You are on page 1of 11

Articles

https://doi.org/10.1038/s41565-020-0765-7

2D metal–organic framework for stable perovskite


solar cells with minimized lead leakage
Shengfan Wu1, Zhen Li1, Mu-Qing Li   1,2, Yingxue Diao   1,3, Francis Lin   1, Tiantian Liu1, Jie Zhang1,
Peter Tieu   4, Wenpei Gao5, Feng Qi1, Xiaoqing Pan   5,6,7, Zhengtao Xu   1 ✉, Zonglong Zhu   1 ✉ and
Alex K.-Y. Jen   1,3,8 ✉

Despite the notable progress in perovskite solar cells, maintaining long-term operational stability and minimizing potentially
leaked lead (Pb2+) ions are two challenges that are yet to be resolved. Here we address these issues using a thiol-functionalized
2D conjugated metal–organic framework as an electron-extraction layer at the perovskite/cathode interface. The resultant
devices exhibit high power conversion efficiency (22.02%) along with a substantially improved long-term operational stability.
The perovskite solar cell modified with a metal–organic framework could retain more than 90% of its initial efficiency under
accelerated testing conditions, that is continuous light irradiation at maximum power point tracking for 1,000 h at 85 °C. More
importantly, the functionalized metal–organic framework could capture most of the Pb2+ leaked from the degraded perovskite
solar cells by forming water-insoluble solids. Therefore, this method that simultaneously tackles the operational stability and
lead contamination issues in perovskite solar cells could greatly improve the feasibility of large-scale deployment of perovskite
photovoltaic technology.

T
he certified power conversion efficiency (PCE) of organic– well-defined hybrid materials. So far, there are only very few cases
inorganic lead halide perovskite solar cells (PVSCs) has that report the use of MOFs in PVSCs, indicating the very embry-
reached a new high at 25.2%, which can rival those of onic stage of this area. Among the limited studies, MOFs have been
market-dominating inorganic photovoltaic technologies, includ- applied as scaffolds to template the growth of perovskites or as addi-
ing polycrystalline silicon (p-Si), thin film copper indium gallium tives/surface modifiers to passivate the defects of perovskites for
selenide and cadmium telluride1–5. The demonstrated high PCE enhancing the device performance and stability18,21. However, it is
indicates that it is no longer a limiting factor for commercializing challenging to use a MOF as an efficient charge-transporting layer
PVSCs. As such, research efforts in this field have been gradually (CTL) because most of them are quite insulating with low carrier
shifted to focus on the reliability and stability issues of PVSCs under mobilities19. Additionally, the MOFs utilized in PVSCs are mostly
ambient conditions, light soaking and thermal stresses6–9. By apply- unfunctionalized prototypes that lack the diverse chemical func-
ing composition and crystal engineering, encapsulation, electrode tionalities needed for integrating with PVSCs.
selections, defect passivation and interfacial modifications, the To address these problems, we have developed a simple and
device lifetime can be prolonged by excluding external environ- effective strategy by applying a 2D conjugated MOF that pos-
mental tensions10–13. Nevertheless, long-term stability of PVSCs sesses n-type electrical behaviour and suitable energy levels as an
under operational conditions has yet to be realized, which remains electron-extraction layer (EEL) at the perovskite/cathode interface.
a critical hurdle to overcome before their large-scale implementa- The use of Zr(iv) ions to engage selectively the hard carboxyl groups
tion. Moreover, compared to the thermally and chemically stable allows dense free-standing thiol arrays to be built around the Zr(iv)–
cadmium telluride having low solubility in water (with a solubility oxo cluster to offer steric shielding and stabilization22,23. The dense
product constant (Ksp) below 10−30) and presenting a lower toxic- array of thiol groups on the MOF is particularly important because of
ity threat to the environment, lead (Pb)-containing compounds in their capabilities in rigidifying the networks with crosslinked disul-
perovskites (or degraded perovskites) have a much higher Ksp of fide linkages and in trapping heavy metal ions (such as lead, mercury
10−8 in rainwater11,14–16. Therefore, to improve PVSC operational and so on). In addition, the thiol-functionalized MOF can also form
stability effectively and eliminate/minimize Pb2+ leakage from the intimate contact with the silver electrode to extract electrons effi-
degraded perovskite absorbers, an integrated solution with proper ciently from PVSCs by reducing the contact resistance and trapping
materials and device configurations is required to enable PVSCs for mobile Pb2+ ions at the perovskite/electrode interface to mitigate the
practical applications. potential impact of the device on environmental sustainability.
Metal–organic frameworks (MOFs) have been recently explored
in the perovskite photovoltaic community due to their potential Characterization of 2D MOF and perovskite/MOF films
for enhancing photovoltaic performance17–21. This stems from the As shown in Fig. 1a, a new linker molecule H3L3 was first synthesized
relative ease in tuning their functionalities as a versatile class of (Supplementary Figs. 1–4). The peak numbers and assignments

Department of Chemistry, City University of Hong Kong, Kowloon, Hong Kong. 2Frontier Institute of Science and Technology (FIST), Xi’an Key Laboratory
1

of Sustainable Energy and Materials Chemistry, Xi’an Jiaotong University, Shanxi, China. 3Department of Materials Science and Engineering, City University
of Hong Kong, Kowloon, Hong Kong. 4Department of Chemistry, University of California-Irvine, Irvine, CA, USA. 5Department of Materials Science and
Engineering, University of California-Irvine, Irvine, CA, USA. 6Department of Physics and Astronomy, University of California-Irvine, Irvine, CA, USA. 7Irvine
Materials Research Institute, University of California-Irvine, Irvine, CA, USA. 8Department of Materials Science and Engineering, University of Washington,
Seattle, WA, USA. ✉e-mail: zhengtao@cityu.edu.hk; zonglzhu@cityu.edu.hk; alexjen@cityu.edu.hk

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

a b c d
O OH

HS SH

HS SH

HO O

O SH SH OH 5 nm 5 nm

H3L3 ZrL3

e f
500
0.6
0.4 r = 0.999994 r = 0.999991
N2 uptake (cm3 g–1 STP)

400 0.4 Boiling water treated ZrL3


0.2
300 0.2
Activited ZrL3
0 0
0 0.02 0.04 0.06 0 0.02 0.04 0.06
200
As-made ZrL3
100
528 m2 g–1 369 m2 g–1 Calculated from the structural model of ZrL3
0
0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0 10 20 30 40
Relative pressure (P/P0) Relative pressure (P/P0) 2θ (°)

g h i
6
Pero Pero Pero/ZrL3
ZrL3-as made 100 Pero/PC61BM Pero/PC61BM/ZrL3
Pero/ZrL3
5 Band gap = 2.53 eV 12 Pero/PC61BM

Normalized intensity (a.u.)


PL intensity (×105 counts)

Pero/PC61BM/ZrL3
4 9 10
0

intensity (a.u.)
10–1
α (S)

Normalized
3 10
–1

6
2 –2
10
3
1
10–2 0 25 50 75 100 125 150
Time (ns)
0 0
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 650 700 750 800 850 0 150 300 450 600 750
Energy (eV) Wavelength (nm) Time (ns)

Fig. 1 | Characterization of ZrL3 and the perovskite films with different EELs. a, The thiol-decked carboxyl linker H3L3. b, Structural model of corresponding
2D MOF ZrL3. c, Bright field image. d, Model of ZrL3. e, N2 sorption isotherm at 77 K and Brunauer–Emmett–Teller plot (insets) for the activated ZrL3 before
(left) and after (right) boiling-water-treatment. f, PXRD patterns of ZrL3. g, UV-vis diffuse reflectance spectrum of the as-made ZrL3. h,i, PL (h) and TRPL
(i) spectra of the perovskite films with different EELs.

in the NMR spectra confirmed the molecular structure of the should be noted that their ratios here cannot be pinpointed by the
H3L3 linker and indicated the high purity of the H3L3 sample used current data as their small weight fractions do not impact signifi-
for assembling the Zr(iv)-based MOF product, ZrL3. The corre- cantly the elemental and TGA results.
sponding polycrystalline framework solid, ZrL3 (Fig. 1b), was then The morphology, micro and atomic structures of the 2D MOF
solvothermally prepared, using ZrOCl2∙8H2O as the metal source, were also characterized by scanning electron microscopy (SEM) and
formic acid as the modulator and ethylene dithiol for stabilizing the transmission electron microscopy (TEM). The SEM image of ZrL3
thiol groups on the organic linker. The peak S–H stretch at around features the morphology of thin platelets with dimensions between
2,555 cm−1 can be clearly observed from the FT-IR spectra of both 100 and 500 nm (Supplementary Fig. 10). The low magnification
H3L3 and ZrL3 (Supplementary Figs. 5 and 6), indicating the suc- image, fast Fourier transform and electron diffraction show the dis-
cessfully installed thiol groups in the ZrL3 networks. The powder tinct hexagonal arrays of the 2D layers in ZrL3, as demonstrated
X-ray diffraction (PXRD) pattern of as-made ZrL3 (Supplementary in Supplementary Fig. 11. Moreover, atomic resolution high angle
Fig. 7) can be indexed onto hexagonal unit cells (parameters: annular dark field images from aberration corrected scanning trans-
a = 19.98 Å, c = 7.24 Å (see Supplementary Note 1). The short c axis mission electron microscopy indicate the direct (eclipsed) stacking
suggests a single-layer structure based on Zr6O8 clusters as it com- of the MOF layers (Supplementary Fig. 12). Individual Zr atoms
pares well with the corresponding interlayer distance in a known are seen within the Zr clusters at high magnification. The surface
structure. Elemental analysis and thermogravimetric analysis clusters of a thin region were found to be more diffuse compared
(TGA) results fit the formula Zr6O4(OH)4(L3)2(OH)6(H2O)27(DMF) to the interior. The organic linkers between Zr clusters were seen
(DMF = N,N-dimethylformamide) for ZrL3 (Supplementary Fig. 8 in the bright field image and the surface of ZrL3 was also shown
and Note 2). The linker/Zr6 cluster ratio thus determined suggests to be terminated by the organic linkers as in the model (Fig. 1c,d).
full occupancy for the linker sites in ZrL3. The capping sites can TEM studies also suggest superior stability of the 2D networks as
be occupied by the formate (HCOO−) and HO−/H2O species, but it ZrL3 was stable under prolonged parallel electron beam radiation

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
(~30 e− Å−2 s−1) while imaging. The stability of the ZrL3 lattice is PC61BM and Ag, a much higher FF of 76.42% along with a higher
further verified by gas sorption studies and PXRD. The N2 sorption PCE of 20.34% was achieved (Supplementary Figs. 21 and 22). These
isotherm (77 K) features dominant type-I characteristics of micro- results suggest the enhanced electron extraction and hole-blocking
porous solids (Fig. 1e), with additional uptake at high P/P0 regions ability originated from ZrL3, which is consistent with the TRPL
that can be attributed to capillary condensation in mesopores (for results. To extract electrons even more efficiently, a bis-C60 surfac-
example, arising from the inter-particle/layer spaces). tant (previously demonstrated to be able to tune the cathode work
The Brunauer–Emmett–Teller surface area is calculated to be function for improving electron extraction in both organic photo-
528 m2 g−1 for ZrL3, which agrees with the simulated accessible voltaic devices and PVSCs (ref. 27–30)) was further incorporated with
solvent surface of 594 m2 g−1 (Supplementary Fig. 13). To probe ZrL3 as a hybrid EEL. Although the mobilities of the ZrL3:bis-C60
respectively the hydrolytic and thermal stability, ZrL3 was placed hybrid EEL are similar to those of ZrL3, the reduced energy barrier
in boiling water for 24 hours and it was also heated at 100 °C in air with the cathode will facilitate electron collection (Supplementary
for 7 days. The boiling-water-treated sample still showed substan- Fig. 23).
tial porosity as per gas sorption studies, even though the surface As expected, the PCE of the device with the ZrL3:bis-C60 EEL
area was somewhat decreased (369 m2 g−1, see also Supplementary (denoted as M-PVSC) could be further increased to 22.02% with
Figs. 14 and 15 for the density functional theory analysis of the N2 an FF of 81.28%, open-circuit voltage (Voc) of 1.20 V and negligible
sorption data). Moreover, the PXRD pattern of ZrL3 after either hysteresis. This is among the highest values reported for the pla-
boiling-water-treatment or continual heating at 100  °C in air nar inverted PVSCs (Fig. 2c, Supplementary Figs. 24 and 25 and
remained strong, indicating the retention of the crystalline phase Table 2). The short-circuit current density (Jsc) obtained from the
(Fig. 1f and Supplementary Fig. 16). These results demonstrated the J–V curve matched well with the integrated value from the external
superior moisture and thermal stability of ZrL3. The optical band- quantum efficiency (EQE) (Fig. 2d), demonstrating the good reli-
gap of ZrL3 was determined and calculated to be 2.53 eV from its ability of the results. The ‘champion’ (best) cell of M-PVSC was fur-
ultraviolet-visible (UV-vis) diffuse reflectance spectrum, as shown ther held at the maximum power point (MPP) with a bias voltage
in Fig. 1g. Additionally, the absorption spectra were also obtained of 1.03 V under continuous 1 sun illumination (AM 1.5G, 100 mW
to study the effect of ZrL3 on light harvesting of the perovskite layer cm−2) to obtain a stabilized photocurrent of 20.97 mA cm−2, which
(Supplementary Fig. 17). No obvious change could be found in the corresponds to a stabilized PCE of 21.60% (Fig. 2e). More impor-
absorption spectra with varied thicknesses of ZrL3 and the bandgap tantly, the efficiency of the encapsulated M-PVSC was certified
of the mixed-cation (Cs/formamidinium (FA)/methylammonium to be 21.3% by an independent accredited institute (the National
(MA)) perovskite was calculated to be 1.60 eV from the Tauc plot. Institute of Metrology, China), which is close to the value measured
The electron- and hole-only devices were fabricated to study the in house (Supplementary Fig. 26).
electrical properties of ZrL3 by the space-charge-limited-current To elucidate the potential roles ZrL3 and bis-C60 played in the
method (Supplementary Figs. 18 and 19 and Note 3)24. As a result, hybrid EEL for device efficiency, the PVSCs using only bis-C60
the electron- and hole-mobilities were calculated to be 2.81 × 10−7 as the EEL were fabricated as the reference device (denoted as
m2 V−1 s−1 and 5.36 × 10−9 m2 V−1 s−1, respectively. The relatively r-PVSC), which affords a slightly lower PCE of 21.32% than that of
higher electron mobility indicated that ZrL3 is more of a n-type M-PVSC (Supplementary Fig. 27 and Table 3). The device perfor-
material, which would be appropriate for extracting electrons mance of the PVSCs with different EELs is summarized in Fig. 2f
in PVSCs. To investigate the interfacial charge carrier dynam- and Supplementary Tables 3–5. There are no obvious enhancements
ics between perovskite and different EELs, steady-state photolu- in Voc and Jsc, while the enhanced PCE for M-PVSCs mainly comes
minescence (PL) and time-resolved photoluminescence (TRPL) from the FF improvements (Supplementary Fig. 28). To explore the
spectra were obtained for the perovskite films coated with ZrL3, effect of EEL morphology on the improved FF, atomic force micros-
phenyl-C61-butyric acid methyl ester (PC61BM) and PC61BM/ZrL3 copy was employed to characterize the morphology of bis-C60,
layers, respectively. As can be seen in Fig. 1h, the pure perovskite ZrL3 and ZrL3:bis-C60 prepared on perovskite/PC61BM film. The
film showed the strongest PL intensity due to the efficient radia- roughness of perovskite/PC61BM was slightly reduced from 10.02 to
tive recombination of the photogenerated carriers, while it was 8.86 nm after coating with bis-C60 (Supplementary Fig. 29).
apparently quenched upon contacting with the EELs. We note By contrast, the film roughness of both perovskite/PC61BM/
that the perovskite film showed substantially quenched PL when ZrL3 and perovskite/PC61BM/ZrL3:bis-C60 films were increased
it contacted with ZrL3, suggesting efficient charge extraction from to 36.31 nm and 35.45 nm, respectively and the morphologies of
perovskite to ZrL3 (ref. 25,26). This phenomenon is further con- perovskite/PC61BM/ZrL3 and perovskite/PC61BM/ZrL3:bis-C60
firmed by the TRPL spectra, as shown in Fig. 1i and summarized were quite similar. These results implied that the film morphol-
in Supplementary Table 1. The perovskite/ZrL3 film exhibited biex- ogy of hybrid EEL was dominated by ZrL3 and it should not be the
ponential decay with a fast (1.6 ns) and a slow (13.8 ns) component reason for the varied FF between the devices with different EELs.
after being excited by a pulsed laser. This decay is much shorter Furthermore, given that the hole-blocking ability of the EEL can sig-
than for the perovskite-only sample (137.8 and 632.8 ns). In addi- nificantly affect the FF in photovoltaic devices, the energy levels of
tion, the perovskite/PC61BM/ZrL3 film exhibited a slightly faster PL ZrL3, bis-C60 and ZrL3:bis-C60 were then calculated and are sum-
decay than the perovskite/PC61BM film, which might be attributed marized in Supplementary Table 6 (also Supplementary Figs. 30 and
to the cascade energy alignment in perovskite/PC61BM/ZrL3 (see 31 and Note 4). It was found that bis-C60 possesses a larger energy
Supplementary Fig. 20 and Note 4). barrier for the holes than ZrL3. Therefore, it can potentially provide
better hole-blocking ability which is consistent with the higher FF
Device architecture and photovoltaic performance observed in r-PVSC than the devices with ZrL3 EEL alone.
The inverted planar PVSCs with the device configuration of indium On the other hand, it is known that the thiol group can inter-
tin oxide (ITO) glass/poly(triarylamine) (PTAA)/perovskite/ act with Ag by forming S–Ag bonds, which can reduce the contact
PC61BM/EEL/silver (Ag) were fabricated and the corresponding resistance at the interface31–34. Thus, it is expected that the S–Ag
energy alignment is provided in Fig. 2a,b. The device without any bonds would form between the dense thiol and disulfide groups
EEL showed a relatively low PCE of 18.80% with a fill factor (FF) of in ZrL3 and the Ag electrode to enable M-PVSCs to have better
71.21%, which was attributed to the non-ideal electron extraction EEL-Ag contact than r-PVSCs. This allows M-PVSCs to have lower
from PC61BM to the electrode and the insufficient hole-blocking contact resistance and higher FF, which is consistent with our pre-
ability. Encouragingly, by introducing ZrL3 as the EEL between vious report33. Based on these considerations, the high efficiency

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

a b c
25

Current density (mA cm–2)


20
3.9
4.1

Energy level (–eV)


60 4.3
ZrL3:bis-C 15 Voc = 1.20 V

PTAA
PC 61BM Jsc = 22.58 mA cm–2

Perovskite
Ag
ITO

PC61BM
FF = 81.28%

ZrL3:bis-C60
10
PCE = 22.02%
Perovskite 4.8
5.2 5
PTAA 5.5
te
ITO substra 0
0 0.2 0.4 0.6 0.8 1.0 1.2
Voltage (V)

d e f
100 25 25 12
Average PCE = 20.85%
80 20 10
20
Stablized at 20.97 mA cm–2
8
Integrated Jsc

J (mA cm–2)

60 15
EQE (%)

15

Counts
6
40 10 10
4
20 5 5
EQE 2
Integrated Jsc In N2-filled glovebox
0 0 0
300 400 500 600 700 800 0 10 20 30 40 50 60 19 20 21 22 23
Wavelength (nm) Time (min) PCE (%)

Fig. 2 | Device structure and performance of the PVSCs with ZrL3:bis-C60 EEL. a, Device structure of inverted PVSC. b, Energy level alignment of the used
materials. c, J–V curve of the champion PVSC with ZrL3:bis-C60 EEL. d, EQE spectrum and integrated current density of the champion PVSC with ZrL3:bis-C60
EEL. e, Stabilized power output at the MPP for the PVSC with ZrL3:bis-C60 EEL. f, PCE histogram obtained from 30 PVSCs with ZrL3:bis-C60 EEL.

achieved for M-PVSCs should be mainly attributed to bis-C60, Long-term device stability
while ZrL3 with the dense thiol groups plays a role in reducing the To verify the feasibility of using a ZrL3:bis-C60 EEL in enhancing
contact resistance, thus slightly improving the FF. device stability, a comprehensive study was carried out on the encap-
The resultant PVSCs were also characterized by light intensity (I) sulated PVSCs in both humid air and nitrogen atmosphere. The
dependent photovoltage and photocurrent density. Supplementary M-PVSC showed superior shelf stability in an ambient environment
Fig. 32 shows the linear relationships between Jsc and I in (with a RH of 75%), maintaining 90% of its initial efficiency after
double-logarithmic scale and the fitted α values are close to 1, which 1,100 hours (Fig. 3a). By contrast, the PCE of r-PVSC dropped sig-
demonstrates the efficient carrier transport and the few bimolecular nificantly to <50% of its original value. The significantly enhanced
recombinations for both devices. Additionally, the slopes were cal- stability can be attributed to the strong coordinating ability of Zr(iv)
culated from the semi-natural logarithmic plot of Voc versus I. The ions and the dense thiol and crosslinked disulfide linkages in ZrL3,
M-PVSC exhibited a slightly lower slope (1.41 kBT/q) than r-PVSC which protect perovskites against moisture and oxygen38.
(1.57 kBT/q), where kB, T and q are Boltzmann’s constant, absolute Long-term operational stability of the devices was tested under
temperature and elementary charge, respectively. These results a white LED lamp with the light intensity equivalent to 1 sun in
indicate that the trap-assisted Shockley–Read–Hall recombination a N2-filled glovebox (Fig. 3b). Specifically, the operational stability
is reduced in M-PVSC because of the better contact between ZrL3 was tested under an accelerated ageing condition: (1) no UV filter
and the Ag electrode, which contributes to faster charge extraction for the light source and (2) the devices were monitored at an ele-
in M-PVSC. vated temperature of 85 °C to estimate their potential for long-term
In a real-use scenario, perovskites and CTLs in a complete PVSC stability. The ZrL3-incorporated cell was tracked at the MPP and
are susceptible to hydration and oxidation due to abundant mois- the corresponding PCEs obtained from the J–V curve were con-
ture and an oxygen atmosphere35–37. To highlight the superior sta- stantly monitored during the test (Fig. 3c). The PCE of M-PVSC
bility of ZrL3, perovskite/PC61BM films were prepared initially on only dropped slightly to 92% of its initial value after 1,000 hours of
glass and coated with either bis-C60 or ZrL3:bis-C60 hybrid EEL ageing. Thus, M-PVSC is among the most stable inverted PVSCs
to investigate their ability to protect perovskites against moisture tested under the MPP tracking condition at 85 °C. Moreover, it is
and oxygen. The perovskite/PC61BM film with only bis-C60 EEL observed that the constantly recorded PCEs from the J–V curves of
bleached almost completely after being placed in an ambient envi- M-PVSC during MPP tracking showed a similar decay trend, dem-
ronment with a relative humidity (RH) of 75% for over two weeks onstrating the good reliability of the testing (Fig. 3c). In the litera-
(Supplementary Fig. 33). In sharp contrast, no obvious degrada- ture, small-molecule based CTLs/EELs tend to crystallize and form
tion was observed for the perovskite/PC61BM film protected by the aggregates under thermal stress. This creates pathways for direct
ZrL3:bis-C60 hybrid EEL. Besides, the contact angle with a drop interactions between perovskite and metal electrode and so impairs
of water on top of ZrL3:bis-C60 protected perovskite/PC61BM film the device performance37. The superior stability of ZrL3 under ther-
rose from 68.37 to 81.23° (Supplementary Fig. 34), implying ZrL3 mal stress significantly enhanced the thermal stability of M-PVSC,
can indeed help prevent moisture from penetrating into perovskite which was further verified by the stability profile of the PVSCs with
film, which is consistent with its enhanced film stability. only ZrL3 as the EEL (Supplementary Fig. 35).

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
a d
Ag ETL Perovskite HTL ITO Ag ETL Perovskite HTL ITO
20 ZrL3:bis-C60
105 Bis-C60
105
I I
16 104 104
PCE (%)

Intensity

Intensity
12 103 Ag 103 PbI2
PbI2 Ag
2
8 10 102
M-PVSC
r-PVSC In humid air 101 101 Zr
4
0
10 100
0 150 300 450 600 750 900 1,050
0 500 1,000 1,500 0 500 1,000 1,500
Time (h)
Sputtering time (s) Sputtering time (s)

b e Ag ETL Perovskite HTL ITO


Ag ETL Perovskite HTL ITO
1.0 105 ZrL3:bis-C60
105 Bis-C60
Normalized PCE

I I
0.8 104 104

Intensity
Intensity
0.6 103 Ag
PbI2
103 Ag PbI2

M-PVSC 10 2
102
0.4
r-PVSC MPP in N2-filled glovebox 101 101 Zr
0.2 Lead leakage
10 0
100
0 200 400 600 800 1,000
0 500 1,000 1,500 0 500 1,000 1,500
Time (h)
Sputtering time (s) Sputtering time (s)

c f
H2O

20 H2O

H2O

16
PCE (%)

Lead immobilization
Zr
L3

via S–Pb bond


12
:b C 61
is BM
P

-C
60

8
Pe
ro
vs
kit

Pb Pb
e

Pb
C O S Pb Zr
4
0 200 400 600 800 1,000
Time (h)

Fig. 3 | Long-term stability studies for the PVSCs with bis-C60 (r-PVSC) or ZrL3:bis-C60 (M-PVSC) EEL. a, Shelf stability of the encapsulated PVSCs in
humid air with RH of 75%. b, Long-term operational stability of the encapsulated PVSCs under MPP tracking at 85 °C under continuous light irradiation
with a white LED lamp (100 mW cm−2) in a N2-filled glovebox. c, PCEs constantly recorded from the J–V curves of M-PVSC during MPP tracking. d,e,
ToF-SIMS depth profiles of the pristine (d) and aged (e) PVSCs with bis-C60 (left) or ZrL3:bis-C60 (right) EEL. f, Schematic of degradation process of
PVSCs and the immobilization effect of ZrL3 on leaked Pb2+ ions.

Lead contamination of PVSCs deionized water with a pH value of ~5.6 (simulating acidic rain)
To probe the possible sources and mechanisms that cause the deg- to compare the concentration of water soluble Pb components
radation in our devices, the depth profiles of Pb, Ag and I ions between M-PVSCs and r-PVSCs, as shown in Fig. 4c. Pb2+ testing
throughout the un-encapsulated PVSCs were investigated by using paper was used to perform a quick check on the Pb2+ concentra-
time-of-flight secondary-ion mass spectrometry (ToF-SIMS) (Fig. tion in the contaminated water and it was found that the aqueous
3d,e). The Pb2+ signal could be clearly observed to shift towards extracts from the r-PVSC showed a much higher Pb2+ concentration
the Ag electrode through the electron-transporting layer in the compared with those of the M-PVSC.
r-PVSC after it was aged under the accelerated condition in air An inductively coupled plasma optical emission spectroscopy
(85 °C and 85% RH). This indicates that the perovskite films can (ICP-OES) instrument was further used to quantify accurately
be easily decomposed into Pb-containing by-products and ammo- the Pb2+ concentrations (Fig. 4d and Supplementary Fig. 36).
nium salts under such an aggressive ageing condition, resulting in The Pb2+ concentration decreased drastically to an average value
Pb2+ diffusion into other layers35,39. Very encouragingly, almost no of 7.6 ppm for M-PVSCs compared to 38.4 ppm for r-PVSCs and
Pb2+ ions could be detected in the Ag layer of M-PVSC and they 38.1 ppm for pure perovskite film, which means over 80% of the
were “trapped” at the ZrL3:bis-C60/Ag interface to form Pb(ii)– leaked Pb2+ ions from the degraded perovskite could be captured
ZrL3 complexes. Unlike the physical encapsulation methods used by thiol-functionalized ZrL3 to form water-insoluble Pb(ii)–ZrL3
in reducing lead leakage, this in situ chemical adsorption of Pb2+ by complexes. This substantially reduced lead leakage was attributed
ZrL3:bis-C60 EEL in the device is a much more effective and sus- to the chemical reaction between the dense array of thiol and
tainable process for long-term practical applications (Fig. 3f). disulfide groups on ZrL3 with leaked Pb2+ ions from the degraded
To quantify the adsorption ability of ZrL3 for Pb2+ ions, sorption perovskites. Although this Pb(ii)-trapping process can be further
kinetics and sorption isotherm measurements were obtained (Fig. optimized, our results demonstrate a promising strategy to improve
4a,b and Supplementary Notes 5 and 6). A distribution coefficient simultaneously PVSC long-term operational stability and reduce/
Kd of around 105 ml g−1, a substantial value of 355 mg g−1 for the prevent Pb2+ leakage into the environment. This will greatly facili-
Pb adsorption capacity (qmax) and a sorption rate constant (pseudo tate the large-scale deployment of highly efficient PVSCs for clean
first-order) of 0.103 min−1 were achieved. These results strongly energy applications.
support the idea that ZrL3 EEL can effectively adsorb Pb2+ by rapid
chemical reactions between thiols and Pb2+ to minimize the leak- Conclusion
age of Pb2+ ions into the environment. To investigate the lead leak- We have developed an effective strategy to address simultaneously
age that occurred in the degraded PVSCs, they were immersed in the most challenging operational stability and lead contamination

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

a b
10 25
250

Absorbed amount (mg g–1)


20

Absorbed amount (mg g–1)


8 1.2 2.8
Pb concentration (ppm)

200
15

ln(Ct)
0.8 2.4

Ce/qe (g L–1)
6 10
0.4 150 2.0

5 1.6
4 0.0
100
0 2 4 6 8 10 12 14 16 1.2
0 t (min)
0.8
2 50
0 60 120 180 240 300 360 420
Time (min) 150 300 450 600 750
Ce (mg L–1)
0 0
0 60 120 180 240 300 360 420 0 150 300 450 600 750 900

Time (min) Equilibrium concentration (Ce) (mg L–1)

c d
60
r-PVSC
Lead concentration (ppm)
50 M-PVSC
0 ppm

40

M-PVSC M-PVSC
10 ppm 30

20
20 ppm
10

r-PVSC r-PVSC 0
40 ppm
1 2 3 4 5 6 7 8 9 10 11 12
Immersed into deionized water
Devices

Fig. 4 | Lead contamination of the PVSCs with bis-C60 (r-PVSC) or ZrL3:bis-C60 (M-PVSC) EEL. a, Pb(ii) sorption kinetics of an activated ZrL3 sample
(20.0 mg) at an initial Pb(ii) concentration of 10.0 ppm. The two insets are the corresponding absorbed Pb(II) amount and the Lagergren first-order kinetic
plot of the Pb(II) sorption, respectively. b, Pb(ii) Langmuir sorption isotherm of the activated sample of ZrL3. The inset is the linear fit with the Langmuir
adsorption model. c, Degraded PVSCs with bis-C60 (r-PVSC) or ZrL3:bis-C60 (M-PVSC) EEL are immersed into the deionized water with a pH value of
~5.6 and Pb ion testing paper is used to quickly detect the Pb concentration in the contaminated water. d, Pb concentration in the contaminated water was
accurately detected by ICP-OES measurements.

issues in PVSCs by incorporating a functionalized 2D MOF as 4. Yoshikawa, K. et al. Silicon heterojunction solar cell with interdigitated back
an EEL into a PVSC. The resultant inverted PVSC achieved a contacts for a photoconversion efficiency over 26%. Nat. Energy 2, 17032 (2017).
high PCE of >22% (a certified value of 21.3%) and substantially 5. Kim, M. et al. Methylammonium chloride induces intermediate phase
stabilization for efficient perovskite solar cells. Joule 3, 2179–2192 (2019).
improved long-term operational stability. More importantly, the 6. Bai, S. et al. Planar perovskite solar cells with long-term stability using ionic
thiol-containing MOF can capture most of the leaked Pb2+ (over liquid additives. Nature 571, 245–250 (2019).
80%) from degraded PVSCs by forming a water-insoluble com- 7. Christians, J. A. et al. Tailored interfaces of unencapsulated perovskite solar
plex. These combined results showed the promise of simultane- cells for >1,000 hour operational stability. Nat. Energy 3, 68–74 (2018).
ously realizing stable and environmental-friendly PVSCs through 8. Li, N. et al. Cation and anion immobilization through chemical bonding
enhancement with fluorides for stable halide perovskite solar cells. Nat.
multi-functional MOFs. Energy 4, 408–415 (2019).
9. Wang, L. et al. A Eu3+-Eu2+ ion redox shuttle imparts operational durability to
Online content Pb-I perovskite solar cells. Science 363, 265–270 (2019).
Any methods, additional references, Nature Research report- 10. Niu, T. et al. High performance ambient-air-stable FAPbI3 perovskite solar
cells with molecule-passivated Ruddlesden–Popper/3D heterostructured film.
ing summaries, source data, extended data, supplementary infor-
Energy Environ. Sci. 11, 3358–3366 (2018).
mation, acknowledgements, peer review information; details of 11. Jiang, Y. et al. Reduction of lead leakage from damaged lead halide perovskite
author contributions and competing interests; and statements of solar modules using self-healing polymer-based encapsulation. Nat. Energy 4,
data and code availability are available at https://doi.org/10.1038/ 585–593 (2019).
s41565-020-0765-7. 12. Alharbi, E. A. et al. Atomic-level passivation mechanism of ammonium salts
enabling highly efficient perovskite solar cells. Nat. Commun. 10, 3008 (2019).
13. Chen, J., Zhao, X., Kim, S.-G. & Park, N.-G. Multifunctional chemical linker
Received: 9 December 2019; Accepted: 13 August 2020; imidazoleacetic acid hydrochloride for 21% efficient and stable planar
Published: xx xx xxxx perovskite solar cells. Adv. Mater. 31, 1902902 (2019).
14. Zayed, J. & Philippe, S. Acute oral and inhalation toxicities in rats with
References cadmium telluride. Int J. Toxicol. 28, 259–265 (2009).
1. Jiang, Q. et al. Surface passivation of perovskite film for efficient solar cells. 15. Yu, Y., Hong, Y., Gao, P. & Nazeeruddin, M. K. Glutathione modified gold
Nat. Photon. 13, 460–466 (2019). nanoparticles for sensitive colorimetric detection of Pb2+ ions in rainwater
2. Yoo, J. J. et al. An interface stabilized perovskite solar cell with high stabilized polluted by leaking perovskite solar cells. Anal. Chem. 88, 12316–12322 (2016).
efficiency and low voltage loss. Energy Environ. Sci. 12, 2192–2199 (2019). 16. Hailegnaw, B., Kirmayer, S., Edri, E., Hodes, G. & Cahen, D. Rain on
3. Jung, E. H. et al. Efficient, stable and scalable perovskite solar cells using methylammonium lead iodide based perovskites: Possible environmental
poly(3-hexylthiophene). Nature 567, 511–515 (2019). effects of perovskite solar cells. J. Phys. Chem. Lett. 6, 1543–1547 (2015).

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
17. Wu, H. B. & Lou, X. W. Metal-organic frameworks and their derived 29. Li, C.-Z. et al. Effective interfacial layer to enhance efficiency of polymer solar
materials for electrochemical energy storage and conversion: Promises and cells via solution-processed fullerene-surfactants. J. Mater. Chem. 22,
challenges. Sci. Adv. 3, eaap9252 (2017). 8574–8578 (2012).
18. Li, M. et al. Doping of [In2(phen)3Cl6]·CH3CN·2H2O indium-based 30. Zhu, Z., Chueh, C.-C., Lin, F. & Jen, A. K. Y. Enhanced ambient stability of
metal-organic framework into hole transport layer for enhancing perovskite efficient perovskite solar cells by employing a modified fullerene cathode
solar cell efficiencies. Adv. Energy Mater. 8, 1702052 (2018). interlayer. Adv. Sci. 3, 1600027 (2016).
19. Ryu, U. et al. Nanocrystalline titanium metal–organic frameworks for highly 31. Gan, W., Xu, B. & Dai, H.-L. Activation of thiols at a silver nanoparticle
efficient and flexible perovskite solar cells. ACS Nano 12, 4968–4975 (2018). surface. Angew. Chem. Int. Ed. 50, 6622–6625 (2011).
20. Lee, C. C., Chen, C. I., Liao, Y. T., Wu, K. C. & Chueh, C. C. Enhancing 32. Fenter, P. et al. Structure of octadecyl thiol self-assembled on the silver (111)
efficiency and stability of photovoltaic cells by using perovskite/Zr-MOF surface: an incommensurate monolayer. Langmuir 7, 2013–2016 (1991).
heterojunction including bilayer and hybrid structures. Adv. Sci. 6, 33. Laibinis, P. E. et al. Comparison of the structures and wetting properties of
1801715 (2019). self-assembled monolayers of n-alkanethiols on the coinage metal surfaces,
21. Chang, T. H. et al. Planar heterojunction perovskite solar cells incorporating copper, silver, and gold. J. Am. Chem. Soc. 113, 7152–7167 (1991).
metal–organic framework nanocrystals. Adv. Mater. 27, 7229–7235 (2015). 34. Yip, H.-L., Hau, S. K., Baek, N. S. & Jen, A. K. Y. Self-assembled monolayer
22. Yee, K.-K. et al. Effective mercury sorption by thiol-laced metal–organic modified ZnO/metal bilayer cathodes for polymer/fullerene
frameworks: in strong acid and the vapor phase. J. Am. Chem. Soc. 135, bulk-heterojunction solar cells. Appl. Phys. Lett. 92, 193313 (2008).
7795–7798 (2013). 35. Tiep, N. H., Ku, Z. & Fan, H. J. Recent advances in improving the stability of
23. Wong, Y. L., Diao, Y., He, J., Zeller, M. & Xu, Z. A thiol-functionalized perovskite solar cells. Adv. Energy Mater. 6, 1501420 (2016).
UiO-67-type porous single crystal: Filling in the synthetic gap. Inorg. Chem. 36. Wang, S. et al. Unveiling the role of tBP–LiTFSI complexes in perovskite solar
58, 1462–1468 (2019). cells. J. Am. Chem. Soc. 140, 16720–16730 (2018).
24. Ren, H. et al. Efficient and stable Ruddlesden–Popper perovskite solar 37. Li, M. et al. Interface modification by ionic liquid: A promising candidate for
cell with tailored interlayer molecular interaction. Nat. Photon. 14, indoor light harvesting and stability improvement of planar perovskite solar
154–163 (2020). cells. Adv. Energy Mater. 8, 1801509 (2018).
25. Lu, J. et al. Interfacial benzenethiol modification facilitates charge transfer 38. Chen, W., Xu, L., Feng, X., Jie, J. & He, Z. Metal acetylacetonate series in
and improves stability of cm-sized metal halide perovskite solar cells with up interface engineering for full low-temperature-processed, high-performance,
to 20% efficiency. Energy Environ. Sci. 11, 1880–1889 (2018). and stable planar perovskite solar cells with conversion efficiency over 16%
26. Tan, F. et al. In situ back-contact passivation improves photovoltage and fill on 1 cm2 scale. Adv. Mater. 29, 1603923 (2017).
factor in perovskite solar cells. Adv. Mater. 31, e1807435 (2019). 39. Yang, J. & Kelly, T. L. Decomposition and cell failure mechanisms in lead
27. Zhu, Z., Chueh, C.-C., Li, N., Mao, C. & Jen, A. K. Y. Realizing efficient halide perovskite solar cells. Inorg. Chem. 56, 92–101 (2017).
lead-free formamidinium tin triiodide perovskite solar cells via a sequential
deposition route. Adv. Mater. 30, 1703800 (2018). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
28. Zhu, Z. et al. Highly efficient and stable perovskite solar cells enabled by published maps and institutional affiliations.
all-crosslinked charge-transporting layers. Joule 2, 168–183 (2018). © The Author(s), under exclusive licence to Springer Nature Limited 2020

Nature Nanotechnology | www.nature.com/naturenanotechnology


Articles Nature Nanotechnology

Methods ZrOCl2∙8H2O (92 mg, 0.29 mmol) and formic acid (≥ 88%, 490 mg, 9.37 mmol)
Materials. The starting materials, reagents and solvents for MOF synthesis were were ultrasonically dissolved in another 3.0 ml DMF. The two mixtures were added
purchased from commercial sources, including J&K, Sigma-Aldrich and Anaqua to a 25 ml Pyrex glass ampoule. The ampoule was flame-sealed and heated in an
and were used without further purification. Caesium iodide (CsI), formamidinium oven at 120 °C for 48 h, followed by programmed cooling to room temperature for
iodide (FAI) and methylammonium bromide (MABr) were purchased from 12 h. The pale brown powder formed was collected by filtration, then washed with
Dysol. Lead iodide (PbI2) and lead bromide (PbBr2) were purchased from TCI. DMF (3 ml × 10) and acetone (3 ml × 10) and dried under nitrogen flow for weight
Phenyl-C61-butyric acid methyl ester (PC61BM) was purchased from ADS. measurement (90 mg). The above powder was denoted as as-made ZrL3.
Poly[bis(4-phenyl)(2,4,6-trimethylphenyl) amine] (PTAA) and bathocuproine
(BCP, purity of 99.9%) were purchased from Xi’an Polymer Light Technology Preparation of activated ZrL3. In a nitrogen-filled glovebox, an as-made ZrL3
Corporation. The solvents, including DMF, dimethyl sulfoxide (DMSO), sample (~80 mg) and DMF (~8 ml) were added to a 15 ml glass vial and heated at
isopropanol (IPA) and chlorobenzene (CB) were purchased from J&K and used as 60 °C for 12 h. After cooling to room temperature, the supernatant was replaced
received. Toluene was purchased from Tian Hang Technology Co., Ltd. Silver with with fresh DMF (~8 ml) and heated at 60 °C for 12 h. The above solvent exchange
high purity was purchased from commercial sources. and heating procedures were repeated three times. After that, the above activation
procedure was repeated using CH3CN. The vacuum-dried powder was denoted as
Synthesis of methyl 2,6-bis(benzylthio)-4-bromobenzoate (S1). 4-Bromo- activated ZrL3.
2,6-difluorobenzoic acid (4.74 g, 0.02 mol) and potassium carbonate (27.6 g, Elemental analysis found (C (26.70%), H (4.286%), N (0.47%)) and a
0.20 mol) were loaded in a 250 ml two neck round-bottom flask. N-Methyl- fitting formula can be determined to be Zr6O4(OH)4(L3)2(OH)6(H2O)27(DMF)
2-pyrrolidone (100 ml) was added to the flask and the mixture was stirred and (molecular mass 2,596.45), which gives a calculated profile as (C (26.37%), H
bubbled with nitrogen at room temperature for 30 min. Benzyl mercaptan (6.0 ml, (3.92%), N (0.54%), S (14.82%), Zr (21.08%)). The activated ZrL3 sample was
0.05 mol) was then added to the mixture under nitrogen. The mixture was stirred also characterized by TGA in both N2 and air flows (Supplementary Fig. 8, see
at 80 °C for 4 days. After cooling to room temperature, methyl iodide (3.0 ml, also Supplementary Fig. 9 for an informative comparison of TGA data). The TGA
48.19 mmol) was added to the mixture and the mixture was stirred at room measurement conducted in an air flow showed 25.1% ZrO2 residue, equivalent
temperature for 30 min. The resulting mixture was poured into distilled water to a Zr content of 18.6%. Both TGA and elemental analysis results agree with the
(800 ml) and extracted with dichloromethane (DCM) (3 × 200 ml). The combined evacuated, dehydrated formula Zr6O8(C27H15O6S6)2O (molecular mass 1947).
organic layer was dried over anhydrous MgSO4, evaporated on a rotary evaporator
and purified by column chromatography (eluent: DCM/hexane, 1:5) to afford S1 Boiling-water-treatment of ZrL3. To a 100 ml round-bottom flask charged with a
as a: white solid (6.96 g, 76% yield based on 4-bromo-2,6-difluorobenzoic acid); 1H stirring bar, an activated ZrL3 sample (120 mg) and deionized water (30 ml) were
NMR (300 MHz, CDCl3): δ = 7.34–7.18 (m, 12H), 4.05 (s, 4H), 3.91 (s, 3H) ppm; added and the mixture was refluxed (using an oil bath at 120 °C) for 24 h. After
13
C NMR (75 MHz, CDCl3): δ = 167.35, 138.72, 136.35, 135.67, 133.41, 129.23, cooling to room temperature, the resultant MOF solid was collected by filtration
128.69, 127.70, 123.22, 52.70, 40.48 ppm. and dried under vacuum at 105 °C for 12 h. The above powder was denoted as
boiling-water-treated ZrL3.
Synthesis of dimethyl 1,3,5-tris[3′,5′-bis(benzylthio)-4′-(methoxycarbonyl)
phenyl] benzene (M3). Molecule S1 (2.00 g, 4.36 mmol), bis(pinacolato)diboron Perovskite solar cell fabrication. The pre-patterned ITO (15 Ω sq−1) glass
(1.107 g, 4.36 mmol), PdCl2(PPh3)2 (77 mg, 0.11 mmol) and anhydrous potassium substrates were sequentially cleaned by sonication with detergent (Decon 90/
acetate (941 mg, 9.59 mmol) were loaded in a 250 ml two neck round-bottom deionized water, v/v = 1:1), deionized water, acetone and IPA for 15 min,
flask and dried under vacuum for 30 min. The flask was then connected to a respectively. The cleaned ITO glass substrates were then transferred into an
nitrogen atmosphere. Anhydrous 1,4-dioxane (30 ml, bubbled with nitrogen for oven at a temperature of 100 °C and treated with O2 plasma for 10 min before
10 min beforehand) was added to the flask. The mixture was stirred at 90 °C for use. Then, the hole-transporting layer, that is, 10 mg ml−1 PTAA in toluene
12 h. After cooling to room temperature, K3PO4 aqueous solution (2.3 M, 5.0 ml (doped with 1 wt% of 2,3,5,6-tetrafluoro-7,7,8,8-tetracyanoquinodimethane
bubbled with nitrogen for 10 min beforehand) was added, followed by the addition (F4-TCNQ)), was spin-coated onto the ITO glass substrate at 4,000 rpm for 30 s
of 1,3,5-tribromobenzene (400 mg, 1.27 mmol). The mixture was stirred at 90 °C and this was subsequently annealed at 150 °C for 10 min. After cooling for 5 min,
for 12 h. After cooling to room temperature, the mixture was poured into water the mixed-cation perovskite film was then deposited on PTAA by a one-step
(400 ml) and extracted with DCM (3 × 100 ml). The combined organic phases spin-coating method. The perovskite precursor solution was prepared by mixing
were washed with distilled water (3 × 200 ml). The organic phase was then dried CsI, FAI, MABr, PbI2 and PbBr2 in DMF/DMSO (v/v = 5:1) to a chemical formula
over anhydrous MgSO4 and evaporated with a rotary evaporator. The solid residue of Cs0.05(FA0.85MA0.15)0.95Pb(I0.85Br0.15)3.
obtained was purified by flash column chromatography (eluent: DCM/n-hexane/ Considering the poor wetting ability of the PTAA surface, 50 μl DMF was first
ethyl acetate, 1:1:0.01) to obtain M3 as a: white solid (987 mg, 64% yield based on spun onto the PTAA layer to improve the wetting ability before depositing the
1,3,5-tribromobenzene); 1H NMR (300 MHz, CDCl3): δ = 7.27 (s, 6H), 7.16–7.26 perovskite layer. Then, 60 μl of perovskite precursor solution was spin-coated on
(m, 30H), 6.88 (s, 3H), 4.11 (s, 12H), 4.03 (s, 9H) ppm; 13C NMR (75 MHz, CDCl3): PTAA at 1,000 rpm for 10 s and 5,000 rpm for 30 s. During the second step, 110 μl
δ = 168.01, 142.17, 141.46, 140.49, 137.12, 133.02, 132.58, 129.26, 128.58, 127.58, CB as anti-solvent was quickly dripped onto the centre of the perovskite film 15 s
125.36, 52.70, 41.11 ppm. before the end of the spin-coating process and then annealed at 60 °C for 5 min and
100 °C for 30 min. The electron-transporting layer, that is PC61BM (20 mg ml−1 in
Synthesis of 1,3,5-tris[3′,5′-bis(benzylthio)-4′-carboxyphenyl]benzene (MA3). CB), was spun onto the perovskite layer with a speed of 2,000 rpm for 30 s, followed
Molecule M3 (485 mg, 0.40 mmol) was dissolved in tetrahydrofuran (7.0 ml) in a by annealing at 90 °C for 10 min. As for the devices with a MOF EEL, ZrL3 solution
100 ml two neck round-bottom flask, followed by adding KOH solution (8.9 M, was first prepared by dispersing ZrL3 into IPA at different concentrations followed
7.0 ml in MeOH/H2O, v-v = 1:1). The mixture was stirred at 70 °C for 24 h. After by ultrasonication and stirring. Then, the ZrL3 (or ZrL3 doped with 10 wt% of
cooling to room temperature, 10% HCl(aq.) was added to the resulting mixture to bis-C60) solution was spun onto the PC61BM layer at a speed of 2,000 rpm for 60 s.
attain pH < 2. A yellow precipitate formed was filtered, washed extensively with For the reference devices, bis-C60 (2 mg ml−1 in IPA) was spun onto the PC61BM
distilled water and suction-dried on a filter paper to afford MA3 as a: brown solid with a speed of 2,000 rpm for 30 s. All the spin-coating processes were conducted in a
(455 mg, 97% yield based on M1); 1H NMR (300 MHz, DMSO-d6): δ = 7.55 (s, 6H), N2-filled glovebox with contents of O2 and H2O < 1 ppm at a controlled temperature
7.52 (s, 3H), 7.16–7.26 (m, 18H), 7.32–7.35 (m, 12H), 4.35 (s, 12H) ppm; 13C NMR of 20–25 °C. Finally, 120 nm Ag was thermally evaporated in a high vacuum chamber
(75 MHz, DMSO-d6): δ = 168.25, 140.59, 140.33, 138.13, 137.19, 133.64, 129.09, (<5 × 10−6 Torr) through a metal shadow mask (aperture area of 0.13 cm2 is defined
128.45, 127.23, 125.52, 37.95 ppm. by the overlap between the pre-patterned ITO and the back metal electrode).

Synthesis of 1,3,5-tris(3′,5′-bimercapto-4′-carboxyphenyl)benzene (H3L3). Characterizations and measurements. Elemental analysis was performed
Molecule MA3 (234 mg, 0.20 mmol), anhydrous AlCl3 (640 mg, 4.80 mmol), dry with a Vario EL III CHN elemental analyser. Infrared spectra in the range
DCM (17.0 ml) and dry toluene (2.0 ml) were mixed in a reaction tube in a N2-filled 400–4,000 cm−1 were recorded on a Nicolet Avatar 360 FT-IR spectrophotometer.
glove box. After the reaction tube was taken out, the tube was connected to a N2 Solution 1H NMR and 13C NMR spectra were recorded at room temperature on
manifold and stirred at room temperature for 2 h. Afterwards, 10% HCl (aq., 10 ml, Bruker superconducting-magnet high-field NMR spectrometers with working
bubbled with N2 for five min beforehand) was added to the mixture, followed by frequencies of 300 and 400 MHz, using tetramethylsilane as the internal standard.
stirring at room temperature for 2 h. The yellow precipitate formed was collected Chemical shifts (δ) are expressed in ppm relative to the residual solvent reference
by suction filtration and washed with 10% HCl (aq.), distilled water, then DCM (for example chloroform-d 1H: 7.26 ppm, 13C: 77.0 ppm). Coupling constants are
and dried under vacuum to afford H3L3 as a: brown solid (103 mg, 82% yield based expressed in hertz. TGA was carried out in a nitrogen or air stream using Perkin
on MA3); 1H NMR (400 MHz, acetone-d6): δ = 8.01 (s, 3H), 7.78 (s, 6H) ppm; 1H Elmer thermal analysis equipment (STA 6000) with a heating rate of 2 °C min−1,
NMR (400 MHz, DMF-d7): δ = 8.11 (s, 3H), 7.96 (s, 6H) ppm; 13C NMR (100 MHz, with an empty Al2O3 crucible being used as the reference. The porosity and
DMF-d7): δ = 169.14, 142.78, 141.27, 137.32, 129.61, 127.05, 126.85 ppm. surface area analyses were performed using a Quantachrome Autosorb iQ gas
sorption analyser. Samples were outgassed at 0.03 Torr with a 2 °C min−1 ramp to
Polycrystalline framework solid of ZrL3. H3L3 (30 mg, 0.048 mmol) and 10 drops 120 °C and held at 120 °C for 12 h. The samples were then held in vacuum until
of 1,2-ethanedithiol (170 mg, 1.81 mmol) were dissolved in DMF (3.0 ml), while the analyses started to run. Pore analysis was performed using N2 at 77.35 K (P/P0

Nature Nanotechnology | www.nature.com/naturenanotechnology


Nature Nanotechnology Articles
range of 2 × 10−7–0.995). PXRD data were collected in the reflection mode at room References
temperature on a Philips X’Pert diffractometer equipped with a CPS 180 detector 40. Cardona, C. M., Li, W., Kaifer, A. E., Stockdale, D. & Bazan, G. C.
using monochromated Cu-Kα (λ = 1.5418 Å) radiation. The X-ray tube operated Electrochemical considerations for determining absolute frontier orbital
at a voltage of 30 kV and a current of 30 mA. The amounts of the metal ions were energy levels of conjugated polymers for solar cell applications. Adv. Mater.
determined using a PerkinElmer Optima™ 8000 ICP-OES instrument. 23, 2367–2371 (2011).
Scanning transmission electron microscopy images were obtained on a 41. Wang, T. C. et al. Rendering high surface area, mesoporous metal-organic
double-aberration-corrected JEOL JEM-300CF instrument at 300 kV. UV-vis frameworks electronically conductive. ACS Appl. Mater. Interfaces 9,
absorption spectra were recorded on a Perkin Elmer UV-vis spectrometer 12584–12591 (2017).
(model Lambda 2S). PL and TRPL spectra were recorded on an FLS980
spectrofluorometer (Edinburgh Instruments) with a Xe lamp and a pulsed
excitation laser, respectively. Ultraviolet photoelectron spectrum studies were Acknowledgements
carried out on a VG ESCALAB 220i-XL surface analysis system equipped with a The work was supported by the APRC Grant of the City University of Hong Kong
He discharge lamp (hv = 21.22 eV). The morphology of the samples was monitored (9380086, 9610421), an ECS grant from the Hong Kong Research Grants Council
by SEM (QUATTRO S, Thermal Fisher Scientific). Atomic force microscopy was (21301319) and Innovation and Technology Support Programme (ITS/497/18FP,
carried out using a Nanoscope IIIa instrument (Veeco). Cyclic voltammetry was GHP/021/18SZ), the Guangdong Major Project of Basic and Applied Basic Research
measured on a CH Instruments CHI760E electrochemical analyser/workstation (no. 2019B030302007) and Guangdong-Hong Kong-Macao Joint Laboratory of
using a conventional three-electrode cell with an ITO strip (5 mm × 35 mm) coated Optoelectronic and Magnetic Functional Materials (no. 2019B121205002). TEM work
with ZrL3 as the working electrode, a Pt wire counter electrode and a Ag/AgCl was conducted using the facilities in the Irvine Materials Research Institute (IMRI) at the
reference electrode at a scan rate of 50 mV s−1 using 0.1 M tetrabutylammonium University of California-Irvine and supported by the NSF under grants (CBET-1159240
hexafluorophosphate in dry DCM as the electrolyte and referenced to the FeCp2/ and DMR-1506535). This work was also supported by an ARG grant (Project 9667168)
FeCp2+ (FeCp2 = Fe(C5H5)2) redox couple (−5.10 eV below vacuum level)40. from the City University of Hong Kong.
The working electrode was prepared using a similar method as reported in the
literature41 by casting a dispersion of ZrL3 in isopropanol onto the ITO glass.
ToF-SIMS measurement was perfomed using a TOF-SIMS V instrument (ION‐
Author contributions
Z.Z. and A.K.-Y.J. conceived the ideas and designed the project. Z.Z., A.K.-Y.J. and Z.X.
TOF GmbH, Cameca IMS 4F), where a 3 keV Cs+ ion beam was used for erosion
directed and supervised the research. S.W. fabricated and characterized devices. Z.L.,
and a 25 keV Bi+ pulsed primary ion beam was used for the analysis. The area of
T.L. and J.Z. also contributed to device fabrication, characterizations and helped S.W.
analysis was 48 × 48 µm2 while the sputtering area was 300 × 300 µm2.
analyse the data. Z.X., M.-Q.L. and Y.D. designed and synthesized the materials. P.T. and
The current density–voltage (J–V) characteristics of the photovoltaic devices
W.G. performed the TEM measurement and P.T., W.G. and X.P. analysed the TEM data.
were measured in a N2-filled glovebox at room temperature using a Keithley 2400
F.L. conducted cyclic voltammetry measurements and analysed the data. F.L. and F.Q.
SourceMeter under simulated sunlight from a solar simulator (EnliTech, SS-F5). A
designed the NMR characterization of the digested MOF sample and analysed the data.
National Renewable Energy Laboratory calibrated silicon solar cell (with a KG-2
S.W., Z. Z. and A.K-Y.J. drafted and finalized the manuscript. All the authors revised the
filter) was used to obtain the AM 1.5G (100 mW cm−2) light intensity from the solar
manuscript.
simulator. The PVSCs were covered with a shading mask with an aperture area of
0.104 cm2 to ensure the accuracy of the current density obtained from the J–V curves.
The J–V measurements were carried out in sweep mode with reverse (from 1.23 V to Competing interests
−0.02 V) and forward (from −0.02 V to 1.23 V) scans with a scan rate of 10 mV s−1. Authors declare no competing interests.
EQE was carried out using an EQE measurement system (EnliTech, QE-R).

Reporting Summary. Further information on research design is available in the Additional information
Nature Research Reporting Summary linked to this article. Supplementary information is available for this paper at https://doi.org/10.1038/
s41565-020-0765-7.

Data availability Correspondence and requests for materials should be addressed to Z.X., Z.Z. or
The data that support the plots within this paper and other findings of this study A.K.-Y.J.
are available from the corresponding author upon reasonable request. Reprints and permissions information is available at www.nature.com/reprints.

Nature Nanotechnology | www.nature.com/naturenanotechnology


nature research | solar cells reporting summary
Corresponding author(s): Zhengtao Xu, Zonglong Zhu and Alex K.-Y. Jen

Solar Cells Reporting Summary


Nature Research wishes to improve the reproducibility of the work that we publish. This form is intended for publication with all accepted papers
reporting the characterization of photovoltaic devices and provides structure for consistency and transparency in reporting. Some list items might
not apply to an individual manuscript, but all fields must be completed for clarity.
For further information on Nature Research policies, including our data availability policy, see Authors & Referees.

` Experimental design
Please check: are the following details reported in the manuscript?
1. Dimensions
Yes The area of the tested solar cells is 0.13 cm2 (see "Methods" section in main text).
Area of the tested solar cells
No
Yes The device area is defined by the overlap between the pre-patterned ITO and back
Method used to determine the device area metal electrode, where the metal electrode is thermally evaporated.
No
2. Current-voltage characterization
Current density-voltage (J-V) plots in both forward Yes Supplementary Figs. 22, 25 and 27.
and backward direction No

Voltage scan conditions Yes The J-V measurements were conducted with sweep mode with reverse (from 1.23 V
For instance: scan direction, speed, dwell times No to -0.02 V) and forward (from -0.02 V to 1.23 V) scan with a scan rate of 10 mV s-1
(see "Methods" section in main text).

Test environment Yes All J-V measurements were conducted in N2-filled glovebox at room temperature
For instance: characterization temperature, in air or in glove box No (see "Methods" section in main text).

Protocol for preconditioning of the device before its Yes No preconditioning protocol was used.
characterization No

Stability of the J-V characteristic Yes The steady-outputs were measured with time evolution of the maximum power point
Verified with time evolution of the maximum power point or with No in Figs. 2E and 3B.
the photocurrent at maximum power point; see ref. 7 for details.

3. Hysteresis or any other unusual behaviour


Description of the unusual behaviour observed during Yes No unusual behaviour was observed during the characterization.
the characterization No
Yes Supplementary Figs. 22, 25 and 27.
Related experimental data
No

4. Efficiency
External quantum efficiency (EQE) or incident Yes Fig. 2D.
photons to current efficiency (IPCE) No

A comparison between the integrated response under Yes The Jsc obtained from J-V measurement is consistent with the Jsc integrated from
the standard reference spectrum and the response No EQE spectrum. (Fig. 2C, 2D).
measure under the simulator

For tandem solar cells, the bias illumination and bias Yes Not relevant.
November 2017

voltage used for each subcell No

5. Calibration
Light source and reference cell or sensor used for the Yes Simulated AM 1.5 G irradiation was from solar simulator (EnliTech, SS-F5, Taiwan). A
characterization No silicon solar cell (with a KG-2 filter) was used as reference cell (see "Methods" section
in main text).

1
Confirmation that the reference cell was calibrated Yes The reference cell was calibrated by National Renewable Energy Laboratory (more
and certified No see "Methods" section in main text).

nature research | solar cells reporting summary


Calculation of spectral mismatch between the Yes The test light source for the device measurement matches well with the one-sun light
reference cell and the devices under test No spectrum and reference cell. The spectral mismatch factor is 1 for the devices

6. Mask/aperture
Yes The size of the mask used for J-V measurements is 0.104 cm2 (See "Methods" section
Size of the mask/aperture used during testing in main text).
No

Variation of the measured short-circuit current Yes All devices has been measured with identical masks
density with the mask/aperture area No

7. Performance certification
Identity of the independent certification laboratory Yes Certified by the National Institute of Metrology (NIM, China).
that confirmed the photovoltaic performance No

A copy of any certificate(s) Yes Certificate is provided in Supplementary Fig. 26.


Provide in Supplementary Information No

8. Statistics
Yes 30 solar cells for each device configurations were tested and statistic for efficiency
Number of solar cells tested distribution demonstration (Fig. 2F and Supplementary Fig. 28).
No
Yes Fig. 2F and Supplementary Fig. 28.
Statistical analysis of the device performance
No

9. Long-term stability analysis


Type of analysis, bias conditions and environmental Yes Long-term stability for the cells was measured and analyzed at different conditions
conditions No including shelf stability in humid air and MPP tracking at elevated temperature in
For instance: illumination type, temperature, atmosphere N2-filled glovebox (detailed conditions are described in main text).
humidity, encapsulation method, preconditioning temperature

November 2017

You might also like