A hierarchical composite of Sb2S3 nanorods grown on zinc oxide (ZnO) nanofiber was prepared, and the formation of comb-shaped Sb2S3 nanorod arrays on the ZnO nanofibers was confirmed. It was found that the size of the diameter and the density of the nanorods are regulatable by changing the concentration of polyvinyl pyrrolidone as an additive for the growth of Sb2S3 nanorod on ZnO nanofiber. The obtained Sb2S3 nanorod arrays were applied as a light absorber for thin-film solar cells composed of glass-fluorine-doped tin oxide/compact ZnO/ZnO nanofibers−ZnS/Sb2S3 nanorod arrays/poly(3-hexylthiophene-2,5-diyl)/MoOx/Ag. The rectification ratio and photocurrent generation efficiency of the comb-shaped Sb2S3 nanorod arrays were improved as compared with the heterojunction of randomly stacked Sb2S3 nanorods. Smaller series resistance (Rs) of 8.13 Ω cm−2 and an ideality factor (n) of 2.84 with the comb-shaped Sb2S3 nanorod arrays than those of the randomly stacked ones of Rs = 15.01 Ω cm−2 and n = 3.83 also indicated superior charge extraction property and suppressed recombination of the comb-shaped Sb2S3 nanorod arrays at the interface.

Research on thin-film solar cells that use copper indium gallium selenide (Sulfide) (CuInGeSe or CuInGeS), cadmium telluride (CdTe), and antimony sulfide (Sb2S3) to replace silicon has been developed, and Sb2S3 as a binary chalcogenide has recently become one of the promising materials for the photovoltaic light absorber.

Sb2S3 has a bandgap of ∼1.7 eV and a large absorption coefficient (>105 cm−1) in visible light,1,2 together with its low toxicity, stability, and abundance on earth, making it a potential material for the photovoltaic light absorber. The power conversion efficiency (PCE) of 7.5% was obtained in the bulk heterojunction solar cells based on the Sb2S3-sensitized mesoporous TiO2 films.3 

The most commonly discovered Sb2S3 solid state solar cell architecture consists of a transparent conducting oxide coated glass substrate, a compact semiconductor metal oxide layer, an Sb2S3 layer, a hole transporting layer, and metal contacts.3–16 The working principle of such sort of solar cells can be illustrated as follows: under illumination, the active layer—Sb2S3 layer is excited, producing exciton pair. Thereafter, the electrons are injected into the conduction band of n-type semiconduction oxides, such as titanium dioxide (TiO2) or zinc oxide (ZnO), while the holes are transported to the opposite direction into the p-type hole transporting materials, such as poly[3-hexylthiophene-2,5-diyl (P3HT)] or 2,2′,7,7′-tetrakis(N,N-di-p-methoxyphenyl-amine)9,9′-spirobifluorene (Spiro-OMeTAD). Finally, the electrons and holes are collected at conductive electrodes, usually fluorine doped tin oxide (FTO) or indium doped tin oxide (ITO) and metal electrode, respectively.16 However, for planar structure devices, relatively small hole diffusion length (LD = 180 ± 60 nm) in Sb2S3 restricted the injection of holes into hole transport materials.17 Moreover, a large number of grain boundaries in the nanocrystalline films will lead to charge recombination, resulting in reduced performance.2,16,18 These issues have led to the exploration of solar cell structures to enhance the charge transport properties.

Using one dimensional (1D) nanostructure as a substitution for the planar contact between the active layer and hole or electron transport layer is one of the promising strategies to suppress the charge recombination and provide a direct pathway along the long axis of 1D nanostructure for electron transport.18–24 Electrons are considered to be several orders of magnitude faster to transport in 1D nanostructures, such as nanorods, nanowires, nanofibers, and nanocolumns.25–28 In this context, a combination of Sb2S3 with 1D electron transport material has been developed; for example, Ying et al. used Sb2S3-sensitized TiO2 nanorod arrays and prepared solid state solar cells with a PCE of 5.37%.29 Sun et al. designed ZnO nanorod arrays with Cu-doped Sb2S3 quantum dot and prepared an Sb2S3 quantum dot sensitized solar cell with a PCE of 3.14%.30 Parize et al. reported a chemical spray pyrolysis method to cover uniform ultra-thin Sb2S3 as a light absorbing shell on ZnO/TiO2 core–shell nanowire and achieved a PCE of 2.3%.31 Li et al. prepared an Sb2S3 nanocrystal coated TiO2 dendritic structure for a hybrid solar cell and achieved a PCE of 1.56%.32,33

On the other hand, antimony chalcogenides, such as Sb2S3 and Sb2Se3 with an orthorhombic structure, have inherent anisotropic crystal structures.2 However, only a few studies have reported on the application of those inherent anisotropic properties of the Sb2S3 crystal structure for photovoltaics. The Sb2S3 crystal is constructed by 1D ribbon like (Sb4S6)n chains, and it has been proved that the carrier can easily transport along the ribbon but is hard to jump between ribbons.2,16 In order to apply the anisotropic properties of Sb2S3 crystal structure and advantages of 1D nanostructure for improvement of electron transport in solar cells, we designed a comb-shaped Sb2S3 nanorod array on ZnO nanofibers as a light absorber for the photovoltaic device. Electrospun ZnO nanofiber scaffold was coated on compact ZnO, and Sb2S3 nanorod was grown on ZnO nanofibers by hydrothermal method.

Preparation of glass-FTO/compact-ZnO (c-ZnO)/ZnO nanofibers-ZnS, the chemical bath deposition of Sb2S3 seed layer, electron spinning method for preparation of ZnO nanofiber, hydrothermal method for the growth of Sb2S3 nanorod arrays, and the preparation procedures of thin-film solar cells are described in the supplementary material.34,35–37 

Measurement of x-ray diffraction (XRD) patterns, observation of scanning electron microscope (SEM) images and energy dispersive spectroscopy (EDS) mappings, measurement of x-ray photoelectron spectra (XPS), UV-visible absorption spectra, and photoemission yield, and measurement of current density–voltage (JV) curves and external quantum efficiency (EQE) were carried out using previously reported apparatuses.34 

The optical image and SEM images of the pristine ZnO nanofibers and the Sb2S3 seeds deposited on the ZnO nanofibers are shown in Figs. 1(a)1(d). Rod like parts of the Sb2S3 seeds were observed in Fig. 1(d). Sb2S3 nanorods were grown on ZnO nanofibers in the presence of 8 mg ml-1 polyvinylpyrrolidone (PVP, Molecular Weight = 1300000, 0.4 g PVP in 50 ml precursor solution), and the composite structure was confirmed in SEM images in Figs. 1(e) and 1(f).

FIG. 1.

Optical photo images of (a) ZnO nanofibers, (b) Sb2S3 seed layer coated on ZnO nanofibers, and SEM images of (c) ZnO nanofibers, (d) Sb2S3 seed layer coated on ZnO nanofibers, and (e) and (f) Sb2S3 nanorod arrays grown on ZnO nanofibers.

FIG. 1.

Optical photo images of (a) ZnO nanofibers, (b) Sb2S3 seed layer coated on ZnO nanofibers, and SEM images of (c) ZnO nanofibers, (d) Sb2S3 seed layer coated on ZnO nanofibers, and (e) and (f) Sb2S3 nanorod arrays grown on ZnO nanofibers.

Close modal

Sb2S3 compounds with an orthorhombic structure have an inherent anisotropic crystal structure. The Sb2S3 crystal is formed by (Sb4S6)n chains, and each chain is combined together by Van der Waals force as shown in Fig. 2(a), causing the carrier easier to transport through the chains rather than hopping through the inter-chains. Surfaces of Sb2S3, which are parallel to the [001] direction, such as (100), (010), (110), and (120) surfaces, have no dangling bonds and have lower formation energies than (hk1) surfaces. It has also been confirmed by a computational study that, as long as the ribbons are suitably oriented, the grain boundaries will be terminated by the intrinsically benign surfaces [for example, (100), (010), (110), and (120) planes], and the recombination loss would be minimized,2 as shown in Fig. 2(b). This effective carrier transport ability along the (Sb4S6)n ribbon and suppressed recombination nature makes the 1D Sb2S3 nanostructure an excellent light absorber, and it is to offer photo response and device performance.9 

FIG. 2.

Schematic diagram of (a) four Sb2S3 prime cells stacking along [001] direction and (b) comb-shaped Sb2S3 nanorod arrays on ZnO nanofiber. Note that all the atoms at the edge of these ribbons are saturated and introduce no recombination loss at the grain boundaries once they are oriented vertically onto the substrates.

FIG. 2.

Schematic diagram of (a) four Sb2S3 prime cells stacking along [001] direction and (b) comb-shaped Sb2S3 nanorod arrays on ZnO nanofiber. Note that all the atoms at the edge of these ribbons are saturated and introduce no recombination loss at the grain boundaries once they are oriented vertically onto the substrates.

Close modal

PVP is used for the preparation of metal-based nanomaterials since the oxygen atom in PVP can coordinate with metal and act as a capping ligand, which will restrict the growth of nanocrystals in a certain direction. In this context, different amount (0, 1, 2, 4, and 8 mg ml−1) of PVP was added to the precursor solution for the hydrothermal process, and the morphological changes of Sb2S3 nanorod arrays on ZnO nanofiber composites were observed by SEM as indicated in Figs. 3(a)3(e), and their corresponding optical photos are shown in Fig. 3(f). It can be observed that the thickness of Sb2S3 nanorods covered on the ZnO nanofibers is ∼1 µm, which is several times larger than the diameter of ZnO nanofiber, and the ZnO nanofiber scaffold could not be observed since it has been hidden by Sb2S3 nanorods. With the increasing amount of PVP used in the hydrothermal synthesis, the aspect ratios of the ZnO nanofiber increased. When the PVP concentration used was 2 and 4 mg ml−1, comb-shaped structures were obtained. As shown in Fig. 3(g), Sb2S3 nanorods grown on ZnO nanofibers have two predominant orientations, while these two orientations are vertical to each other. This unique structure disappears when the PVP amount was increased to 8 mg ml−1. When the PVP concentration was 8 mg ml−1, the Sb2S3 nanorods covered densely on the ZnO nanofibers and it is hard to tell the predominant orientation of Sb2S3 nanorods.

FIG. 3.

SEM images of comb-shaped Sb2S3 nanorod arrays grown on ZnO nanofibers prepared with (a) 0 g, (b) 0.05 g, (c) 0.1 g, (d) 0.2 g, and (e) 0.4 g of PVP in 50 ml of the precursor solution during the hydrothermal process at (i) lower, (ii) higher magnifications, (f) optical image of the above five samples, and (g) SEM images of Sb2S3 nanorod arrays grown on ZnO nanofibers prepared with 0.1 g of PVP through hydrothermal reaction for 2 h in two different positions where Sb2S3 nanorod arrays are (i) fully and (ii) not fully covered ZnO nanofibers.

FIG. 3.

SEM images of comb-shaped Sb2S3 nanorod arrays grown on ZnO nanofibers prepared with (a) 0 g, (b) 0.05 g, (c) 0.1 g, (d) 0.2 g, and (e) 0.4 g of PVP in 50 ml of the precursor solution during the hydrothermal process at (i) lower, (ii) higher magnifications, (f) optical image of the above five samples, and (g) SEM images of Sb2S3 nanorod arrays grown on ZnO nanofibers prepared with 0.1 g of PVP through hydrothermal reaction for 2 h in two different positions where Sb2S3 nanorod arrays are (i) fully and (ii) not fully covered ZnO nanofibers.

Close modal

The mechanisms of the formation of such a hierarchical structure can be supposed as follows: As shown in Fig. 4(a), the precursor of S and Sb started to react with each other and formed Sb2S3 nanocrystals randomly on the seed layer, which has already been coated on ZnO nanofibers. Since the (hk0) faces have no dangling bonds as mentioned above and have smaller formation energy than (hk1), Sb2S3 nanocrystals will naturally grow to be rod like structures. It also indicates that the Sb2S3 nanorod is less likely to grow parallel along the central axis of one nanofiber [Fig. 4(c-ii)] but grow vertically to the central axis of one nanofiber [Fig. 4(c-i)] because of no dangling bonds in (hk0) faces and no covalent bond between Sb2S3 and ZnO.

FIG. 4.

Schematic diagram of (a) growth mechanisms of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers, (b) cross-sectional view of one ZnO nanofiber attached with comb-shaped Sb2S3 nanorod arrays, in which the Sb2S3 nanorods can only grow along x axis instead of z axis because of steric hindrance, (c-i) Sb2S3 nanorod vertically grown on ZnO nanofiber, (c-ii) parallel growth of Sb2S3 nanorod on ZnO nanofiber, and (d) comb-shaped Sb2S3 nanorod arrays with different diameters on ZnO nanofibers with different concentration of PVP in the precursor solution.

FIG. 4.

Schematic diagram of (a) growth mechanisms of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers, (b) cross-sectional view of one ZnO nanofiber attached with comb-shaped Sb2S3 nanorod arrays, in which the Sb2S3 nanorods can only grow along x axis instead of z axis because of steric hindrance, (c-i) Sb2S3 nanorod vertically grown on ZnO nanofiber, (c-ii) parallel growth of Sb2S3 nanorod on ZnO nanofiber, and (d) comb-shaped Sb2S3 nanorod arrays with different diameters on ZnO nanofibers with different concentration of PVP in the precursor solution.

Close modal

Initially formed Sb2S3 nanorods have a predominant growth orientation along with the seeds coated on the nanofibers as shown in Fig. 5(a). In some areas, the Sb2S3 nanorods can grow along only with some certain direction because other directions are blocked by other ZnO nanofibers (steric hindrance) as shown in Fig. 4(b). However, once some Sb2S3 nanorods are formed on nanofiber, subsequent Sb2S3 nanorods are preferred to grow nearby because of lower formation energy and Ostwald Ripening effect38 and preferred to form locally aligned orientation, which can be confirmed by the SEM image of the intermediate state of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers as shown in Fig. 3(g). The hydrothermal reaction was terminated at a different time, and the SEM images of those samples are shown in Figs. 5(b)5(f). It can be concluded that the nanorods start to form at 30 min and the predominant growth directions are limited by the seeds layer, and the nanorods are preferred to grow aligned next to each other. EDS mapping and line scanning were characterized as indicated in Fig. S2 as proof of the composite nanostructure.

FIG. 5.

SEM images of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers obtained with different hydrothermal reaction times for (a) 0 min, (b) 0.5 h, (c) 1 h, (d) 1.5 h, (e) 2.5 h, and (f) 4 h.

FIG. 5.

SEM images of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers obtained with different hydrothermal reaction times for (a) 0 min, (b) 0.5 h, (c) 1 h, (d) 1.5 h, (e) 2.5 h, and (f) 4 h.

Close modal

As presented in Fig. 4(d), when the PVP concentration in precursor solution was low, obtained Sb2S3 nanorods were large in both radius and length, and it can not be observed hierarchical nanostructure. With the increment of PVP concentration, the Sb2S3 nanorods were smaller and grew vertically on the nanofiber to the central axis of that nanofiber. This phenomenon can be ascribed to the steric hindrance, which prohibits the growth caused by the capping effect from PVP molecules,39 as presented in Fig. S3.

XRD patterns were characterized to figure out the predominant orientation of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers as shown in Fig. 6.

FIG. 6.

(a) XRD patterns and (b) texture coefficient of (hk0) faces, and (c) texture coefficient of (hk1) faces of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers prepared with 0, 0.05, 0.1, 0.2, and 0.4 g of PVP in 50 ml of the precursor solution in the hydrothermal reaction.

FIG. 6.

(a) XRD patterns and (b) texture coefficient of (hk0) faces, and (c) texture coefficient of (hk1) faces of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers prepared with 0, 0.05, 0.1, 0.2, and 0.4 g of PVP in 50 ml of the precursor solution in the hydrothermal reaction.

Close modal

The texture coefficient of each (hk0) and (hk1) face is calculated by the following equation:

TChkl=I(hkl)/I0(hkl)1NNI(hkl)/I0(hkl).
(1)

Those values of the samples prepared with various amounts of PVP used in the hydrothermal reaction are plotted in Figs. 6(b) and 6(c). In terms of (hk0) faces, the samples prepared with 0.1 g PVP have the highest texture coefficient, while increasing or decreasing the PVP concentration will induce a lower texture coefficient of (hk0) faces. Since the (hk0) faces are vertical to (hk1) faces, the texture coefficients of (hk1) peaks have an opposite trend.

This can be explained by the growth mechanisms of such comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers as mentioned above in Fig. 4. Since the ZnO nanofibers are likely to stack on each other and cause steric hindrance for the growth of Sb2S3 nanorods as shown in Fig. 4(b), Sb2S3 nanorods predominantly grow along the x-axis instead of the z-axis. Moreover, when the PVP concentration was too low, the nanocrystals were larger and ZnO nanofibers were submerged under Sb2S3 nanorods, while the size of nanocrystals was smaller and diminish the influence of steric hindrance.

It can be confirmed that PVP concentration affects the nanocrystal size of Sb2S3 by the values of FWHM of some certain peaks of the XRD patterns as shown in Figs. S4 and S5. The (330) peaks of each sample were fitted by the gaussian method and compared in Fig. S6a. The respective grain sizes were estimated by using Scherrer’s equation,

D=Kλβcosθ,
(2)

where D is the mean size of the ordered (crystalline) domains, which may be smaller or equal to the grain size and may be smaller or equal to the particle size; K is a dimensionless shape factor, with a value close to unity. λ is the x-ray wavelength, β is FWHM in radians, and θ is the Bragg angle. The calculated crystalline sizes are plotted in Fig. S6b. It can be found that the crystalline sizes of Sb2S3 nanocrystals decreased as the PVP concentration used in the precursor solution increased. A similar trend also exists in the (130) peak as depicted in Figs. S6c and S6d.

Figure 7(d) indicates that the band energy level diagram of five samples of Sb2S3 nanorod arrays on ZnO nanofibers prepared with different concentrations of PVP and was estimated from the absorption edges of their UV-visible absorption spectra [Fig. 7(a)], the Tauc’s plots [Fig. 7(b)], and the plots of the emission yield vs photoenergy profiles [Fig. 7(c)].

FIG. 7.

(a) Absorption spectra, (b) Tauc’s plots of (αhv)1/2 vs photo energy, (c) emission yield vs photo energy profiles, (d) band energy level diagram of Sb2S3 nanorod arrays grown on ZnO nanofibers with different amounts of PVP in the precursor solution, and (e) JV curves and (f) dark J-V curves of solar cells composed of glass-FTO/c-ZnO/ZnO nanofibers−ZnS/Sb2S3 nanorod arrays/P3HT/MoOx/Ag, in which the Sb2S3 nanorod arrays grown on ZnO nanofibers were prepared with different amount of PVP in the precursor solution.

FIG. 7.

(a) Absorption spectra, (b) Tauc’s plots of (αhv)1/2 vs photo energy, (c) emission yield vs photo energy profiles, (d) band energy level diagram of Sb2S3 nanorod arrays grown on ZnO nanofibers with different amounts of PVP in the precursor solution, and (e) JV curves and (f) dark J-V curves of solar cells composed of glass-FTO/c-ZnO/ZnO nanofibers−ZnS/Sb2S3 nanorod arrays/P3HT/MoOx/Ag, in which the Sb2S3 nanorod arrays grown on ZnO nanofibers were prepared with different amount of PVP in the precursor solution.

Close modal

XPS of the above five samples of Sb2S3 nanorod arrays on ZnO nanofibers prepared with different concentrations of PVP was characterized as shown in Fig. S7. The spin–orbit coupled doublet Sb 3d core level was split into 3d5/2 and 3d3/2, and the separation of the 3d doublet by 9.3 eV can be attributed to the charge state of Sb3+.40 The Sb 3d peak in the XPS spectrum of the Sb2Se3 crystals can be deconvoluted into several peaks. It is worth noticing that the FWHM of Sb 3d peaks are fixed to be 0.86 eV for all spectra in the fitting process. It has already been confirmed that the signal of the Sb 3d5/2 peak and Sb3/2 is composed of Sb–S bonds from Sb2S3 and Sb–O from Sb2O3 defects.41 It is also confirmed that a non-negligible O 1s peak was observed in the Sb3d orbitals of the samples, which was caused by the existence of the –OH group.42–44 The Sb–O defects were caused by non-radiative recombination, which will be resulted in PCE loss when used in photovoltaic devices. The spectra of Sb 3d of five tested samples show clear Sb–O defects. The lower Sb–S to Sb–O atomic ratio indicates more defects; hence, the comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers prepared with 0.1 g PVP had fewer defects than others as shown in Fig. S7f.

Thin film solar cells composed of glass-FTO/c-ZnO/ZnO nanofibers−ZnS/Sb2S3 nanorod arrays/P3HT/MoOx/Ag were prepared with the above five samples of Sb2S3 nanorod arrays on ZnO nanofibers obtained with different concentrations of PVP, and the JV curves of the best performed solar cells were indicated in Fig. 7(e). Short circuit current density (Jsc), open circuit voltage (Voc), fill factor (FF), PCE, series resistance (Rs), and shunt resistance (Rsh) of the champion device were summarized in Table I.

TABLE I.

Photovoltaic characteristics of solar cells composed of glass-FTO/c-ZnO/ZnO nanofibers−ZnS/Sb2S3 nanorod arrays/P3HT/MoOx/Ag (champion device).

PVP (g)Jsc (mA cm−2)Voc (V)FFPCE (%)Rs (Ω)Rsh (Ω)
2.767 0.203 0.317 0.178 1148.0 2153.2 
0.05 5.355 0.217 0.285 0.331 781.5 3332.9 
0.1 6.293 0.309 0.340 0.662 646.3 3513.8 
0.2 4.322 0.305 0.473 0.473 747.0 3473.2 
0.4 3.929 0.271 0.379 0.403 701.1 3403.5 
PVP (g)Jsc (mA cm−2)Voc (V)FFPCE (%)Rs (Ω)Rsh (Ω)
2.767 0.203 0.317 0.178 1148.0 2153.2 
0.05 5.355 0.217 0.285 0.331 781.5 3332.9 
0.1 6.293 0.309 0.340 0.662 646.3 3513.8 
0.2 4.322 0.305 0.473 0.473 747.0 3473.2 
0.4 3.929 0.271 0.379 0.403 701.1 3403.5 

It can be found that the lowest PCE was attained when the Sb2S3 nanorod arrays on ZnO nanofibers prepared without PVP were used. This is mainly caused by the high leakage current as shown in the dark J-V curves in Fig. 7(f). As the SEM image and schematic diagram shown in Figs. 3(a) and 4(d), the radius of Sb2S3 nanorods is larger than that of ZnO nanofiber, and the gap between the nanorods is quite large, which may be induced by the direct contact between P3HT and ZnO nanofiber. On the other hand, the device using comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers prepared with 0.1 g PVP showed the highest PCE, which may be attributed to its relatively high carrier transport ability and relatively fewer Sb–O defects (Fig. S7f).

The champion device of using the comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers as a light absorbing layer was also compared with devices using planar Sb2S3 layer and randomly stacked Sb2S3 nanorods as a light absorber, and the JV characterization of the champion devices are shown in Figs. 8(a) and 8(b) and Table II.

FIG. 8.

(a) JV curves, (b) dark JV curves, (c) EQE profiles, and (d) absorption spectra of the light absorber of the solar cells using randomly stacked Sb2S3 nanorods, comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers, and planar Sb2S3 as the light absorber.

FIG. 8.

(a) JV curves, (b) dark JV curves, (c) EQE profiles, and (d) absorption spectra of the light absorber of the solar cells using randomly stacked Sb2S3 nanorods, comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers, and planar Sb2S3 as the light absorber.

Close modal
TABLE II.

Photovoltaic characteristics of solar cells with three different light absorbers (champion device).

Light absorberJsc (mA cm−2)Voc (V)FFPCE (%)RS (Ω)RSH (Ω)
Randomly stacked Sb2S3 nanorods 3.315 0.302 0.321 0.374 1075.4 1784.5 
Comb-shaped Sb2S3 nanorod arrays 6.293 0.309 0.340 0.662 646.28 3513.8 
Planar Sb2S3 layer 1.952 0.095 0.372 0.060 1132.4 1341.2 
Light absorberJsc (mA cm−2)Voc (V)FFPCE (%)RS (Ω)RSH (Ω)
Randomly stacked Sb2S3 nanorods 3.315 0.302 0.321 0.374 1075.4 1784.5 
Comb-shaped Sb2S3 nanorod arrays 6.293 0.309 0.340 0.662 646.28 3513.8 
Planar Sb2S3 layer 1.952 0.095 0.372 0.060 1132.4 1341.2 

It was found that the device using planar Sb2S3 as a light absorbing layer showed the lowest PCE. It can also be found that the device using the comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers exhibit better performance than planar absorber as well as randomly stacked Sb2S3 nanorods. From the EQE curves as shown in Fig. 8(c), the randomly stacked nanorods-based device showed higher EQE than that of the comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers in the region from 310 to 330 nm, while the comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers-based device showed much higher EQE from 330 to 750 nm. The absorption spectra of the above three light absorbing layers-based solar cells were also compared as depicted in Fig. 8(d), and the cross-sectional SEM images of those tested light absorbers-based solar cells were shown in Fig. S8. It is clear that the profiles of the EQE curves in Fig. 8(c) are reflected by the absorbance in Fig. 8(d), which was directly reflected by the total thickness of the layers of each component. The higher EQE of the device using Sb2S3 nanorods on ZnO nanofibers as light absorbers might be ascribed to the fast carrier transport from Sb2S3 to ZnO.

An equivalent circuit model of the solar cells was introduced to evaluate the improvement of electron transport efficiency by Eq. (3),

J=J0expVRSJnVth+GVJL.
(3)

Vth was obtained from thermal energy kBT/e, where kB is the Boltzmann constant, T is the absolute temperature, and e is the elementary charge, which is generally considered to be 0.0257 V at 298 K. JL is the photocurrent density, which can be directly approximated as the current density of the device when voltage is 0 V (JL = Jsc). RS is the series resistance and G (=Rsh−1) is the parallel conductance. J0 and n are the reversed saturation current density at the dark state and ideality factor of the diodes, respectively. These parameters can be solved separately by transforming Eq. (3) into the three following Eqs. (4)(6):

dJdV=1nVth×J0expVRsJnVth+G,
(4)
dVdJ=Rs+nVthJ+JLGV,
(5)
lnJ+JLGV=1nVth×VRsJ+lnJ0.
(6)

The parallel conductance G calculated by Eq. (4) is shown in Fig. 9(a), and the randomly stacked Sb2S3 nanorods device has the smallest parallel conductance of 8.40 mS cm−2 compared to that of 10.85 and 11.23 mS cm−2 for Sb2S3 nanorod decorated ZnO nanofiber solar cell and planar Sb2S3 device, respectively.

FIG. 9.

Calculation of characteristic parameters in the equivalent circuit: (a) shunt conductance, G; (b) series, Rs, and ideality factor, n; and (c) dark state saturation current, J0.

FIG. 9.

Calculation of characteristic parameters in the equivalent circuit: (a) shunt conductance, G; (b) series, Rs, and ideality factor, n; and (c) dark state saturation current, J0.

Close modal

The series resistance Rs calculated by Eq. (5) is shown in Fig. 9(b) with 8.13, 15.01, and 13.46 Ω cm−2 for Sb2S3 nanorod decorated ZnO nanofiber solar cell, randomly stacked Sb2S3 nanorods and planar Sb2S3 devices, respectively. It was found that the Sb2S3 nanorod decorated ZnO nanofiber based device showed the smallest Rs, implying the improvement in charge extraction capability. The value of ideality factor n is obtained from Eq. (5) by the slope of the dashed line in Fig. 9(b), reflects the recombination in the diode, with n being 2.84, 3.83, and 2.01 for Sb2S3 nanorod decorated ZnO nanofiber solar cell, randomly stacked Sb2S3 nanorods, and planar Sb2S3 devices, respectively. The smaller n indicates the more suppression of recombination at the interface. In other words, there is still quite a lot of recombination occurring in the interface for Sb2S3 nanorods on ZnO nanofibers-based solar cells, which might be caused by the insufficient contact between each component. The maximum Voc could be predicted by Shockley–Queisser limit model45 as can be expressed by Eq. (7),

VocSQ=VthlnJscSQJ0BB+1,
(7)

where J0BB is the reverse saturation current density calculated by considering merely the black body radiation of the solar cell at room temperature. The trapping and re-emission process of non-equilibrium carriers by deep-level defects in the device can cause severe recombination current and boost the J0, which can be considered to be the loss in Voc.46 The values of J0 of each device were also calculated by Eq. (6) as indicated in Fig. 9(c). Calculated J0 of Sb2S3 nanorods ZnO nanofibers-based solar cell was 2.47 × 10−5 mA cm−2, which is smaller than that of the other two devices, one order smaller than that of the planar device, indicating that the trap-mediated recombination was effectively suppressed and resulted in a much larger Voc.

Low PCE of a such device using the comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers as light absorbers may be brought by the insufficient contact between ZnO nanofibers and compact ZnO layer after the growth of Sb2S3 nanorods as shown in Fig. S9. Some nanofibers as marked in the red dashed circle cannot contact directly with the compact ZnO layer. These parts were occupied by Sb2S3 nanorods as confirmed by the cross-sectional SEM image and EDS mapping as shown in Fig. S10. It is difficult for free carriers generated from these parts to travel through a long path and overcome a huge barrier to the FTO electrode. Those parts also blocked contact with other parts, like P3HT, which may also reduce the quantum efficiency.

Preparation for comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers through hydrothermal growth of Sb2S3 nanorods on ZnO nanofiber was proposed. The size of the diameter of the nanorod and the extent of formation of comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers were controlled by changing the concentration of PVP as the additive in the precursor solution. Spectral properties of Sb2S3 nanorod arrays on ZnO nanofibers were characterized, and the photovoltaic characteristics of thin-film solar cells were evaluated with the device structure of glass-FTO/c-ZnO/ZnO nanofibers−ZnS/Sb2S3 nanorod arrays/P3HT/MoOx/Ag. It was found that the rectification ratio and photocurrent generation efficiency of the comb-shaped Sb2S3 nanorod arrays on ZnO nanofibers were improved as compared with those of randomly stacked Sb2S3 nanorods. Smaller series resistance and ideality factor of the comb-shaped Sb2S3 nanorod arrays than those of the randomly stacked ones also indicated superior charge extraction properties and suppressed recombination at the interface.

See the supplementary material for experimental detail and the corresponding other experimental data.

This work was supported by the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan as the MEXT KAKENHI (Grant Number: 17H03536).

The authors have no conflicts to disclose.

B. Zhou: Conceptualization (lead); Data curation (equal); Writing – original draft (equal). T. Sagawa: Conceptualization (supporting); Supervision (lead); Writing – review & editing (lead).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
M. Y.
Versavel
and
J. A.
Haber
, “
Structural and optical properties of amorphous and crystalline antimony sulfide thin-films
,”
Thin Solid Films
515
(
18
),
7171
7176
(
2007
).
2.
R.
Kondrotas
,
C.
Chen
, and
J.
Tang
, “
Sb2S3 solar cells
,”
Joule
2
(
5
),
857
878
(
2018
).
3.
J.
Han
,
X.
Pu
,
H.
Zhou
,
Q.
Cao
,
S.
Wang
,
Z.
He
,
B.
Gao
,
T.
Li
,
J.
Zhao
, and
X.
Li
, “
Synergistic effect through the introduction of inorganic zinc halides at the interface of TiO2 and Sb2S3 for high-performance Sb2S3 planar thin-film solar cells
,”
ACS Appl. Mater. Interfaces
12
(
39
),
44297
44306
(
2020
).
4.
L.
Wang
,
D.
Li
,
K.
Li
,
C.
Chen
,
H.
Deng
,
L.
Gao
,
Y.
Zhao
,
F.
Jiang
,
L.
Li
,
F.
Huang
,
Y.
He
,
H.
Song
,
G.
Niu
, and
J.
Tang
, “
Stable 6%-efficient Sb2Se3 solar cells with a ZnO buffer layer
,”
Nat. Energy
2
(
4
),
1
9
(
2017
).
5.
X.
Wen
,
C.
Chen
,
S.
Lu
,
K.
Li
,
R.
Kondrotas
,
Y.
Zhao
,
W.
Chen
,
L.
Gao
,
C.
Wang
, and
J.
Zhang
, “
Vapor transport deposition of antimony selenide thin film solar cells with 7.6% efficiency
,”
Nat. Commun.
9
(
1
),
2179
(
2018
).
6.
H.
Deng
,
Y.
Zeng
,
M.
Ishaq
,
S.
Yuan
,
H.
Zhang
,
X.
Yang
,
M.
Hou
,
U.
Farooq
,
J.
Huang
,
K.
Sun
,
R.
Webster
,
H.
Wu
,
Z.
Chen
,
F.
Yi
,
H.
Song
,
X.
Hao
, and
J.
Tang
, “
Quasiepitaxy strategy for efficient full‐inorganic Sb2S3 solar cells
,”
Adv. Funct. Mater.
29
(
31
),
1901720
(
2019
).
7.
R.
Nie
,
H. S.
Yun
,
M. J.
Paik
,
A.
Mehta
,
B. W.
Park
,
Y. C.
Choi
, and
S. I.
Seok
, “
Efficient solar cells based on light‐harvesting antimony sulfoiodide
,”
Adv. Energy Mater.
8
(
7
),
1701901
(
2018
).
8.
R.
Tang
,
X.
Wang
,
C.
Jiang
,
S.
Li
,
W.
Liu
,
H.
Ju
,
S.
Yang
,
C.
Zhu
, and
T.
Chen
, “
n-type doping of Sb2S3 light-harvesting films enabling high-efficiency planar heterojunction solar cells
,”
ACS Appl. Mater. Interfaces
10
(
36
),
30314
30321
(
2018
).
9.
Z.
Li
,
X.
Liang
,
G.
Li
,
H.
Liu
,
H.
Zhang
,
J.
Guo
,
J.
Chen
,
K.
Shen
,
X.
San
,
W.
Yu
,
R.
Schropp
, and
Y.
Mai
, “
9.2%-efficient core-shell structured antimony selenide nanorod array solar cells
,”
Nat. Commun.
10
(
1
),
125
(
2019
).
10.
Y.
Zeng
,
K.
Sun
,
J.
Huang
,
M. P.
Nielsen
,
F.
Ji
,
C.
Sha
,
S.
Yuan
,
X.
Zhang
,
C.
Yan
,
X.
Liu
,
H.
Deng
,
Y.
Lai
,
J.
Seidel
,
N.
Ekins-Daukes
,
F.
Liu
,
H.
Song
,
M.
Green
, and
X.
Hao
, “
Quasi-vertically-orientated antimony sulfide inorganic thin-film solar cells achieved by vapor transport deposition
,”
ACS Appl. Mater. Interfaces
12
(
20
),
22825
22834
(
2020
).
11.
S.-J.
Lee
,
S.-J.
Sung
,
K.-J.
Yang
,
J.-K.
Kang
,
J. Y.
Kim
,
Y. S.
Do
, and
D.-H.
Kim
, “
Approach to transparent photovoltaics based on wide band gap Sb2S3 absorber layers and optics-based device optimization
,”
ACS Appl. Energy Mater.
3
(
12
),
12644
12651
(
2020
).
12.
X.
Jin
,
Y.
Fang
,
T.
Salim
,
M.
Feng
,
S.
Hadke
,
S. W.
Leow
,
T. C.
Sum
, and
L. H.
Wong
, “
In situ growth of [hk1]‐oriented Sb2S3 for solution‐processed planar heterojunction solar cell with 6.4% efficiency
,”
Adv. Funct. Mater.
30
(
35
),
2002887
(
2020
).
13.
W.
Lin
,
W.-T.
Guo
,
L.
Yao
,
J.
Li
,
L.
Lin
,
J.-M.
Zhang
,
S.
Chen
, and
G.
Chen
, “
Zn(O,S) buffer layer for in situ hydrothermal Sb2S3 planar solar cells
,”
ACS Appl. Mater. Interfaces
13
(
38
),
45726
45735
(
2021
).
14.
W.
Wang
,
X.
Wang
,
G.
Chen
,
L.
Yao
,
X.
Huang
,
T.
Chen
,
C.
Zhu
,
S.
Chen
,
Z.
Huang
, and
Y.
Zhang
, “
Over 6% certified Sb2(S,Se)3 solar cells fabricated via in situ hydrothermal growth and postselenization
,”
Adv. Electron. Mater.
5
(
2
),
1800683
(
2018
).
15.
Y.
Zhao
,
S.
Wang
,
C.
Jiang
,
C.
Li
,
P.
Xiao
,
R.
Tang
,
J.
Gong
,
G.
Chen
,
T.
Chen
, and
J.
Li
, “
Regulating energy band alignment via alkaline metal fluoride assisted solution post‐treatment enabling Sb2(S,Se)3 solar cells with 10.7% efficiency
,”
Adv. Energy Mater.
12
(
1
),
2103015
(
2022
).
16.
U. A.
Shah
,
S.
Chen
,
G. M. G.
Khalaf
,
Z.
Jin
, and
H.
Song
, “
Wide bandgap Sb2S3 solar cells
,”
Adv. Funct. Mater.
31
(
27
),
2100265
(
2021
).
17.
J. A.
Christians
,
D. T.
Leighton
, and
P. V.
Kamat
, “
Rate limiting interfacial hole transfer in Sb2S3 solid-state solar cells
,”
Energy Environ. Sci.
7
(
3
),
1148
1158
(
2014
).
18.
M.
Batmunkh
,
T. J.
Macdonald
,
C. J.
Shearer
,
M.
Bat‐Erdene
,
Y.
Wang
,
M. J.
Biggs
,
I. P.
Parkin
,
T.
Nann
, and
J. G.
Shapter
, “
Carbon nanotubes in TiO2 nanofiber photoelectrodes for high‐performance perovskite solar cells
,”
Adv. Sci.
4
(
4
),
1600504
(
2017
).
19.
L. E.
Greene
,
M.
Law
,
B. D.
Yuhas
, and
P.
Yang
, “
ZnO−TiO2 core−shell nanorod/P3HT solar cells
,”
J. Phys. Chem. C
111
(
50
),
18451
18456
(
2007
).
20.
Y.-J.
Lee
,
D. S.
Ruby
,
D. W.
Peters
,
B. B.
McKenzie
, and
J. W. P.
Hsu
, “
ZnO nanostructures as efficient antireflection layers in solar cells
,”
Nano Lett.
8
(
5
),
1501
1505
(
2008
).
21.
Y.
Hames
,
Z.
Alpaslan
,
A.
Kösemen
,
S. E.
San
, and
Y.
Yerli
, “
Electrochemically grown ZnO nanorods for hybrid solar cell applications
,”
Solar Energy
84
(
3
),
426
431
(
2010
).
22.
D. Y.
Son
,
J. H.
Im
,
H. S.
Kim
, and
N. G.
Park
, “
11% efficient perovskite solar cell based on ZnO nanorods: An effective charge collection system
,”
J. Phys. Chem. C
118
(
30
),
16567
16573
(
2014
).
23.
E.
Galoppini
,
J.
Rochford
,
H.
Chen
,
G.
Saraf
,
Y.
Lu
,
A.
Hagfeldt
, and
G.
Boschloo
, “
Fast electron transport in metal organic vapor deposition grown dye-sensitized ZnO nanorod solar cells
,”
J. Phys. Chem. B
110
(
33
),
16159
16161
(
2006
).
24.
K.
Mahmood
,
A.
Khalid
,
S. W.
Ahmad
, and
M. T.
Mehran
, “
Indium-doped ZnO mesoporous nanofibers as efficient electron transporting materials for perovskite solar cells
,”
Surf. Coat. Technol.
352
,
231
237
(
2018
).
25.
G. S.
Han
,
H. S.
Chung
,
D. H.
Kim
,
B. J.
Kim
,
J.-W.
Lee
,
N.-G.
Park
,
I. S.
Cho
,
J.-K.
Lee
,
S.
Lee
, and
H. S.
Jung
, “
Epitaxial 1D electron transport layers for high-performance perovskite solar cells
,”
Nanoscale
7
(
37
),
15284
15290
(
2015
).
26.
L.
Yang
and
W. W.-F.
Leung
, “
Application of a bilayer TiO2 nanofiber photoanode for optimization of dye‐sensitized solar cells
,”
Adv. Mater.
25
(
12
),
1792
1795
(
2013
).
27.
F. J.
Ramos
,
M.
Oliva-Ramirez
,
M. K.
Nazeeruddin
,
M.
Grätzel
,
A. R.
González-Elipe
, and
S.
Ahmad
, “
Nanocolumnar 1-dimensional TiO2 photoanodes deposited by PVD-OAD for perovskite solar cell fabrication
,”
J. Mater. Chem. A
3
(
25
),
13291
13298
(
2015
).
28.
L.
Yang
and
W. W.-F.
Leung
, “
Application of a bilayer TiO2 nanofiber photoanode for optimization of dye‐sensitized solar cells
,”
Adv. Mater.
23
(
39
),
4559
4562
(
2011
).
29.
C.
Ying
,
F.
Guo
,
Z.
Wu
,
K.
Lv
, and
C.
Shi
, “
Influence of surface modifier molecular structures on the photovoltaic performance of Sb2S3‐sensitized TiO2 nanorod array solar cells
,”
Energy Technol.
8
(
6
),
1901368
(
2020
).
30.
Z.
Sun
,
Z.
Peng
,
Z.
Liu
,
J.
Chen
,
W.
Li
,
W.
Qiu
, and
J.
Chen
, “
Band energy modulation on Cu-doped Sb2S3-based photoelectrodes for charge generation and transfer property of quantum dot–sensitized solar cells
,”
J. Nanoparticle Res.
22
(
9
),
1
9
(
2020
).
31.
R.
Parize
,
A.
Katerski
,
I.
Gromyko
,
L.
Rapenne
,
H.
Roussel
,
E.
Kärber
,
E.
Appert
,
M.
Krunks
, and
V.
Consonni
, “
ZnO/TiO2/Sb2S3 core–shell nanowire heterostructure for extremely thin absorber solar cells
,”
J. Phys. Chem. C
121
(
18
),
9672
9680
(
2017
).
32.
Y.
Li
,
Y.
Wei
,
K.
Feng
,
Y.
Hao
,
J.
Pei
, and
B.
Sun
, “
Preparation of Sb2S3 nanocrystals modified TiO2 dendritic structure with nanotubes for hybrid solar cell
,”
Mater. Res. Express
5
(
6
),
065903
(
2018
).
33.
Y.
Li
,
Y.
Wei
,
K.
Feng
,
Y.
Hao
,
J.
Pei
,
Y.
Zhang
, and
B.
Sun
, “
Introduction of PCPDTBT in P3HT: Spiro-OMeTAD blending system for solid-state hybrid solar cells with dendritic TiO2/Sb2S3 nanorods composite film
,”
J. Solid State Chem.
276
,
278
284
(
2019
).
34.
B.
Zhou
,
T.
Hayashi
,
K.
Hachiya
, and
T.
Sagawa
, “
Preparation of Sb2S3 nanorod arrays by hydrothermal method as light absorbing layer for Sb2S3-based solar cells
,”
Thin Solid Films
757
,
139389
(
2022
).
35.
R.
Chen
,
J.
Cao
,
Y.
Duan
,
Y.
Hui
,
T. T.
Chuong
,
D.
Ou
,
F.
Han
,
F.
Cheng
,
X.
Huang
,
B.
Wu
, and
N.
Zheng
, “
High-efficiency, hysteresis-less, UV-stable perovskite solar cells with cascade ZnO–ZnS electron transport layer
,”
J. Am. Chem. Soc.
141
(
1
),
541
547
(
2019
).
36.
S.
Messina
,
M. T. S.
Nair
, and
P. K.
Nair
, “
All-chemically deposited solar cells with antimony sulfide-selenide/lead sulfide thin film absorbers
,”
MRS Online Proc. Libr.
1012
,
413
418
(
2007
).
37.
Q.
Han
,
L.
Chen
,
M.
Wang
,
X.
Yang
,
L.
Lu
, and
X.
Wang
, “
Low-temperature synthesis of uniform Sb2S3 nanorods and its visible-light-driven photocatalytic activities
,”
Mater. Sci. Eng., B
166
(
1
),
118
121
(
2010
).
38.
B.
Jia
and
L.
Gao
, “
Growth of well-defined cubic hematite single crystals: Oriented aggregation and Ostwald ripening
,”
Cryst. Growth Des.
8
(
4
),
1372
1376
(
2008
).
39.
I. A.
Safo
,
M.
Werheid
,
C.
Dosche
, and
M.
Oezaslan
, “
The role of polyvinylpyrrolidone (PVP) as a capping and structure-directing agent in the formation of Pt nanocubes
,”
Nanoscale Adv.
1
(
8
),
3095
3106
(
2019
).
40.
G.-X.
Liang
,
Z.-H.
Zheng
,
P.
Fan
,
J.-T.
Luo
,
J.-G.
Hu
,
X.-H.
Zhang
,
H.-L.
Ma
,
B.
Fan
,
Z.-K.
Luo
, and
D.-P.
Zhang
, “
Thermally induced structural evolution and performance of Sb2Se3 films and nanorods prepared by an easy sputtering method
,”
Sol. Energy Mater. Sol. Cells
174
,
263
270
(
2018
).
41.
Q.
Wang
,
Z.
Chen
,
J.
Wang
,
Y.
Xu
,
Y.
Wei
,
Y.
Wei
,
L.
Qiu
,
H.
Lu
,
Y.
Ding
, and
J.
Zhu
, “
Sb2S3 solar cells: Functional layer preparation and device performance
,”
Inorg. Chem. Front.
6
(
12
),
3381
3397
(
2019
).
42.
T. J.
Whittles
,
T. D.
Veal
,
C. N.
Savory
,
A. W.
Welch
,
F. W.
de Souza Lucas
,
J. T.
Gibbon
,
M.
Birkett
,
R. J.
Potter
,
D. O.
Scanlon
,
A.
Zakutayev
, and
V. R.
Dhanak
, “
Core levels, band alignments, and valence-band states in CuSbS2 for solar cell applications
,”
ACS Appl. Mater. Interfaces
9
(
48
),
41916
41926
(
2017
).
43.
P.
Büttner
,
F.
Scheler
,
C.
Pointer
,
D.
Döhler
,
M. K. S.
Barr
,
A.
Koroleva
,
D.
Pankin
,
R.
Hatada
,
S.
Flege
,
A.
Manshina
,
E. R.
Young
,
I.
Mínguez-Bacho
, and
J.
Bachmann
, “
Adjusting interfacial chemistry and electronic properties of photovoltaics based on a highly pure Sb2S3 absorber by atomic layer deposition
,”
ACS Appl. Energy Mater.
2
(
12
),
8747
8756
(
2019
).
44.
Z.
Deng
,
D.
Chen
,
F.
Tang
,
J.
Ren
, and
A. J.
Muscat
, “
Synthesis and purple-blue emission of antimony trioxide single-crystalline nanobelts with elliptical cross section
,”
Nano Res.
2
(
2
),
151
160
(
2009
).
45.
W.
Shockley
and
H. J.
Queisser
, “
Detailed balance limit of efficiency of p‐n junction solar cells
,”
J. Appl. Phys.
32
(
3
),
510
519
(
1961
).
46.
Y.
Qi
,
Y.
Li
, and
Q.
Lin
, “
Engineering the charge extraction and trap states of Sb2S3 solar cells
,”
Appl. Phys. Lett.
120
(
22
),
221102
(
2022
).

Supplementary Material