growth and characterization. Adv. Mater. 27, 5176–5183 (2015).
Photoluminescence. Photoluminescence spectra were measured using a 450 nm 53. Zhou, J., Chu, Y. & Huang, J. Photodetectors based on two-dimensional
laser diode as the excitation source in a back-scattering configuration at room layer-structured hybrid lead iodide perovskite semiconductors. ACS Appl.
temperature. The laser beam was focused to a spot with a diameter of ∼10 µm. The Mater. Interfaces 8, 25660–25666 (2016).

NATURE PHOTONICS | www.nature.com/naturephotonics

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.

You might also like