You are on page 1of 12

Solar Energy 188 (2019) 1209–1220

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Phase transition behavior and defect analysis of CuSbS2 thin films for T
photovoltaic application prepared by hybrid inks

Shahara Banua,b, Yunae Choa, Kihwan Kima, Seung Kyu Ahna, Jihye Gwaka, Ara Choa,b,
a
New and Renewable Energy Research Division, Photovoltaic Laboratory, Korea Institute of Energy Research (KIER), 152 Gajeong-ro, Yuseong-gu, Daejeon 305-343,
South Korea
b
University of Science and Technology (UST), 217 Gajeong-ro, Yuseong-gu, Daejeon, South Korea

A R T I C LE I N FO A B S T R A C T

Keywords: CuSbS2 (CAS) thin films were deposited via a non-vacuum hybrid ink method onto a Mo/soda lime glass (SLG)
CuSbS2 solar cell substrate. To fabricate the CAS films, Cu-Sb precursors were spin coated and then annealed with sulfur powder.
Hybrid ink During annealing, different amounts of sulfur powder were used to control the crystal orientation of the films. X-
S-flux ray diffraction (XRD) patterns were employed to examine the crystal orientation by calculating the texture co-
Phase transition behavior
efficient and Lotgering factor. It was found that the S-flux amount during sulfurization was a critical parameter
Recombination
Defects
for controlling the crystal orientation and phase transition of the CAS structure. Originally, CAS has an or-
thorhombic structure; however, if the S-flux was higher than the optimum, a pseudo-phase transition from
orthorhombic to cubic was observed. In addition, the electrical characteristics and defect properties were con-
ducted for the solar cells prepared with various S-flux to understand the difference in the photovoltaic perfor-
mances affected by the structural change. Admittance spectroscopy revealed that the defect levels were shal-
lower in the CAS solar cell with an orthorhombic structure, which could have contributed to the better
photovoltaic performance than that of the cubic structure. The CAS solar cell deposited with low S-flux exhibited
dominant VS2+ defects; however, for excessive S-flux, CuSb2− became prominent.

1. Introduction concentrations in the range of 1016–1018 cm−3 (Banu et al., 2016;


Willian de Souza Lucas et al., 2017; Yang et al., 2014), and a hole
Over the decades, earth-abundant, cost-efficient, and en- mobility of 0.1–1.0 cm2 V−1 s−1 (Welch et al., 2015a,b). It also shows
vironmentally friendly constituent absorber materials have attracted p-type conductivity owing to the dominant copper vacancies (VCu)
significant interest as substitutes of commercialized chalcogenide based (Willian de Souza Lucas et al., 2017; Yang et al., 2014), low tempera-
Cu(InGa)Se2 (CIGS) and CdTe thin-film solar cells. For the design of ture grain growth (Yang et al., 2014), and high spectroscopic-limited
earth-abundant photovoltaics (PV) with environmentally friendly and maximum efficiency (SLME) of 22.9% (Yu et al., 2013).
inexpensive constituents, a wide variety of research has been carried From the inspiration of its favorable fundamental optoelectronic
out on absorber materials, such as CuSbS2 (Cho et al., 2017), Fe2S properties and promising applications, the CAS-based solar cell was
(Berry et al., 2012; Moon et al., 2014), Sb2S3 (Chen et al., 2017), SnS constructed first in 2005 (Rodriguez-Lazcano et al., 2005). Since then,
(Banu et al., 2017; Sinsermsuksakul et al., 2014), and Cu2SnS3 (Tiwari CAS solar cells have been demonstrated successfully via various ap-
et al., 2017), that contain earth-abundant and low-toxicity constituent proaches. However, so far it has been struggling with low device effi-
materials with simpler structures. Among these, CuSbS2 (CAS) has at- ciency mostly less than or near 1% (Macías et al., 2017; Wan et al.,
tracted attention as a promising solar cell absorber material because it 2016; Willian de Souza Lucas et al., 2016; Yang et al., 2014). Minimal
is environmentally friendly, cost efficient, has earth abundant materials studies have reported greater than 2% efficiency, such as 2.55%, with a
and suitable optoelectronic properties for thin-film solar cells. It fea- device made by magnetron sputtering followed by a sulfurization pro-
tures an appropriate band gap of 1.38–1.65 eV (Banu et al., 2016; cess (Zhang et al., 2018). In addition, a 3.13% device was prepared with
Welch et al., 2015a,b; Yang et al., 2014) for photovoltaic solar energy electrochemical deposition of metal stacks followed by subsequent
conversion, a high absorption coefficient larger than 105 cm−1 in the sulfurization (Septina et al., 2014). In previous study, we reported
visible region (Welch et al., 2015a,b; Yu et al., 2013), tunable hole 3.22% conversion efficiency for a CAS device using a hybrid ink method


Corresponding author at: Photovoltaic Laboratory, Korea Institute of Energy Research (KIER), 152 Gajeong-ro, Yuseong-gu, Daejeon 305-343, South Korea.
E-mail address: icemua@kier.re.kr (A. Cho).

https://doi.org/10.1016/j.solener.2019.07.019
Received 22 April 2019; Received in revised form 10 June 2019; Accepted 6 July 2019
0038-092X/ © 2019 International Solar Energy Society. Published by Elsevier Ltd. All rights reserved.
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

in 2016, which is the highest reported efficiency until now using this and sheet resistance of 0.18 Ω/□.
absorber (Banu et al., 2016). However, the efficiency of CAS solar cell is
still low after over a decade of research. Therefore, further optimization 2.2. Preparation of the hybrid inks
of device fabrication and characterization are required for the im-
provement of device performance. The hybrid ink was prepared using Cu(I)Ac, Sb(Ac)3, MeOH, and
To understand the behavior of this absorber material extensively, MEA. In a glove box, Cu(I)Ac and Sb(Ac)3 each were mixed with MEA
we systematically investigated the influence of Cu/Sb ratio to fabricate and MeOH in 10-mL vials separately and were sealed with the lids.
phase pure CAS solar cell and its subsequent effects on photovoltaic Then, both mixtures were taken out from the glove box and sonicated
performance in another study (Banu et al., 2019). For the CIGS solar separately in an ultrasonicator (Branson 2510) for approximately
cells, Cu/(Ga + In) (CGI) is one of the most critical and important 30 min at room temperature to obtain complete dissolution. The Sb and
fabrication condition for making the absorber layer as a p-type because Cu precursor solution colors were clear and greenish-blue, respectively.
the free carriers in CIGS-based materials are generated from the Cu After the dissolution, both vials were returned to the glove box and
vacancies (VCu) (Lee et al., 2016). Similarly, the p-type conductivity of blended together. The molar ratio of Cu(I)Ac to Sb(Ac)3 was fixed at
the CAS solar cell is responsible for the free carriers induced by VCu 1:1.2, which optimized in our previous study to make phase pure
(Yang et al., 2014). Conversely, it is very delicate to make pure CAS CuSbS2 (Banu et al., 2019). The molar ratio of both Cu(I)Ac and Sb(Ac)3
phase owing to the presence of other stable binary and ternary phases with MEA was 1:6 to achieve adequate coordination between the MEA
of Cu-Sb-S compounds, such as Sb2S3, Cu3SbS3, Cu3SbS4, and and precursors (Cu and Sb). The weight ratio of Cu(I)Ac and Sb(Ac)3
Cu12Sb4S13 (Banu et al., 2019; Cho et al., 2017; Yu et al., 2013). with MeOH was 1:5 to adjust the viscosity of the precursor ink.
Nevertheless, CuSbS2 single phase was fabricated successfully by con-
trolling the Cu/Sb ratio without the formation of other unexpected 2.3. Thin film deposition
phases (Banu et al., 2019).
Traditionally, characterization or analysis can help to understand The precursor inks were coated onto Mo-coated (1 μm) soda lime
the limitation of solar cell performance and provide further information glass substrates by spin coating (ACE-200, Dong Ah Tech., Korea) at
for methods to improve the efficiency of the cell. Hence, the Cu/Sb ratio 2000 rpm for 20 s. Then, the samples were dried immediately on a hot
was first optimized to create an appropriate amount of VCu, and then, plate at 80 °C, 120 °C, and then 200 °C for 5 min each in air to evaporate
the recombination mechanism and defect properties were investigated the solvent and induce thermal decomposition of MEA. The spin coating
for the optimum Cu/Sb-ratio device. From these results, deep defects in and drying processes were repeated four more times to obtain the re-
the absorber at 0.23 eV above the valance band maximum were ob- quired thickness. The film deposition procedures were performed in
served, which were responsible for the CuSb antisite defects. It was ambient air.
suggested that the efficiency was still low due to the CuSb defects even
though VCu was optimized. Therefore, for further improvement of the 2.4. Sulfurization
cell performance, the CuSb defects need to be scaled down.
To reduce the CuSb defects, different other conditions were tried. The precursor films were subsequently sulfurized in a rapid thermal
Based on the reports of the CZTS solar cell, self p-type doping in CZTS is annealing (RTA) system in which a graphite container was inserted
related to VCu with CuZn antisite defects (Banu et al., 2019), and the inside. The sulfurization chamber was equipped with a temperature,
energy level of CuZn can be changed by the S concentration owing to the time, and pressure controller. The precursor samples were placed inside
strong hybridization between S 3p and Cu 3d (Duan et al., 2013). the graphite container along with the sulfur powder, which was then
Therefore, it is predicted that the deep defects level could be controlled placed inside the RTA heat treatment chamber. The setup of the RTA
by changing the S-flux during the preparation of the CAS solar cells. sulfurization equipment is shown in reference Cho et al. (2013a). In this
Thus, studying the role of sulfur flux on the CAS solar cells could be study, the S amount (0.1–0.5 g) was chosen as main experimental
fruitful to fabricate the device with fewer defects. variables to control the S-flux inside the graphite container. For all
In this regard, we report a comparative study of CAS solar cells cases, sulfurization was performed at 450 °C for 30 min, which has
prepared with various S-flux using hybrid ink with an optimized Cu/Sb previously reported optimum growth conditions (Banu et al., 2016).
ratio. However, while we were trying to change S-flux, surprisingly a After loading the samples, the chamber was initially evacuated to a low
phase transition behavior was observed in the CAS solar cells. This vacuum pressure of 10−3 Torr using a rotary pump. Then, the back-
pseudo-phase transition could be related to the preferred orientation of ground pressure was regulated to 1.02 atm (1 atm + 1/3 psi) by in-
CIGS and will be discussed in detail. In addition, the cell characteristics jecting nitrogen gas, and the time used to increase the temperature
deduced from the current density-voltage and external quantum effi- between the set points was fixed to 5 min.
ciency measurements and electrical defect properties from the capaci-
tance based measurements, including capacitance-voltage (C-V), ad- 2.5. Device fabrication
mittance spectroscopy (AS), and drive-level capacitance profiling
(DLCP), were investigated to correlate with the phase transition beha- The solar cells were fabricated according to the conventional Mo/
vior. CuSbS2/CdS/i-ZnO/ITO/Al structure. A 60-nm CdS buffer layer was
deposited onto the sulfurized films via chemical bath deposition (CBD),
2. Experimental section as reported previously (Lee et al., 2015). After depositing the buffer
layer, intrinsic ZnO (~50 nm) and indium tin oxide (~150 nm) window
2.1. Reagents layers were deposited by radio frequency (RF) sputtering. Finally, an Al
grid of a thickness of approximately 800 nm was deposited as a current
Copper (I) acetate, C2H3CuO2 (Cu(I)Ac, 97%), antimony (III) collector using thermal evaporation. The active area of the completed
acetate, C6H9O6Sb (Sb(Ac)3, 99.99%), monoethanolamine, cells was 0.2225 cm2.
NH2CH2CH2OH (MEA, 99.0%), methanol, CH3OH (MeOH, 99.8%), and
sulfur powder (99.98%,) were purchased from Sigma-Aldrich. All che- 2.6. Characterization and analysis
micals were used as received without further purification and stored in
a nitrogen-filled glove box to restrain from the degradation caused by The chemical composition of the sulfurized films was analyzed using
humidity or air. Mo-coated soda lime glasses were fabricated by a DC energy dispersive X-ray spectroscopy (EDS; EDAX Genesis apex, accel-
magnetron sputtering system with a thickness of approximately 1 μm erating voltage of 30 kV, collection time of 100 s). The roughness and

1210
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

morphology of the surface were investigated using atomic force mi- Ra values for the 0.4 and 0.5 g films were similar, the AFM images of
croscopy (AFM, Park System XE-100, 0.6 Hz scan rate) and high-re- them (Fig. 1(d) and (e)) were quite different. Therefore, other surface
solution scanning electron microscopy (HRSEM; XL30SFEG Phillips Co., analytical methods were performed to distinguish this difference.
Holland at 10 kV), respectively. The crystal structure of the films was Along with AFM, surface SEM images were also measured for the
identified using X-ray diffraction spectroscopy (XRD; Rigaku Japan, D/ better understanding of the films morphology. Fig. 2 shows the surface
MAX-2500) with a Cu Kα radiation source (λ = 1.5406 Å). The current- morphologies of thin films before and after sulfurization with different
voltage characteristics of the devices were measured using a source amounts of S-flux. As-deposited film (Fig. 2(a)) exhibited very dense
meter (K2450, Keithley) under irradiation of simulated AM 1.5 sunlight and compact amorphous morphology without voids and cracks. In ad-
(K201, Mcscience) with a calibration standard (PVM 524, PV dition, after sulfurization, all the films prepared in different S-flux en-
Measurement Inc.). The external quantum efficiency (EQE) was de- vironments showed micron-size grains, and as the S-flux increased, the
termined using an incident photon-to-electron conversion efficiency grains became bigger due to the acceleration of S diffusion into the
(IPCE) measurement unit (PV Measurements, Inc., USA). The capaci- precursor films. From the SEM images, it was observed that the film
tance-voltage (C-V) measurements, admittance spectroscopy (AS), and prepared with 0.1 g S-flux (Fig. 2(b)) showed pinholes, whereas upon
drive-level capacitance profile (DLCP) measurements of the fabricated increasing the S-flux from 0.2 to 0.4 g, the general morphological
cells were performed with an LCR meter (Agilent 4284A, Keysight). The properties of the films such as the compactness and grain size of the
temperature of the cells was controlled by cooling (by flowing liquid films, were found to be improved without having any pinholes or voids
N2) or heating using a temperature controller (LTSE350-P, Linkam). (Fig. 2(c)–(e)). However, an excessive addition of S-flux (0.5 g) dis-
rupted the formation of dense film, resulting in a film with a notable
3. Results and discussion amount of voids (Fig. 2(f)), which was consistent with the corre-
sponding AFM results. Therefore, the SEM analysis also showed a si-
Single phase CAS thin-film solar cells were prepared via a non-va- milar trend with the AFM images of the films prepared with 0.4 and
cuum hybrid ink method. The hybrid ink consisted of metal precursors 0.5 g S-flux.
(Cu(I)Ac and Sb(Ac)3), chelating agents (MEA), and a solvent (MeOH). To investigate the effect of S-flux on the crystalline behavior of the
In the hybrid ink, the chelating agents and metal precursors make a CAS thin films, X-ray diffraction was carried out, and the corresponding
metal-chelate complex that controls the reaction rate for the formation patters are shown in Fig. 3. XRD patterns of all sulfurized films showed
of dense thin layers (Cho et al., 2014, 2013a,b,c, 2012). Using this only the orthorhombic CuSbS2 (JCPDS card no. 44-1417) phase. Among
advantage of the hybrid ink, dense CAS thin films were successfully the XRD peaks of orthorhombic CuSbS2 phase, the signature peaks
prepared, and the influences of the S-flux on the formation of CAS thin known as (1 1 1), (4 1 0), and (3 0 1) are located at approximately
film were investigated. 28–30°, shown as a color shade in Fig. 3. For the CAS thin films pre-
As mentioned before, in this study, the main experimental variable pared with an S-flux of approximately 0.1–0.4 g, the conventional
is the S-flux, which is controlled by various S-amounts during the sul- (1 1 1), (4 1 0), and (3 0 1) preferred orientations were detected. How-
furizations (thermal treatment). To date, numerous studies have been ever, as the S-flux increased from 0.4 g to 0.5 g, the intensities of typical
reported regarding the thermal treatment under different conditions, (1 1 1), (4 1 0), and (3 0 1) diffraction peaks decreased significantly, and
such as H2S (Septina et al., 2014) gas and Sb2S3 vapor (Chen et al., the intensities of the other orientation peaks, such as (2 0 0), (4 0 0),
2015; Saragih et al., 2017; Willian de Souza Lucas et al., 2016) to im- and (8 0 0) became more prominent. These XRD results also showed
prove the film quality. However, it is notable that thermal treatments good agreement with the AFM topography and SEM micrographs be-
were operated in vacuum condition for all the above-mentioned studies. cause the surface roughness and morphology for 0.5 g (Figs. 1 and 2)
It is worth mentioning that during sulfurization, S-flux and S-pressure were different from the others. For simplicity, the (1 1 1), (4 1 0), and
both are important to convert Cu and Sb precursors to CAS thin films (3 0 1) preferred orientations were denoted as (111/301) and the or-
using our method. Hence, the main factor to change the S-flux was the ientations of (2 0 0), (4 0 0) and (8 0 0) were denoted as (2 0 0) in this
variation of the S-amounts with ambient S-pressure during the thermal study.
heat treatment (sulfurizations). For CuInSe2 (CIS) and CIGS, two different preferred orientations,
(1 1 2) and (2 2 0)/(2 0 4), have been already reported (Cho et al.,
3.1. Morphology and structure 2013b; Kim et al., 2018; Rudmann et al., 2003), and the preferred or-
ientation in CIS and CIGS was controlled by regulating the Se-flux
To determine the compositions of the films, an EDS analysis of the (Chaisitsak et al., 2002; Gwak et al., 2015; Hanna et al., 2003). These
sulfurized films was performed. It revealed that there was not a notable reports relevant with our study, as illustrated in Fig. 3. Two commonly
change of the atomic ratios of Cu/Sb (0.76–0.83) with the variation of used calculation methods of the preferred orientation, such as the
the S-flux. An AFM analysis was performed to examine the surface Lotgering factor (LF) and texture co-efficient (TC), were used to quan-
texture and roughness (Ra) of the sulfurized CAS thin films prepared tify the degree of preferred orientation.
with various S-flux and is shown in Fig. 1. It was inspected from the To calculate LF and TC, XRD intensities of (hkl) plane with 2θ/θ
surface texture shown in Fig. 1(a)–(e), that the grain size increased and mode were used, and the representative graphs are shown in Fig. 4. The
expanded with the addition of S-flux. The films grown with 0.1–0.4 g S- LF and TC are defined by the following equations.
flux showed a similar surface texture in terms of grains regardless of the
size, whereas, after an additional increase of S-flux, no distinct grains dV AkT ⎡ 1 ⎤
= Rs +
were observed due to the decomposition of the grains, as the grains dJ q ⎢⎣ J − JSC ⎥
⎦ (1)
become larger coalescence of the grains was occurred. This decom-
where P(hkl) is the ratio of integrated intensities of (hkl) reflection to
position of grains for the films prepared with 0.5 g S-flux (Fig. 1(f))
the sum of the reflections in the scanned range, P0(hkl) represents the
could be explained by the excess S-flux which might affect the structure
corresponding ratio for a randomly orientated CAS phase (JCPDS card
of the absorber during the sulfurization processes. Moreover, this
no. 44-1417) (Lotgering, 1959), and
structural change could initiate voids on the surface of the films. From
Fig. 1(f), it was observed that the Ra value was approximately the same I(hkl)
I0(hkl)
or slightly increased up to 0.3 g S-flux. However, a significant increase TC(hkl) =
1 n I(hkl)
of the Ra value occurred for the 0.4 g S-flux films and afterwards it n
∑i = 1 I0(hkl) (2)
becomes steady. This indicated an increase of the surface roughness due
to the bigger grains and larger variation of grain heights. Although the where I(hkl) is the measured intensity of (hkl) plane, I0(hkl) is the

1211
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

Fig. 1. AFM images of the sulfurized thin films prepared with different S-flux: (a) 0.1 g, (b) 0.2 g, (c) 0.3 g, (d) 0.4 g, and (e) 0.5 g, and (f) variations of surface average
roughness, Ra with different S-flux.

Fig. 2. SEM images of the surface morphology of CAS thin films (a) as-deposited and grown with different S-flux: (b) 0.1 g, (c) 0.2 g, (d) 0.3 g, (e) 0.4 g, and (f) 0.5 g.

intensity of standard CAS phase, and n is the number of (hkl) diffraction Based on these calculations and the morphology from AFM and
peaks (Wang et al., 2014a,b). In this study, six influential peaks of SEM, this difference could affect to the structure of CAS; therefore, the
(2 0 0), (4 0 0), (1 1 1), (4 1 0), (3 0 1), and (8 0 0) were used for the lattice parameters of the films were calculated using the three strongest
calculations. Fig. 4(a) and (b) shows the LF and CF values of the CAS peaks for each film XRD and presented in Fig. 4(c). The CAS thin films
thin film as a function of S-flux. From Fig. 4(a) and (b), the LF and TC prepared with S-flux up to 0.4 g showed the orthorhombic crystal
for (1 1 1) and (3 0 1) decreased; however, that for (2 0 0) increased structure as usual. However, with the addition of more S (> 0.4 g),
with the addition of S-flux. From Eqs. (1) and (2), it is inferred that surprisingly a phase transition behavior was observed, which seemed to
when LF and TC are equal to 0 and 1, respectively, the films are entirely be a cubic structure from orthorhombic. Hence, this phase transition
non-oriented with the standard data. Conversely, the films with a behavior was defined as a “pseudo”-phase transition because the phase
higher deviation from 0 and 1 means highly oriented to other planes was not entirely transformed to cubic and still there were small sig-
than the standard one. Therefore, Fig. 4(a) and (b) illustrate that the nature peaks from the orthorhombic phase. This pseudo-phase transi-
film prepared with 0.4 g S-flux results the best match with the standard tion occurred because of the volume expansion of the unit cell resulting
data having (111/301) preferred orientations. On the other hand, the from the allocation of the excess S atom.
film prepared with a 0.5 g S-flux was highly oriented to the (2 0 0) Therefore, from the above discussion, it is attributed that a single-
plane. phase CAS could be formed for an optimum ink composition and

1212
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

of orientation in CAS was more severe than that in CIS/CIGS because


the phase transition behavior was observed in excess S-flux condition.
Briefly, to make pure orthorhombic CAS phase, 0.4 g S-flux was the
optimum and a pseudo-phase transition can be occurred when the S-
flux exceed this limit.

3.2. Device performance and temperature-dependent PV characterizations

CAS devices were fabricated by depositing CdS/i-ZnO/ITO/Al layers


on the CuSbS2 absorbers grown with various S-flux. Fig. 5 shows the
illuminated current density-voltage (JV) and external quantum effi-
ciency (EQE) curves for the devices with different amounts of S-flux,
with the calculated band gap (Eg) from the corresponding EQE shown in
the inset. The device parameters extracted from the JV measurements
are summarized in Table 1. Upon increasing the S-flux, the device
Fig. 3. X-ray diffraction pattern of CAS thin films grown with different S-flux. A parameters, such as JSC, VOC, and η, were increased and the FF was
reference pattern of orthorhombic CuSbS2 (PDF#44-1417) is included at the almost constant up to the optimum amount of S-flux (to make orthor-
bottom. hombic CAS phase) due to the improved carrier collection, structure,
and morphology. However, when the amount of S-flux exceeded the
sulfurized condition, such as temperature and pressure, regardless of optimum level, excessive S-flux negatively affected the device perfor-
the S-flux. However, the preferred orientation of CAS could be regu- mance owing to the disrupted morphology of the film. In particular, the
lated by the S-flux, similar to CIS/CIGS system, where Se-flux plays variation of JSC was more prominent in the JV curves, as shown in
critical role to control the orientations. It is worth noting that the effect Table 1. The change of JSC was further analyzed by EQE curves shown
in Fig. 5(b). EQE also followed the same trend as the device

Fig. 4. (a) LF, (b) CF for (1 1 1), (3 0 1), and (2 0 0) planes of the CAS thin films as a function of S-flux, and (c) evolution of the unit cell parameters with S-flux in CAS
thin films during the sulfurizations.

1213
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

Fig. 5. (a) Light JV and (b) EQE curves of CAS devices with various S-flux. The inset shows the calculated band gap from the corresponding EQE.

Table 1 performance, i.e., was improved until an optimum S-flux of 0.4 g which
Device parameters for CAS solar cells at room temperature with various S-flux. reflected a better carrier collection through the CAS absorber layer, and
S-flux (g) VOC (V) JSC (mA/cm2) FF (%) η (%) when the S-flux was greater than the optimum, the EQE decreased
drastically. In addition, the optical band gaps of the absorber layers
0.1 0.27 2.33 41.71 0.26 were also deduced from the long-wavelength cutoff of the EQE curves
0.2 0.35 3.45 42.42 0.51
for all devices, as shown in inset of Fig. 5(b). The band gaps of all
0.3 0.35 5.38 39.62 0.74
0.4 0.38 9.86 40.85 1.53
devices were calculated from [hν × ln (1 − EQE)]2 vs. hν plots and
0.5 0.29 3.59 34.61 0.36 presented in Table 2. The calculated band gaps for all devices were
approximately the same and perfectly matched with the band gap of
CuSbS2 (Banu et al., 2019, 2016). To verify these influences, electrical
Table 2 characteristics and defect properties of the absorber layer were eval-
Calculated band gap, Eg from EQE, barrier height, Фb from the fitting of ex- uated in the next section.
ponential (RsT vs. 1000/T) plot, hopping energy, Δ by fitting the exponential To obtain additional insight into the origin of the difference in the
temperature dependent JSC curve, and activation energy, Ea at 0 K from VOC vs. device performance of CAS thin-film solar cells with the variation of the
T curve, of CAS solar cells with different S-flux. S-flux, the carrier transport mechanisms of the devices were analyzed.
S-flux (g) Eg (eV) Фb (meV) Δ (meV) Ea (eV) In this regard, temperature dependent current density-voltage (JV-T)
measurements were performed for all devices in the temperature range
0.1 1.56 209 102.61 0.61
of T = 100–300 K.
0.2 1.55 155 103.03 0.84
0.3 1.54 147 129.12 0.86 It was discussed in the previous section that the variation of η values
0.4 1.53 143 165.45 0.87 for the solar cells made with different S-flux was mainly observed due
0.5 1.57 147 142.65 0.66 to the larger variation of JSC, which gave the insight on the carrier
collection problem. To further elucidate this issue, temperature-de-
pendent JSC and η values were determined from the JV-T measurements
and are shown in Fig. 6. It was noticed from Fig. 6 that JSC and η values

Fig. 6. (a) JSC and (b) η values obtained from the CAS solar cells for various S-flux under different temperatures.

1214
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

Fig. 7. Temperature dependence of the dark series resistance, RS, for the sulfurized thin films prepared with different S-flux: (a) 0.1 g, (b) 0.2 g, (c) 0.3 g, (d) 0.4 g,
and (e) 0.5 g. The inset shows the corresponding back-contact barrier height fitting.

of all studied devices were collapsing toward zero at lower temperature. temperatures are required to activate the defects, and these high-energy
This rapid declination of JSC as well as η, particularly in the low tem- defects can act as recombination centers (Lee et al., 2016). Therefore,
perature region, could be explained by the fast recombination of the deep defects can limit JSC owing to the fast recombination rate. This
photo-generated charges before they were collected (Banu et al., 2019; result will be further supported by AS analysis in the next section.
Lee et al., 2016). On the other hand, using a standard diode analysis of a thin film
Generally, the limitations of JSC due to fast recombination are solar cell, series resistance, RS was extracted from the y-intercept of the
negligible for device containing shallow defects states (close to the dV/dJ vs. 1/J plot as a function of temperature (Fig. 7),
valence or conduction band) because of easy capture and release of dV AkT 1
dJ
= Rs + ⎡ ⎤
q ⎣ J − JSC ⎦
(3)where q is the charge of an electron, V is
carriers by defect sites. However, if the device contains deep defect the applied bias, J is the current density, A is the diode ideality factor, k
states located in the middle of the valence and conduction band, higher

1215
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

is Boltzmanńs constant and T is temperature (Hegedus and Shafarman, increased gradually; however, for a further increase of S-flux, Δ started
2004; Sites and Mauk, 1989). From Fig. 7, increasing behaviors of series to decrease. Therefore, the hopping energy of the device made with
resistance toward a low temperature were observed. This increasing 0.4 g S-flux was significantly higher than that of the other devices. This
behavior of series resistance toward low temperature could be ex- results highlighted the existence of fewer defect states and less re-
plained by the back contact barrier, which could block the carriers to combination loss for the 0.4 g S-flux device, resulting in a higher JSC
transport at the interface of Mo and CAS. Hence, back contact barrier than that of the other S-flux.
height, ɸB, were extracted from the temperature dependent dark series To elucidate the effect of S-flux on recombination behavior, tem-
resistance using the following equation: perature-dependent VOC of the CAS solar cells for various S-flux were
performed and compared (Fig. 8(b)). Usually, the activation energy, Ea
k ϕ
Rs = R 0 + exp ⎛ B ⎞
⎜ ⎟ extracted from the temperature-dependent VOC provides information on
qA* T kT
⎝ ⎠ (4) the dominant recombination process in the device. Ea was determined
where A* is the effective Richardson constant and R0 is the background by extrapolating a plot of VOC vs. T at T = 0 K using the following
series resistance resulting from the contact grid and TCO (Gunawan equation and presented in Table 2.
et al., 2010; Wang et al., 2010). Corresponding fitting curve for ɸB are Ea AkT ⎛ J00 ⎞
shown in the inset of Fig. 7, and the values are summarized in Table 2. VOC = − ln ⎜ ⎟

q q ⎝ JL ⎠ (6)
When the S-flux increased up to an optimum level, ɸB decreased due to
the improved morphology and crystallinity. However, when the S-flux where JL is the photo-generated current density (Cao et al., 2011). For
increased more than the optimum level, it decreased again, which could all devices, the Ea values (Table 2) were significantly lower than their
be the results of structural change in the film due to excessive amount respective bandgap values, suggesting that the recombination me-
of S. Around room temperature, the effect of ɸB at the interface of Mo chanism in these CAS cells was strongly influenced by interface re-
and CAS was not noticeable. However, as the temperature gradually combination (Nadenau et al., 2000; Riedel et al., 2004; Wang et al.,
decreased, the carrier transport was more affected by ɸB. Thus, it can be 2014a,b), which could also cause the low VOC of the devices at room
asserted that the dramatic change of RS depending on temperature was temperature.
another reason for rapid declination of JSC with temperature. At room
temperature, there was no problem regarding solar cell operation be- 3.3. Defect properties
cause carrier could overcome that barrier even though S-flux increased.
However, as temperature decreased, recombination rate was influenced To evaluate the quality of the formed p-n junctions in the fabricated
by the S-flux; therefore, ɸB was different for various S-flux. CAS devices, a room temperature C-V analysis was performed. Fig. 9
Furthermore, such a strong temperature dependence of JSC also can shows the Mott-Schottky plots of the CAS devices with different S-flux,
be explained by hopping-type charge-carrier transport in the device exhibiting common p-type characteristics (negative values of the
(Banu et al., 2019). Hopping energy, Δ is the energy difference between slopes) of the films. The net carrier density, NC-V of the devices was
the localized defect states. Therefore, high value of Δ represents fewer calculated from the slopes of the 1/C2–V plots using the following
defect states in the bulk of the device, resulting in a lower re- equation:
combination. In contrast, a lower value of Δ indicates larger defect
C 3 ⎛ dC ⎞−1
states in the absorber layer, which could accelerate the recombination NC − V =
qεε0 A2 ⎝ dV ⎠ (7)
rate. Thus, to determine the source of the recombination loss as a result
of the higher variation of JSC, Δ was deduced for all devices made with where ε is the dielectric constant of CAS, ε0 is the dielectric constant of
different S-flux by fitting the exponential temperature-dependent JSC vacuum, A is the sample area, and C is the capacitance. Corresponding
curve shown in Fig. 8(a) using the following equation: data listed in Table 3. No notable difference of NC-V among the devices
Δ with different S-flux was found, and the values were in the range of the
JSC\; = \;J00 \;exp ⎛−\; ⎞ previously reported value (Banu et al., 2019). In general, the room
⎝ kT ⎠ (5)
temperature C-V analysis provides information regarding the overall
where Joo is the saturation current density prefactor. The values of Δ are carrier density, including interface and bulk. Therefore, it is unable to
presented in Table 2. It is notable that with the addition of S-flux up to distinguish between the interface and bulk defects. However, it is al-
0.4 g, i.e., until the optimum level for this deposition method, Δ ready verified from the Ea values of temperature dependent VOC

Fig. 8. (a) JSC vs. 1000/T curves to calculate the hopping energy, Δ, and (b) VOC vs. T curves and its linear extrapolation to 0 K indicating the activation energy, Ea of
the recombination process in CAS solar cells with different S-flux.

1216
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

temperature range of 100–300 K, as shown in Fig. 10(a)–(e). The ca-


pacitance-frequency (C-F) spectra of the devices were obtained in the
dark at a zero bias from 100 Hz to 1 MHz. Usually, in AS, the low fre-
quency capacitance provides information for the deep traps and free
carriers densities, while the high-frequency capacitance is responsible
only for the free carrier density. From Fig. 10(a)–(e), it is noticed that
the capacitance of the device with optimum S-flux (Fig. (d)) showed a
smaller deviation as a function of frequency than those of the others,
indicating smaller trap densities in the absorber.
Fig. 10(f) shows an Arrhenius plot for the inflection point ω0 of the
C-F plot (Fig. 10(a)–(e)), where, ω0 was determined from the peak of
the – dC/d(ln(ω)) vs. ω plots. The defect activation energy, Et can be
calculated from the Arrhenius plot of ln(ω0/T2), using the following
equation (Duan et al., 2013; Kim et al., 2018; Chen et al., 2011):

E
ω0 = 2πυ0 T 2 exp ⎛− t ⎞
⎝ kT ⎠ (8)

where υ0 is the emission factor. The activation energy, Et, for the de-
Fig. 9. 1/C2 versus V curve for various S-flux.
vices prepared with 0.1, 0.2, 0.3, 0.4, and 0.5 g of S-flux were estimated
as 0.187, 0.186, 0.195, 0.202, and 0.237 eV, respectively, confirming
Table 3 the presence of deep defects. Comparing the Et values, it was observed
Carrier densities derived from C-V and DLCP for the CAS solar cells with dif- that the value increased with the addition of S-flux. A similar trend also
ferent S-flux.
was observed in CZTSSe solar cells, where it was reported that the
S-flux (g) NC-V (cm−3) NB (cm−3) NIF (cm−3) defect ionization energies increased as the S-concentration increased,
resulting in the down shifting of the valance band (Chen et al., 2011;
0.1 2.69 × 1017 7.97 × 1016 1.89 × 1017
0.2 2.10 × 1017 3.88 × 1016 1.71 × 1017
Duan et al., 2013). On the other hand, Willian de Souza Lucas et al.
0.3 1.73 × 1017 4.00 × 1016 1.33 × 1017 reported two deep defect states in CAS absorber at 0.17 eV below the
0.4 1.66 × 1017 2.94 × 1016 1.37 × 1017 conduction band minimum (CBM) and 0.24 eV above the valance band
0.5 1.82 × 1017 5.81 × 1016 1.29 × 1017 maximum (VBM), which were related to the VS donors and CuSb ac-
ceptors, respectively (Willian de Souza Lucas et al., 2017). Further-
more, according to Zhang et al. and Wei et al., the activation energy of
(Fig. 8(b)) showed that the interface defects were dominant. Thus, in
0.202 eV could be responsible for the VSb defects above the VBM or the
order to identify the defect types, an additional analytical method, such
Cui defects below the CBM (Wei et al., 1998; Zhang et al., 1998).
as DLCP, was required and will be discussed in next section (Heath
However, EDS value proved that all the films were Cu-poor; therefore,
et al., 2004, 2003).
the Cui defects were excluded in this study. Thus, it can be declared that
An AS analysis of all CAS devices was performed to estimate the
0.202 eV for optimum S-flux device was related to the VSb defects.
defect energy level inside the band gap of solar cells over the
Usually, defect levels below the CBM are referred to as n-type defects,

Fig. 10. Admittance spectra (AS) of CAS solar cells with different S-flux: (a) 0.1 g, (b) 0.2 g, (c) 0.3 g, (d) 0.4 g, and (e) 0.5 g measured in the temperature range of
300–100 K, (f) corresponding Arrhenius plot of the inflection point of the capacitance function calculated from the derivative of the AS.

1217
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

Fig. 11. Deep level response of CAS solar cell device with different S-flux: (a) 0.1 g, (b) 0.2 g, (c) 0.3 g, (d) 0.4 g, and (e) 0.5 g measured at the temperature range
140–300 K.

and those above the VBM are referred to as p-type defects. Concisely, Table 4
from the AS results, it was estimated that the low S-flux mainly con- Free carrier densities (P) and defect densities (NT) derived from the DLCP of
tained n-type (VS) defects, and with the increment of S-flux up to the CAS solar cells with different S-flux.
optimum (0.4 g), n-type (VS) defects were gradually suppressed by S- S-flux (g) P (cm−3) NT (cm−3)
flux. However, as the S-flux increased more than the optimum, another
p-type defects CuSb, which were deeper than VSb, were prominent. 0.1 5.03 × 1015 7.47 × 1016
0.2 4.45 × 1015 3.43 × 1016
Therefore, although the films up to 0.4 g (optimum) contained an or-
0.3 3.86 × 1015 3.61 × 1016
thorhombic structure and a similar morphology regardless of the grain 0.4 6.73 × 1015 2.26 × 1016
size, the JSC and η values increased owing to the reduction of the n-type 0.5 8.65 × 1015 4.94 × 1016
defects VS to p-type defects VSb. Nevertheless, for an additional incre-
ment of the S-flux, the JSC and η values decreased due to the structural
change and existence of deeper defects. Moreover, it is also confirmed response from both free carriers and deep defects (NB). Thus, the defect
that the recombination rate was affected by the existing deep defects in density (NT) in the absorber could be extracted by subtracting the NDL
the devices, and, as a results of that, JSC decreased with the decline of values at a high temperature (NB) from a low temperature (P) (Duan
temperature. et al., 2013; Heath et al., 2004). As presented in Table 4, the influence
To identify the different types of defects (bulk/interface defects, free of S-flux on free carrier densities (P) was insignificant. However, the
carriers, and defect densities), DLCP measurements were performed defect densities (NT) of the devices, which were derived from the dif-
using the technique introduced by Heath et al. (2004). DLCP profile of ference between the NDL values at T = 300 K (NB in Table 3) and 140 K
all the CAS devices using a fixed frequency of 10 kHz as a function of (P in Table 4), were found to decrease with the increment of S-flux up to
temperature are shown in Fig. 11 and presented in Table 3. In order to 0.4 g (optimum). Moreover, as the S-flux increased more (> optimum),
distinguish the bulk and interface defects, the results of room tem- NT increased. Briefly, the device with optimum S-flux contained the
perature DLCP profile were compared with C-V data for all the devices. lowest NT in the bulk than those of the other devices, which showed
As it is mentioned that C-V data provides the overall information for the good agreement with the hopping energy, C-F plot (Fig. 10(a)–(e)), and
devices, including the interface and bulk states. However, DLCP is more device performance.
sensitive to bulk states. Thus, the defect densities at the interface (NIF) From the above discussion, it can be interpreted that sufficient S-
were calculated by subtracting the DLCP data (NB) from the C-V data flux is beneficial to reduce the n-type defects. However, an excessive S-
(NC-V) at room temperature, as listed in Table 3. The NIF values of all the flux could cause structural change in the films, which could negatively
devices were an order of magnitude larger than those of NB. Therefore, affect the device performance. Therefore, the optimum S-flux is needed
it is demonstrated that interface recombination was dominant for all to fabricate orthorhombic CAS thin films with improved device per-
CAS solar cells, which was in good agreement with the Ea values from formance. However, the optimum S-flux device also showed a dominant
the VOC vs. T plot. interface recombination. Thus, reducing the interface recombination
Besides, to further elucidate the effect of S-flux on the bulk prop- could be future steps to further improve the CAS device performance.
erties of the films, free carrier (P) and defect densities (NT) of the de-
vices were derived (Table 4). Traditionally, the carrier density (NDL) 4. Conclusions
extracted from a low temperature DLCP profile represents the free
carrier density (P), whereas a high temperature NDL indicates the CAS solar cells were fabricated for various S-flux via a non-vacuum

1218
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

hybrid ink method and subsequent sulfurization process using the op- Energy Mater. Sol. Cells 110, 126–132.
timum ink composition and sulfurization condition (temperature, time, Cho, A., Ahn, S.J., Yun, J.H., Gwak, J., Ahn, S.K., Shin, K., Yoo, J., Song, H., Yoon, K.,
2013c. The growth of Cu2-xSe thin films using nanoparticles. Thin Solid Films 546,
and pressure). The effect of S-flux on the optoelectronic and defect 299–307.
properties of CAS solar cells was investigated by the aim to reduce the Cho, A., Song, H., Gwak, J., Eo, Y.J., Yun, J.H., Yoon, K., Ahn, S.J., 2014. A chelating
defect energy levels. However, with the change of S-flux, surprisingly a effect in hybrid inks for non-vacuum-processed CuInSe2 thin films. J. Mater. Chem. A
2, 5087–5094.
phase transition behavior was observed. The films prepared until 0.4 g Cho, A., Banu, S., Kim, K., Park, J.H., Yun, J.H., Cho, J.S., Yoo, J.S., 2017. Selective thin
(optimum) S-flux showed an orthorhombic structure with improved film synthesis of copper-antimony-sulfide using hybrid ink. Sol. Energy 145, 42–51.
grains, whereas a pseudo-phase transition from an orthorhombic to a Duan, H.S., Yang, W., Bob, B., Hsu, C.J., Lei, B., Yang, Y., 2013. The role of sulfur in
solution-processed Cu2ZnSn(S, Se)4 and its effect on defect properties. Adv. Funct.
cubic structure was observed for the films prepared with excessive S- Mater. 23, 1466–1471.
flux. In addition, below the optimum S-flux, photovoltaic properties, Gunawan, O., Todorov, T.K., Mitzi, D.B., 2010. Loss mechanisms in hydrazine-processed
specifically JSC, were limited owing to the morphology with smaller Cu2ZnSn(Se, S)4 solar cells. Appl. Phys. Lett. 97, 233506.
Gwak, J., Lee, M., Yun, H., Ahn, S., Cho, A., Ahn, S., Hulme, J.P., 2015. Selenium flux
grains. However, for an excessive S-flux, JSC and other device para-
effect on Cu(In, Ga)Se2 thin films grown by a 3-stage co-evaporation process. Isr. J.
meters were mainly affected by the pseudo-phase transition. From the Chem. 55, 1115–1122.
results of the recombination mechanism and defect properties, it was Hanna, G., Mattheis, J., Laptev, V., Yamamoto, Y., Rau, U., Schock, H.W., 2003. Influence
confirmed that all devices showed a dominant interface recombination. of the selenium flux on the growth of Cu(In, Ga)Se2 thin films. Thin Solid Films
431–432, 31–36.
Moreover, the CAS solar cell deposited with a low S-flux exhibited Heath, J.T., Cohen, J.D., Shafarman, W.N., 2004. Bulk and metastable defects in CuIn1-
dominant n-type VS defects, whereas the device containing excessive S- xGaxSe2 thin films using drive-level capacitance profiling. J. Appl. Phys. 95,

flux showed a deeper defect energy level with dominant CuSb defects. 1000–1010.
Heath, J.T., Cohen, J.D., Shafarman, W.N., 2003. Distinguishing metastable changes in
Therefore, controlling the S-flux is important for fabricating orthor- bulk CIGS defect densities from interface effects. Thin Solid Films 431, 426–430.
hombic CAS thin films as well as improving the photovoltaic perfor- Hegedus, S.S., Shafarman, W.N., 2004. Thin-film solar cells: device measurements and
mance by reducing the defects in the bulk state. analysis. Prog. Photovoltaics Res. Appl. 12, 155–176.
Kim, S., Yoo, H., Rana, T.R., Enkhbat, T., Han, G., Kim, J., Song, S., Kim, K., Gwak, J., Eo,
Y., Yun, J.H., 2018. Effect of crystal orientation and conduction band grading of
Acknowledgements absorber on efficiency of Cu(In, Ga)Se2 solar cells grown on flexible polyimide foil at
low temperature. Adv. Energy Mater. 8, 1801501.
Lee, D., Yang, J.Y., Kim, Y.S., Mo, C.B., Park, S., Kim, B.J., Kim, D., Nam, J., Kang, Y.,
This study was conducted under the framework of the Research and 2016. Effects of the Cu/(Ga+In) ratio on the bulk and interface properties of Cu
Development Program of the Korea Institute of Energy Research (KIER) (InGa)(SSe)2 solar cells. Sol. Energy Mater. Sol. Cells 149, 195–203.
(B9-2411-01). This work was also supported by the Korea Institute of Lee, S., Lee, E.S., Kim, T.Y., Cho, J.S., Eo, Y.J., Yun, J.H., Cho, A., 2015. Effect of an-
nealing treatment on CdS/CIGS thin film solar cells depending on different CdS de-
Energy Technology Evaluation and Planning (KETEP), granded fi-
position temperatures. Sol. Energy Mater. Sol. Cells 141, 299–308.
nancial resource from the Ministry of Trade, Industry & Energy Lotgering, F.K., 1959. Topotactical reactions with ferrimagnetic oxides having hexagonal
(MOTIE) of the Republic of Korea (No. 20173010012980, crystal structures-I. J. Inorg. Nucl. Chem. 9, 113–123.
20163010012430). Macías, C., Lugo, S., Benítez, Á., López, I., Kharissov, B., Vázquez, A., Peña, Y., 2017. Thin
film solar cell based on CuSbS2 absorber prepared by chemical bath deposition (CBD).
Mater. Res. Bull. 87, 161–166.
Appendix A. Supplementary material Moon, D.G., Cho, A., Park, J.H., Ahn, S., Kwon, H., Cho, Y.S., Ahn, S., 2014. Iron pyrite
thin films deposited via non-vacuum direct coating of iron-salt/ethanol-based pre-
cursor solutions. J. Mater. Chem. A 2, 17779–17786.
Supplementary data to this article can be found online at https:// Nadenau, V., Rau, U., Jasenek, A., Schock, H.W., Introduction, I., Cugase, G., Rau, U.,
doi.org/10.1016/j.solener.2019.07.019. Nadenau, V., Schock, H.W., Introduction, I., Cugase, G., 2000. Electronic properties
of CuGaSe2-based heterojunction solar cells. Part I. Transport analysis. J. Appl. Phys.
87, 584–593.
References Riedel, I., Parisi, J., Dyakonov, V., Lutsen, L., Vanderzande, D., Hummelen, J.C., 2004.
Effect of temperature and illumination on the electrical characteristics of polymer-
Banu, S., Ahn, S.J., Ahn, S.K., Yoon, K., Cho, A., 2016. Fabrication and characterization of fullerene bulk-heterojunction solar cells. Adv. Funct. Mater. 14, 38–44.
cost-efficient CuSbS2 thin film solar cells using hybrid inks. Sol. Energy Mater. Sol. Rodriguez-Lazcano, Y., Nair, M.T.S., Nair, P.K., 2005. Photovoltaic p-i-n structure of
Cells 151, 14–23. Sb2S3 and CuSbS2 absorber films obtained via chemical bath deposition. J.
Banu, S., Ahn, S.J., Eo, Y.J., Gwak, J., Cho, A., 2017. Tin monosulfide (SnS) thin films Electrochem. Soc. 152, 635–638.
grown by liquid-phase deposition. Sol. Energy 145, 33–41. Rudmann, D., Bilger, G., Kaelin, M., Haug, F., Zogg, H., Tiwari, A.N., 2003. Effects of NaF
Banu, S., Cho, Y., Kim, K., Kyu, S., Sik, J., Gwak, J., 2019. Solar energy materials and solar coevaporation on structural properties of Cu(In, Ga)Se2 thin films. Thin Solid Films
cells effect of Cu content in CuSbS2 thin films using hybrid inks: their photovoltaic 431–432, 37–40.
properties and defect characteristics. Sol. Energy Mater. 189, 214–223. Saragih, A.D., Daniel, A., Kuo, S.D., 2017. Thin film solar cell based on p-CuSbS2 together
Berry, N., Cheng, M., Perkins, C.L., Limpinsel, M., Hemminger, J.C., Law, M., 2012. with Cd-free GaN/InGaN bilayer. J. Mater. Sci. Mater. Electron. 28, 2996–3003.
Atmospheric-pressure chemical vapor deposition of iron pyrite thin films. Adv. Septina, W., Ikeda, S., Iga, Y., Harada, T., Matsumura, M., 2014. Thin film solar cell based
Energy Mater. 2, 1124–1135. on CuSbS2 absorber fabricated from an electrochemically deposited metal stack. Thin
Cao, Q., Gunawan, O., Copel, M., Reuter, K.B., Chey, S.J., Deline, V.R., Mitzi, D.B., 2011. Solid Films 550, 700–704.
Defects in Cu(In, Ga)Se2 chalcopyrite semiconductors: a comparative study of ma- Sinsermsuksakul, P., Sun, L., Lee, S.W., Park, H.H., Kim, S.B., Yang, C., Gordon, R.G.,
terial properties, defect states, and photovoltaic performance. Adv. Energy Mater. 1, 2014. Overcoming efficiency limitations of SnS-based solar cells. Adv. Energy Mater.
845–853. 4, 1–7.
Chaisitsak, S., Yamada, A., Konagai, M., 2002. Preferred orientation control of Cu(In1- Sites, J.R., Mauk, P.H., 1989. Diode quality factor determination for thin-film solar cells.
xGax)Se2 (x≈ 0.28) thin films and its influence on solar cell characteristics. Jpn. J.
Sol. cells 27, 411–417.
Appl. Phys. 41, 507. Tiwari, D., Koehler, T., Klenk, R., Fermin, D.J., 2017. Solution processed single-phase
Chen, S., Walsh, A., Yang, J., Gong, X.G., Sun, L., Yang, P., Chu, J., Wei, S., 2011. Cu2SnS3 films: structure and photovoltaic performance. Sustain. Energy Fuels 1,
Compositional dependence of structural and electronic properties of Cu2ZnSn(S, Se)4 899–906.
alloys for thin film solar cells. Phys. Rev. B 83, 125201. Wan, L., Ma, C., Hu, K., Zhou, R., Mao, X., Pan, S., Helena, L., Xu, J., 2016. Two-stage co-
Chen, W., Kuo, D., Tran, T.H.I., Tuan, A.N.H., 2015. Preparation of CuSbS2 thin films by evaporated CuSbS2 thin films for solar cells. J Alloys Compd. 680, 182–190.
co-sputtering and solar cell devices with band gap-adjustable n-type InGaN as a Wang, J.T., Shi, X.L., Liu, W.W., Zhong, X.H., Wang, J.N., Pyrah, L., Sanderson, K.D.,
substitute of ZnO. J. Electron. Mater. 45, 688–694. Ramsey, P.M., Hirata, M., Tsuri, K., 2014a. Influence of preferred orientation on the
Chen, X., Li, Z., Zhu, H., Wang, Y., Liang, B., Chen, J., Xu, Y., Mai, Y., 2017. CdS/Sb2S3 electrical conductivity of fluorine-doped tin oxide films. Sci. Rep. 4, 3679.
heterojunction thin film solar cells with thermally-evaporated absorber. J. Mater. Wang, K., Gunawan, O., Todorov, T., Shin, B., Chey, S.J., Bojarczuk, N.A., Mitzi, D., Guha,
Chem. C 5, 9421–9428. S., 2010. Thermally evaporated Cu2ZnSnS4 solar cells. Appl. Phys. Lett. 97, 143508.
Cho, A., Ahn, S., Yun, J.H., Gwak, J., Song, H., Yoon, K., 2012. A hybrid ink of binary Wang, W., Winkler, M.T., Gunawan, O., Gokmen, T., Todorov, T.K., Zhu, Y., Mitzi, D.B.,
copper sulfide nanoparticles and indium precursor solution for a dense CuInSe2 ab- 2014b. Device characteristics of CZTSSe thin-film solar cells with 12.6% efficiency.
sorber thin film and its photovoltaic performance. J. Mater. Chem. 22, 17893. Adv. Energy Mater. 4, 1–5.
Cho, A., Ahn, S., Ho Yun, J., Gwak, J., Kyu Ahn, S., Shin, K., Song, H., Hoon Yoon, K., Wei, S., Zhang, S.B., Zunger, A., 1998. Effects of Ga addition to CuInSe2 on its electronic,
2013a. Non-vacuum processed CuInSe2 thin films fabricated with a hybrid ink. Sol. structural, and defect properties. Appl. Phys. Lett. 72, 3199–3201.
Energy Mater. Sol. Cells 109, 17–25. Welch, A.W., Baranowski, L.L., Zawadzki, P., Dehart, C., Johnston, S., Lany, S., Wolden,
Cho, A., Ahn, S., Yun, J.H., Eo, Y.J., Song, H., Yoon, K., 2013b. Carbon layer reduction via C.A., Zakutayev, A., 2015a. Accelerated development of CuSbS2 thin film photo-
a hybrid ink of binary nanoparticles in non-vacuum-processed CuInSe2 thin films. Sol. voltaic device prototypes. Prog. Photovolt.: Res. Appl. 24, 1–29.

1219
S. Banu, et al. Solar Energy 188 (2019) 1209–1220

Welch, A.W., Zawadzki, P.P., Lany, S., Wolden, C.A., Zakutayev, A., 2015b. Solar energy J., 2014. CuSbS2 as a promising earth-abundant photovoltaic absorber material: a
materials & solar cells self-regulated growth and tunable properties of CuSbS2 solar combined theoretical and experimental study. Chem. Mater. 26, 3135–3143.
absorbers. Sol. Energy Mater. Sol. Cells 132, 499–506. Yu, L., Kokenyesi, R.S., Keszler, D.A., Zunger, A., 2013. Inverse design of high absorption
Willian de Souza Lucas, F., Lucas, D.S., Welch, A.W., Baranowski, L.L., Dippo, P.C., thin-film photovoltaic materials. Adv. Energy Mater. 3, 43–48.
Hempel, H., Unold, T., Eichberger, R., Blank, B., Rau, U., Mascaro, L.H., Zakutayev, Zhang, S.B., Wei, S., Zunger, A., 1998. Defect physics of the CuInSe2 chalcopyrite semi-
A., 2016. Effects of thermochemical treatment on CuSbS2 photovoltaic absorber conductor. Phys. Rev. B 57, 9642.
quality and solar cell reproducibility. J. Phys. Chem. C 3, 18377–18385. Zhang, Y., Huang, J., Yan, C., Sun, K., Cui, X., Liu, F., Liu, Z., Zhang, X., Liu, X., Stride,
Willian de Souza, F., Lucas, Peng, H., Johnston, S., Dippo, P.C., Lany, S., Mascaro, L.H., J.A., Green, M.A., Hao, X., 2018. High open-circuit voltage CuSbS2 solar cells
Zakutayev, A., 2017. Characterization of defects in copper antimony disulfide. J. achieved through the formation of epitaxial growth of CdS/CuSbS2 hetero-interface
Mater. Chem. A 21986–21993. by post-annealing treatment. Prog. Photovolt.: Res. Appl. 1–7.
Yang, B., Wang, L., Han, J., Zhou, Y., Song, H., Chen, S., Zhong, J., Lv, L., Niu, D., Tang,

1220

You might also like