You are on page 1of 6

LETTERS

PUBLISHED ONLINE: 20 JUNE 2016 | DOI: 10.1038/NPHOTON.2016.108

Solution-processed solar cells based on


environmentally friendly AgBiS2 nanocrystals
María Bernechea1†, Nichole Cates1†, Guillem Xercavins1, David So1, Alexandros Stavrinadis1
and Gerasimos Konstantatos1,2*

Solution-processed inorganic solar cells are a promising low- comparable to that of CuInxGa(1−x)Se2 (CIGS) (Supplementary
cost alternative to first-generation solar cells1,2. Solution Fig. 2b), with a favourable bandgap of ∼1.3 eV (Fig. 1e and
processing at low temperatures combined with the use of Supplementary Fig. 2c).
non-toxic and abundant elements can help minimize fabrication The as-synthesized AgBiS2 nanocrystals are capped with oleic
costs and facilitate regulatory acceptance. However, at present, acid, a long aliphatic chain molecule with a carboxylic functional
there is no material that exhibits all these features while demon- group. In view of their insulating character, we sought a ligand
strating promising efficiencies. Many of the candidates being exchange that removes oleate ligands to render the deposited films
explored contain toxic elements such as lead or cadmium (perovs- conductive. Following earlier work on PbS optoelectronics, ethane-
kites2,3, PbS4, CdTe5,6 and CdS(Se)7,8) or scarce elements such as dithiol (EDT) was used to replace the oleate ligands and act as a
tellurium or indium (CdTe and CIGS(Se)/CIS9,10). Others require crosslinking molecule1,4. X-ray photoelectron spectroscopy (XPS)
high-temperature processes such as selenization or sintering, or shows that EDT treatment effectively removes the oleate ligands,
rely on vacuum deposition techniques (Sb2S(Se)311–13, SnS14,15 because the high-energy components in the Ag 3d and Bi 4f
and CZTS(Se)16). Here, we present AgBiS2 nanocrystals as a spectra, which are attributed to the bonding of oleic acid in
non-toxic17, earth-abundant18 material for high-performance, untreated AgBiS2 , disappear after EDT treatment (Fig. 2a,b)26–28.
solution-processed solar cells fabricated under ambient conditions Effective ligand exchange is further supported by the disappearance
at low temperatures (≤100 °C). We demonstrate devices with a of the oleate peaks in Fourier transform infrared spectroscopy
certified power conversion efficiency of 6.3%, with no hysteresis (FTIR) spectra (Supplementary Fig. 3) and the significant reduc-
and a short-circuit current density of ∼22 mA cm−2 for an active tion in the amount of oxygen and carbon in the XPS elemental
layer thickness of only ∼35 nm. analysis (Table 1).
Lead chalcogenide quantum dots4 stand out among the most Solar cells were fabricated with EDT-treated AgBiS2 active layers
promising colloidal nanocrystal materials for solar cells because sandwiched between a ZnO electron-transport layer (ETL) and a
they are solution-processable and exhibit most of the above features, thin (∼10 nm) polymer hole-transport layer (HTL) with an active
although the toxicity of lead remains a concern. AgBiS2 , however, area of 3.1 mm2. This structure, shown schematically in Fig. 3a, is
which belongs to the family of I–III–VI2 compounds, comprises similar to the structure used in Bi2S3 n-type nanocrystal solar
environmentally friendly restriction of hazardous substances cells29. The polymer layer, which does not contribute to carrier gen-
(RoHS)-compliant elements. In its nanocrystalline form, AgBiS2 eration (Supplementary Fig. 2d), is intentionally kept thin to mini-
exhibits photoconductivity and favourable thermoelectric properties mize parasitic absorption. The performance of the EDT-treated
and has been used as a sensitizer or counter-electrode in sensitized devices was poor, with an average power conversion efficiency of
solar cells with modest efficiencies19–24. Here, we present colloidal 1.5% (Fig. 2d). To identify the origin of the low performance, we
AgBiS2 nanocrystals that serve both as the photo-absorbing and focused on the Bi 4f XPS spectra. After EDT treatment the
charge-transporting medium in high-performance solid-state, binding energies of the Bi component (158.1 and 163.4 eV) are
solution-processed solar cells. higher than those corresponding to elemental Bi (∼157 and
We have developed a low-temperature hot-injection synthetic ∼162 eV). However, this component appears at quite low binding
route for the synthesis of colloidal AgBiS2 nanocrystals (see energies relative to standard Bi–S ones26, indicating a high concen-
Methods). The nanocrystals are dispersible in most organic solvents, tration of electrons on the Bi atoms. This low-energy component
and the solutions are stable for months without nanocrystal precipi- was previously observed in the Bi 4f spectra of nanocrystalline
tation or loss in device performance (Fig. 1a and Supplementary Bi2S3 films and has been correlated with a higher trap density and
Fig. 1a). As previously reported19–24, the material crystallizes into increased n-type doping29. Ultraviolet photoelectron spectroscopy
the cubic rock salt structure (Fig. 1b, widening and small deviations (UPS) also reveals n-type doping in EDT-treated AgBiS2 (Fig. 2c).
in peak positions are due to strain25) with nanocrystal diameters of We took the view that thiol chemistry may not adequately
4.6 ± 1 nm (Fig. 1c,d). High-resolution transmission electron passivate the AgBiS2 nanocrystals and therefore explored halide
microscopy (HR-TEM) images show the atomic planes, and fast chemistry as a more effective passivation scheme1,4. Treatment of
Fourier transform (FFT) analysis gives d-space values of 0.207, AgBiS2 nanocrystals with tetramethylammonium iodide (TMAI)
0.316 and 0.287 nm, which correspond to the (220), (111) and resulted in the elimination of the Ag peak associated with
(200) interplanar distances of cubic AgBiS2 (Supplementary Ag-oleate, as was also seen after EDT treatment (Fig. 2a). After
Fig. 2a), in agreement with X-ray diffraction (XRD) data. The material TMAI treatment, the Bi 4f component at low binding energies is
exhibits a very high absorption coefficient (105 to 103 cm−1), present, but is smaller than that observed for the EDT-treated

1
ICFO-Institut de Ciències Fotòniques, The Barcelona Institute of Science and Technology, 08860 Castelldefels (Barcelona), Spain. 2 ICREA–Institució
Catalana de Recerca i Estudis Avançats, Passeig Lluís Companys 23, 08010 Barcelona, Spain. †These authors contributed equally to this work.
* e-mail: gerasimos.konstantatos@icfo.es

NATURE PHOTONICS | VOL 10 | AUGUST 2016 | www.nature.com/naturephotonics 521

© 2016 Macmillan Publishers Limited. All rights reserved


LETTERS NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.108

a b

200
1,400

220
1,200

111
1,000

Counts (a.u.)
800

311
600

222
400

400
200

0
10 15 20 25 30 35 40 45 50 55 60 65 70
2θ´

c d 80 e 106
70
4.62 ± 0.97 nm 105
60
50 104

α (cm−1)
Count

40 103
30
102
20
101
10
50 nm 0 100
0 1 2 3 4 5 6 7 8 9 400 500 600 700 800 900 1,000 1,100
Size (nm) Wavelength (nm)

Figure 1 | Properties of AgBiS2 nanocrystals. a, Photograph of a vial containing a AgBiS2 solution in toluene and a solar cell made with the solution.
b, XRD pattern of AgBiS2 and reference pattern for cubic AgBiS2 (red lines, ref. 00-004-0699), confirming the rock salt structure. c, TEM image of AgBiS2
nanocrystals. Scale bar, 50 nm. d, Size distribution histogram and calculated average size (4.62 ± 0.97 nm). e, Absorption coefficient measurement of
solid-state films of AgBiS2 nanocrystals (dotted lines are included as a guide to the eye).

sample (Fig. 2b). The main component, appearing at a high binding with results for Sb2S3 , which has been reported to strongly interact
energy (∼158.7 and ∼164.1 eV), can be attributed to Bi–S and with the thiophene units in polymer HTLs11. In addition, the exter-
Bi–I bonds (the removal of the original oleic acid was confirmed nal quantum efficiency (EQE) spectrum for the best performing
by FTIR analysis, Supplementary Fig. 3)30. This shift of the Bi HTL shows a high infrared contribution that can only be attributed
peak to higher binding energies indicates the removal of electron to AgBiS2 , pointing to better charge extraction for the thiophene-
density from the Bi atoms, pointing to stronger binding with the rich polymers (Supplementary Fig. 6b).
atomic halide ligands and hence better passivation. TMAI treatment Optimized TMAI-treated AgBiS2 solar cells exhibit the best per-
was also found to lead to a more intrinsic semiconductor, according formance with a PTB7 HTL. A cross-sectional focused ion beam
to UPS measurements (Fig. 2c), resulting in an overall significantly scanning electron microscope (FIB-SEM) image and the approxi-
higher average power conversion efficiency of 4.8% (Fig. 2d). The mate energy levels of such a cell are shown in Fig. 3b,c, respectively.
distribution of efficiencies and the average figures of merit for solar These solar cells exhibit efficiencies up to 5.84% in our laboratory
cells treated with EDT and TMAI are shown in Supplementary (Fig. 3d). Initial stability tests show promise, because these solar
Fig. 5 and Supplementary Table 1, respectively. cells are stable in air for several weeks (Supplementary Fig. 1b). A par-
It is noteworthy that both an ETL and an HTL are required to ticularly striking feature is the high JSC of 18 mA cm−2, despite the
achieve high solar-cell performance. Omitting either one of these small active-layer thickness of ∼35 nm. That such high currents
layers results in significant performance reduction (Supplementary can be achieved with such thin active layers highlights the strong
Table 2). To investigate the role of the HTL on device performance, absorption of AgBiS2 nanocrystals and their promise as a photo-
we fabricated TMAI-treated solar cells with a variety of polymer voltaic material in a variety of device structures including tandem
HTLs (Supplementary Fig. 6). The choice of polymer strongly and sensitized solar cells.
impacts the short-circuit current density (JSC), the fill factor (FF), Figure 3e plots EQE with 1 sun light bias and without light bias.
and consequently the solar-cell efficiency. Curiously, the position Integration of the 1 sun EQE with the solar spectrum gives a pre-
of the polymer highest occupied molecular orbital (HOMO) level dicted JSC of 17.7 mA cm−2, in good agreement with the observed
has little impact on the open-circuit voltage (VOC) of the devices JSC of 18.0 mA cm−2. However, the JSC predicted for the EQE
(Supplementary Table 3), which may indicate that the VOC is without light bias is higher (20.2 mA cm−2). To better understand
limited by trap-assisted recombination. No correlation was observed this, in Fig. 3f we show how the AgBiS2 layer thickness affects the
between the solar-cell figures of merit and the polymer mobility or observed JSC , JSC from EQE without light bias and JSC calculated
HOMO level. However, we have observed that thiophene-rich poly- with transfer-matrix-model (TMM) simulations assuming 100%
mers outperform those with few or no thiophene units, indicating internal quantum efficiency. The TMM-simulated JSC exhibits
that thiophene rings facilitate efficient charge transfer from the three regions (labelled in Fig. 3f ): (1) an increase with increasing
AgBiS2 and thus improve charge collection. This is in agreement thickness below ∼40 nm, (2) a dip around 100 nm due to optical

522 NATURE PHOTONICS | VOL 10 | AUGUST 2016 | www.nature.com/naturephotonics

© 2016 Macmillan Publishers Limited. All rights reserved


NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.108 LETTERS
a b 158.2
Untreated 163.6 Untreated
374.5 368.5 368.2 164.4 159.0
374.2

Intensity (a.u.)
158.1
Intensity (a.u.)

367.8 163.4
373.8 EDT EDT

368.0 TMAI 163.5 158.2 TMAI


374.0 164.1 158.7

376 374 372 370 368 366 166 164 162 160 158 156 154
Binding energy (eV) Binding energy (eV)

c −3.5 d 20
EDT TMAI

Current density (mA cm−2)


−4.0 10
Energy (eV)

0
−4.5

−10
TMAI: mean η = 4.8%
−5.0 EDT: mean η = 1.5%
−20
−0.6 −0.4 −0.2 0.0
−5.5 Voltage (V)

Figure 2 | Effect of ligand treatment on AgBiS2 properties and solar-cell performance. a,b, XPS core-level spectra of Ag 3d (a) and Bi 4f and S 2p (b) for
the untreated and ligand-exchanged samples. Red lines represent the fitted components of chemical states of the metals. Blue lines represent the fitted
component of sulfur. c, Energy levels as determined by UPS for AgBiS2 after ligand exchange with EDT and TMAI. Measurement details are given in
Supplementary Fig. 4. d, J–V curves for solar cells treated with TMAI and EDT, as well as the mean efficiency η obtained for each. A histogram showing the
distribution of efficiencies for EDT- and TMAI-treated solar cells and the average figures of merit are shown in Supplementary Fig. 5 and Supplementary
Table 1, respectively.

interference effects, and (3) a second increase at higher thicknesses. process or the use of alternative surface passivation schemes
The JSC from EQE without light bias shows this same trend and should further suppress trap-assisted recombination and increase
exhibits current densities that approach the simulated values. solar-cell performance. We also performed transient photovoltage
Because the TMM simulation assumes the complete collection of (TPV) measurements with light biases ranging from ∼0.2 sun to
all excited charges, these results indicate that charge collection is effi- 1 sun, and fitted the photovoltage decays to determine the carrier
cient at low light intensities. On the other hand, the observed JSC lifetime τ as a function of VOC (Fig. 4c). As expected, the
values, which are from measurements at 1 sun, drop off for high carrier lifetime decreases with increasing light intensity (and
thicknesses, demonstrating that current collection is less efficient at increasing VOC). The carrier lifetime at 1 sun is ∼2 µs, on a par
higher light intensities in devices thicker than 50 nm. These results with that of PbS quantum dot solar cells33. We hypothesize, there-
suggest that an intensity-dependent process is responsible for fore, that current losses in thicker AgBiS2 devices are due to poor
current loss, particularly in solar cells with thick AgBiS2 layers. carrier transport. Further improvements to the synthesis and
To investigate this further, we also measured JSC and VOC as a ligand exchange should be investigated to increase and balance
function of light intensity. The dependence of JSC on light intensity charge transport.
in Fig. 4a is fitted with a power-law expression (JSC ∝ intensityα), To corroborate our findings we sent devices to Newport
where α = 0.88. This nonlinearity (α ≠ 1) is caused by incomplete Corporation for independent certification. In Fig. 5, we present the
extraction of charges before recombination31. Figure 4b shows Newport-certified results of a record-performing device with an effi-
VOC as a function of light intensity. The fit gives an ideality ciency of 6.3% and JSC of 22 mA cm−2 (Fig. 5 and Supplementary
factor of 1.3, which indicates Shockley–Read–Hall (SRH) trap- Fig. 7). AgBiS2 solar cells do not exhibit hysteresis when the voltage
assisted recombination32. Improvements to the ligand-exchange is scanned in the forward and reverse directions (Fig. 5a). The
light-biased EQE spectrum shows a strong photocurrent response
in the favourable region of 300–1,100 nm (Fig. 5b). This certified
Table 1 | Relative amounts of each element found by XPS. result further demonstrates the promise of AgBiS2 as a highly
Sample Ag C O S (2s) I absorbing, solution-processed material for photovoltaic applications.
Untreated 1.5 12.5 1.9 1.6 – In conclusion, we have presented a new class of ultrathin-film
EDT 1.3 2.5 0.3 2.1 – solution-processed colloidal nanocrystal solar cells based on
TMAI 1.4 3.0 0.9 1.6 0.8 AgBiS2 nanocrystals with power conversion efficiencies up to
6.3%. To our knowledge, this is the first efficient inorganic solar
Quantities normalized to Bi.
material that simultaneously meets the demands for non-toxicity,

NATURE PHOTONICS | VOL 10 | AUGUST 2016 | www.nature.com/naturephotonics 523

© 2016 Macmillan Publishers Limited. All rights reserved


LETTERS NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.108

a b c −3.0
MoO3 (~2 nm)/Ag (~120 nm) PTB7
−3.5
AgBiS2
Ag 120 nm
Polymer (~10 nm) −4.0 ZnO

Energy (eV)
PTB7 6 nm
AgBiS2 (~35 nm) AgBiS2 37 nm −4.5 Ag
ZnO 44 nm ITO
ZnO (~45 nm) −5.0
MoO3
ITO 106 nm
−5.5
ITO (~100 nm) 100 nm
−6.0

d 20 e 80 f 30 1 2 3
Current density (mA cm−2)

Light 60
10
Dark

JSC (mA cm−2)


EQE (%) 20

0 40
JSC = 18.0 mA cm−2
10
−10 VOC = 0.50 V 20 Observed (1 sun)
FF = 0.65 Dark: 20.2 mA cm−2 From TMM model (1 sun)
η = 5.84% From EQE (no light bias)
1 sun: 17.7 mA cm−2
−20 0 0
−0.6 −0.4 −0.2 0.0 400 600 800 1,000 0 50 100 150 200
Voltage (V) Wavelength Thickness (nm)

Figure 3 | Solar cell characterization. a, Structure of AgBiS2 solar cells. b, FIB-SEM image of device cross-section. c, Approximate energy levels of solar-cell
materials. d, J–V curves of best solar-cell device. e, EQE without light bias and with 1 sun light bias. f, Observed JSC values from J–V measurements at 1 sun
(filled squares), JSC values calculated from TMM simulations (line) and JSC values predicted from EQE without light bias (open circles), as a function of
AgBiS2 thickness. ITO, indium tin oxide.

a 25 b 0.50 c
10
0.48
20
Carrier lifetime (μs)
JSC (mA cm−2)

α = 0.88 0.46 n = 1.3


15
VOC (V)

0.44
10
0.42

5
0.40
1
0 0.38
0.0 0.5 1.0 0.1 1 0.39 0.42 0.45 0.48
Intensity (sun) Intensity (sun) VOC (V)

Figure 4 | Optoelectronic characterization of TMAI-treated AgBiS2 solar cells. a, JSC as a function of light intensity and the power-law fit (JSC ∝ intensityα).
b, VOC as a function light intensity and the logarithmic fit used to determine the ideality factor n. c, Carrier lifetimes extracted from TPV measurements as a
function of VOC (the line represents the exponential fit).

a b 80

20
Forward
Current density (mA cm−2)

60
Reverse
10
EQE (%)

40
0
JSC = 22.1 mA cm−2
VOC = 0.45 V 20
−10
FF = 0.63
η = 6.31%

−20 0
−0.6 −0.4 −0.2 0.0 300 600 900 1,200
Voltage (V) Wavelength (nm)

Figure 5 | Newport certification of AgBiS2 solar cells. a,b, J–V (a) and light-biased EQE (b) curves. See Supplementary Fig. 7 for the accreditation certificate.

524 NATURE PHOTONICS | VOL 10 | AUGUST 2016 | www.nature.com/naturephotonics

© 2016 Macmillan Publishers Limited. All rights reserved


NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.108 LETTERS
abundance and low-temperature solution processing. AgBiS2 nano- 22. Pejova, B., Nesheva, D., Aneva, Z. & Petrova, A. Photoconductivity and
crystals also have significant potential as ultrathin absorbing layers relaxation dynamics in sonochemically synthesized assemblies of AgBiS2
quantum dots. J. Phys. Chem. C 115, 37–46 (2011).
in other device architectures such as sensitized and tandem solar 23. Guin, S. N. & Biswas, K. Cation disorder and bond anharmonicity optimize the
cells due to the high JSC of our devices despite their very thin thermoelectric properties in kinetically stabilized rocksalt AgBiS2 nanocrystals.
active layers (∼35 nm). Furthermore, better nanocrystal passivation, Chem. Mater. 25, 3225–3231 (2013).
light-trapping schemes and nanostructuring34,35 of the active layer 24. Chen, C., Qiu, X., Ji, S., Jia, C. & Ye, C. The synthesis of monodispersed
should help overcome the limitations of these devices and result AgBiS2 quantum dots with a giant dielectric constant. CrystEngComm 15,
7644–7648 (2013).
in efficiencies beyond the already compelling ones reported herein. 25. Qin, W., Nagase, T., Umakoshi, Y. & Szpunar, J. A. Relationship between
microstrain and lattice parameter change in nanocrystalline materials. Philos.
Methods Mag. Lett. 88, 169–179 (2008).
Methods and any associated references are available in the online 26. Malakooti, R. et al. Shape-controlled Bi2S3 nanocrystals and their plasma
version of the paper. polymerization into flexible films. Adv. Mater. 18, 2189–2194 (2006).
27. Krsmanovi, R. et al. Colloids and surfaces B: biointerfaces adsorption of sulfur
Received 18 December 2015; accepted 28 April 2016; onto a surface of silver nanoparticles stabilized with sago starch biopolymer.
published online 20 June 2016 Colloids Surf. B 73, 30–35 (2009).
28. Dong, T. et al. One-step synthesis of uniform silver nanoparticles capped by
saturated decanoate : direct spray printing ink to form metallic silver films. Phys.
References Chem. Chem. Phys. 11, 6269–6275 (2009).
1. Kramer, I. J. & Sargent, E. H. The architecture of colloidal quantum dot solar
29. Bernechea, M., Cao, Y. & Konstantatos, G. Size and bandgap tunability in Bi2S3
cells: materials to devices. Chem. Rev. 114, 863–882 (2014).
colloidal nanocrystals and its effect in solution processed solar cells. J. Mater.
2. Kazim, S., Nazeeruddin, M. K., Grätzel, M. & Ahmad, S. Perovskite as light
Chem. A 3, 20642–20648 (2015).
harvester: a game changer in photovoltaics. Angew. Chem. Int. Ed. 53,
30. Boopathi, K. M. et al. Solution-processable bismuth iodide nanosheets as
2812–2824 (2014).
hole transport layers for organic solar cells. Sol. Energy Mater. Sol. Cells 121,
3. Zhou, H. et al. Interface engineering of highly efficient perovskite solar cells.
35–41 (2014).
Science 345, 542–546 (2014).
31. Cowan, S. R., Roy, A. & Heeger, A. J. Recombination in polymer–fullerene bulk
4. Chuang, C. M., Brown, P. R., Bulović, V. & Bawendi, M. G. Improved
heterojunction solar cells. Phys. Rev. B 82, 245207 (2010).
performance and stability in quantum dot solar cells through band alignment
32. Kirchartz, T., Deledalle, F., Tuladhar, P. S., Durrant, J. R. & Nelson, J. On the
engineering. Nature Mater. 13, 796–801 (2014).
differences between dark and light ideality factor in polymer:fullerene solar cells.
5. Major, J. D., Treharne, R. E., Phillips, L. J. & Durose, K. A low-cost non-toxic
J. Phys. Chem. Lett. 4, 2371–2376 (2013).
post-growth activation step for CdTe solar cells. Nature 511, 334–337 (2014).
33. Chuang, C.-H. M. et al. Open-circuit voltage deficit, radiative sub-bandgap
6. Panthani, M. G. et al. High efficiency solution processed sintered CdTe
states, and prospects in quantum dot solar cells. Nano Lett. 15,
nanocrystal solar cells: the role of interfaces. Nano Lett. 14, 670–675 (2014).
3286–3294 (2015).
7. Santra, P. K. & Kamat, P. V. Mn-doped quantum dot sensitized solar cells: a
34. Rath, A. K., Bernechea, M., Martinez, L. & Konstantatos, G. Solution-processed
strategy to boost efficiency over 5%. J. Am. Chem. Soc. 134, 2508–2511 (2012).
heterojunction solar cells based on p-type PbS quantum dots and n-type Bi2S3
8. Pan, Z. et al. Near infrared absorption of CdSexTe1–x alloyed quantum dot
nanocrystals. Adv. Mater. 23, 3712–3717 (2011).
sensitized solar cells with more than 6% efficiency and high stability. ACS Nano
35. Rath, A. K. et al. Solution-processed inorganic bulk nano-heterojunctions and
7, 5215–5222 (2013).
their application to solar cells. Nature Photon. 6, 529–534 (2012).
9. Reinhard, P. et al. Review of progress toward 20% efficiency flexible CIGS
solar cells and manufacturing issues of solar modules. IEEE J. Photovolt. 3,
572–580 (2013). Acknowledgements
10. Romanyuk, Y. E. et al. All solution-processed chalcogenide solar cells—from The authors thank J. Osmond for ellipsometry measurements and H. Maeckel for
single functional layers towards a 13.8% efficient CIGS device. Adv. Funct. developing the TMM simulation code and device-characterization set-ups. The research
Mater. 25, 12–27 (2015). leading to these results has received funding from Fundació Privada Cellex and the
11. Im, S. H. et al. Toward interaction of sensitizer and functional moieties in hole- European Community’s Seventh Framework Programme (FP7-ENERGY.2012.10.2.1)
transporting materials for efficient semiconductor-sensitized solar cells. Nano under grant agreement 308997. The authors acknowledge financial support from the
Lett. 11, 4789–4793 (2011). Spanish Ministry of Economy and Competitiveness (MINECO) and the ‘Fondo Europeo de
12. Chang, J. A. et al. Panchromatic photon-harvesting by hole-conducting Desarrollo Regional’ (FEDER) through grant MAT2014-56210-R. This work was also
materials in inorganic–organic heterojunction sensitized-solar cell through supported by AGAUR under the SGR grant (2014SGR1548). N.C. acknowledges
the formation of nanostructured electron channels. Nano Lett. 12, support from Marie Curie Actions FP7-PEOPLE-2013-IIF (project no. 622358).
1863–1867 (2012). G.K. acknowledges financial support from the Spanish Ministry of Economy and
13. Zhou, Y. et al. Thin-film Sb2Se3 photovoltaics with oriented one-dimensional Competitiveness, through the ‘Severo Ochoa’ Programme for Centres of Excellence in
ribbons and benign grain boundaries. Nature Photon. 9, 409–415 (2015). R&D (SEV-2015-0522).
14. Steinmann, V. et al. 3.88% efficient tin sulfide solar cells using congruent thermal
evaporation. Adv. Mater. 26, 7488–7492 (2014). Author contributions
15. Sinsermsuksakul, P. et al. Overcoming efficiency limitations of SnS-based solar M.B. conceived, synthesized and characterized the material, designed experiments and co-
cells. Adv. Energy Mater. 4, 1400496 (2014). wrote the manuscript. N.C. fabricated and characterized the devices, ran and analysed
16. Kim, J. et al. High efficiency Cu2ZnSn(S,Se)4 solar cells by applying a double TMM simulations, designed experiments and co-wrote the manuscript. G.X. and D.S.
In2S3/CdS emitter. Adv. Mater. 26, 7427–7431 (2014). fabricated and characterized the devices. A.S. took the FIB-SEM image. G.K. designed
17. Mohan, R. Green bismuth. Nature Chem. 2, 336 (2010). experiments, supervised the work, directed the study and co-wrote the manuscript. All
18. Vesborg, P. C. K. & Jaramillo, T. F. Addressing the terawatt challenge: scalability authors discussed the results, and have read and agreed to the publication of
in the supply of chemical elements for renewable energy. RSC Adv. 2, this manuscript.
7933–7947 (2012).
19. Liang, N. et al. Homogenously hexagonal prismatic AgBiS2 nanocrystals: Additional information
controlled synthesis and application in quantum dot-sensitized solar cells. Supplementary information is available in the online version of the paper. Reprints and
CrystEngComm 17, 1902–1905 (2015). permissions information is available online at www.nature.com/reprints. Correspondence and
20. Huang, P., Yang, W. & Lee, M. AgBiS2 semiconductor-sensitized solar cells. requests for materials should be addressed to G.K.
J. Phys. Chem. C 117, 18308–18314 (2013).
21. Pejova, B., Grozdanov, I., Nesheva, D. & Petrova, A. Size-dependent properties Competing financial interests
of sonochemically synthesized three-dimensional arrays of close-packed G.K., N.C. and M.B. have filed a provisional patent application with reference number
semiconducting AgBiS2 quantum dots. Chem. Mater. 20, 2551–2565 (2008). EP15176300 on AgBiS2 nanocrystal-based solar cells.

NATURE PHOTONICS | VOL 10 | AUGUST 2016 | www.nature.com/naturephotonics 525

© 2016 Macmillan Publishers Limited. All rights reserved


LETTERS NATURE PHOTONICS DOI: 10.1038/NPHOTON.2016.108

Methods Solar cell fabrication. All solar cell fabrication steps were performed in air. ITO-
Chemicals and materials. All the reactions were carried out using standard Schlenk covered glass substrates (Universität Stuttgart, Institut für Großflächige
techniques. Reagents were purchased from Sigma Aldrich, except Bi(OAc)3 , which Mikroelektronik) were cleaned by ultrasonication in soapy water, acetone and
was purchased from Alfa Aesar. Poly[(4,8-bis-(2-ethylhexyloxy)-benzo(1,2-b:4,5-b′)- isopropanol for 10 min each and dried with nitrogen. A ∼45 nm ZnO layer was then
dithiophene)-2,6-diyl-alt-(4-(2-ethylhexyl)-3-fluorothieno[3,4-b]thiophene-)-2- grown using a sol–gel method. Zinc acetate dehydrate (1 g) was dissolved in 10 ml
carboxylate-2-6-diyl)] (PBE) and poly[4,8-bis(5-(2-ethylhexyl)thiophen-2-yl)benzo- methoxyethanol and 284 µl ethanolamine. The solution was spin-cast onto the ITO-
[1,2-b;4,5-b′]dithiophene-2,6-diyl-alt-(4-(2-ethylhexyl)-3-fluorothieno[3,4-b]- covered glass substrates at 3,000 r.p.m. and heated at 200 °C for 30 min. This process
thiophene-)-2-carboxylate-2-6-diyl)] (PCE-10) were purchased from 1-material, was then repeated to ensure uniform coverage of ZnO on the ITO. Afterwards, a
poly(3-hexylthiophene-2,5-diyl) (P3HT) from Rieke Metals, poly(3,4- ∼35 nm film of AgBiS2 nanocrystals was deposited using three rounds of a layer-by-
ethylenedioxythiophene) polystyrene sulfonate (PEDOT:PSS) from Heraeus Clevios layer (LBL) process using a 20 mg ml–1 solution of AgBiS2 in toluene filtered
and 2,2′,7,7′-tetrakis-(N,N-di-4-methoxyphenylamino)-9,9′-spirobifluorene through a 0.45 µm polytetrafluoroethylene (PTFE) filter. One LBL cycle involved
(spiro-OMeTAD) from Merck. dropping one drop of the AgBiS2 solution onto the ZnO-covered substrate spinning
at 2,000 r.p.m. and waiting 10 s, adding five drops of TMAI (1 mg ml–1 in methanol)
Synthesis of AgBiS2 nanocrystals. For the synthesis of AgBiS2 nanocrystals, or EDT (2% in acetonitrile) and waiting 20 s. For TMAI, this step was repeated once
1 mmol Bi(OAc)3 , 0.8 mmol Ag(OAc) and 17 mmol oleic acid (OA) were pumped with an additional five drops of the TMAI solution and an additional 20 s of waiting.
overnight at 100 °C to form the Bi and Ag oleates and remove oxygen and moisture. The TMAI- and EDT-treated films were then rinsed with methanol or acetonitrile,
The reaction atmosphere was then switched to Ar, and 1 mmol respectively, and then rinsed with toluene. For thinner and thicker AgBiS2 layers,
hexamethyldisilathiane (HMS) dissolved in 5 ml 1-octadecene (ODE) was quickly more or fewer cycles of the LBL process were performed. After AgBiS2 deposition,
injected into the flask, the heating was stopped (without removing the heating the TMAI- and EDT-treated samples were annealed in air for 10 min at 100 and
mantel), and the reaction was allowed to cool slowly. Nanocrystals were isolated after 50 °C, respectively. The samples were then stored in air in the dark overnight before
the addition of acetone and centrifugation, purified by successive dispersion in depositing a thin layer of PTB7 by spin-coating a 5 mg ml–1 solution of PTB7 in
toluene and precipitation with acetone, and finally dispersed in anhydrous toluene. dichlorobenzene at 2,000 r.p.m. Finally, a Kurt J. Lesker Nano36 system was used to
The reaction flasks were protected from light until precipitation. evaporate 2 nm of MoO3 and 120 nm of Ag through a shadow mask to produce solar
cells with a diameter of 2 mm (area of 3.1 mm2). The devices were kept in air in the
Characterization of AgBiS2 nanocrystals. UV–vis absorption measurements were dark for at least one day before measuring.
performed using a Cary 5000 UV–Vis–NIR spectrophotometer in solution using a
glass cuvette with a 1 mm optical path or as films on glass substrates. For the Solar cell characterization. All device characterization was performed in air under
absorption coefficient measurements, an integrating sphere set-up was used. ambient conditions. Current–voltage measurements were performed with a Keithley
Films exchanged with TMAI were grown on glass, the thickness was measured 2400 source meter and a Newport Oriel Sol3A solar simulator with an AM1.5 filter.
using a profilometer, and the absorption coefficient was determined as The intensity of the solar simulator was adjusted using a Hamamatsu S1336 silicon
described elsewhere36. photodiode that had been calibrated at the Fraunhofer Institute of Solar Energy
TEM and XRD measurements were performed in the Scientific and Systems, Freiburg, Germany. The solar cells were measured with and without masks,
Technological Centres of the University of Barcelona (CCiT-UB). TEM micrographs and no difference in power conversion efficiency was observed. For certified cells,
were obtained using a JEOL 2100 microscope operating at an accelerating voltage of appropriate masks have always been used.
200 kV. Samples were prepared by placing two drops of a diluted toluene solution on The EQE was measured using a Newport Cornerstone 260 monochromator, a
a holey carbon-coated grid and allowing the solvent to evaporate in air. The average Thorlabs MC2000 chopper, a Stanford Research SR570 trans-impedance amplifier
diameter was calculated by measuring the diameters of no less than 100 nanocrystals and a Stanford Research SR830 lock-in amplifier. A calibrated Newport 818-UV
from non-aggregated areas. XRD data were collected using a PANalytical X’Pert photodetector was used as a reference. Light bias was provided by a Newport Xe
PRO MPD Alpha1 powder diffractometer in Bragg–Brentano θ/2θ geometry with a arc lamp.
radius of 240 mm, Cu Kα radiation (λ = 1.5406 Å) and a power of Measurements of JSC and VOC as a function of light intensity were performed
45 kV–40 mA. with an Agilent 4000X oscilloscope with the 50 Ω and 1 MΩ input terminals,
UPS and XPS measurements were performed at the Institut Català de respectively. A FiberTech Optica LED lightsource provided the bias light.
Nanociencia i Nanotecnologia (ICN2). XPS measurements were performed TPV measurements were performed with an Agilent 4000X oscilloscope using a
with a Phoibos 150 analyser (SPECS) in ultrahigh-vacuum conditions (base pressure 1 MΩ input terminal and a Vortran Stradus laser with a wavelength of 637 nm.
of 1 × 10–10 mbar) with a monochromatic Kα X-ray source (1,486.74 eV). The A FiberTech Optica LED light source provided the bias light. Carrier lifetime τ
pass energy value used was 10 eV for the high-resolution spectrum, and the and VOC were determined by fitting the exponential photovoltage decays with
energy resolution as measured by the full-width at half-maximum (FWHM) of the V = Aexp(–t/τ) + VOC , where V is the photovoltage, A is an exponential prefactor
Ag 3d5/2 peak for a sputtered Ag foil was 0.55 eV. Intensities were estimated by and t is time.
calculating the integral of each peak, determined by subtracting the Shirley-type The cross-section of the device was imaged using an in-lens secondary electron
background and fitting the experimental curve to a combination of Lorentzian and detector of a Zeiss Auriga FIB-SEM microscope operated at 5 kV. The cross-section
Gaussian lines of variable proportions. Accurate binding energies (0.2 eV) were was prepared in situ by depositing a thin (a few nanometres) layer of Pt on top of the
determined by referencing to the C 1s peak at 284.8 eV. UPS measurements device via a gas injection system integrated in the microscope and subsequently
were performed on a SPECS PHOIBOS 150 electron spectrometer using using the FIB mode of the microscope to mill down a 6 µm × 2 µm rectangular
monochromated HeI radiation (21.2 eV). Samples were prepared by covering ITO region into the top surface.
substrates (∼10 × 10 mm2) with toluene solutions and allowing the solvent to
evaporate in air, or following a layer-by-layer process, as described in the
next section. References
FTIR measurements were performed on a Cary 600 FTIR spectrophotometer in 36. Cesaria, M., Caricato, A. P. & Martino, M. Realistic absorption coefficient of
transmission mode using Nujol mulls between polyethylene sheets. ultrathin films. J. Opt. 14, 105701 (2012).

NATURE PHOTONICS | www.nature.com/naturephotonics

© 2016 Macmillan Publishers Limited. All rights reserved

You might also like