Next Article in Journal
Structural Evolution and Sustainability of Agricultural Trade between China and Countries along the “Belt and Road”
Next Article in Special Issue
Circular Design Principles Applied on Dye-Sensitized Solar Cells
Previous Article in Journal
The Pricing Mechanism Analysis of China’s Natural Gas Supply Chain under the “Dual Carbon” Target Based on the Perspective of Game Theory
Previous Article in Special Issue
Influence/Effect of Deep-Level Defect of Absorber Layer and n/i Interface on the Performance of Antimony Triselenide Solar Cells by Numerical Simulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Back Contact Engineering to Improve CZTSSe Solar Cell Performance by Inserting MoO3 Sacrificial Nanolayers

1
Department of Optoelectronics and Materials Technology, National Taiwan Ocean University, Keelung City 202301, Taiwan
2
Center for Condensed Matter Sciences, National Taiwan University, Taipei 10617, Taiwan
3
Center of Atomic Initiative for New Materials, National Taiwan University, Taipei 10617, Taiwan
4
Department of Physics, National Taiwan University, Taipei 10617, Taiwan
5
Nano Science and Technology Program, Taiwan International Graduate Program, Academia Sinica, Taipei 115, Taiwan
6
Institute of Atomic and Molecular Sciences, Academia Sinica, Taipei 10617, Taiwan
7
Department of Materials Science and Engineering, National Taiwan University, Taipei 10617, Taiwan
8
Department of Vehicle Engineering, National Taipei University of Technology, Taipei 10608, Taiwan
9
Graduate Institute of Applied Science and Technology, National Taiwan University of Science and Technology, Taipei 10617, Taiwan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Sustainability 2022, 14(15), 9511; https://0-doi-org.brum.beds.ac.uk/10.3390/su14159511
Submission received: 3 July 2022 / Revised: 27 July 2022 / Accepted: 1 August 2022 / Published: 3 August 2022
(This article belongs to the Special Issue Toward Cost-Effective and Efficient Alternatives to Si Photovoltaics)

Abstract

:
Earth-abundant Cu2ZnSn(S,Se)4 (CZTSSe) is a promising nontoxic alternative compound for commercially available Cu(In,Ga)(S,Se)2 thin-film solar cells. In this study, a MoO3 nanolayer was applied as a sacrificial layer to optimize the quality of the interface between the CZTSSe and Mo back contact. MoO3 nanolayers can greatly improve CZTSSe grain growth and suppress the formation of some harmful secondary phases, especially the undesirable MoS(e)2. In terms of device performance, the series resistance was reduced from 1.83 to 1.54 Ω·cm2, and the fill factor was significantly enhanced from 42.67% to 52.12%. Additionally, MoO3 nanolayers improved CZTSSe absorber quality by lowering the defect energy levels from 228 to 148 meV. Furthermore, first-principles calculations demonstrate that the partial sulfoselenized MoO3 nanolayers may function as the (p-type) hole-selective contacts at Mo/CZTSSe interfaces, leading to an overall improvement in device performance. Lastly, a CZTSSe solar cell with about 26% improvement (compared with reference cells) in power conversion efficiency was achieved by inserting 5 nm MoO3 sacrificial layers.

1. Introduction

Thin-film chalcogenide-based solar cells such as CdTe [1] and Cu(In,Ga)(S,Se)2 (CIGSSe) [2,3,4] are commercially available and demonstrate over 20% power conversion efficiency (PCE). However, the toxicity of cadmium (Cd), and the scarcity of indium (In) and tellurium (Te) are major concerns for their applications [5,6]. Meanwhile, Earth-abundant and nontoxic CZTSSe, which shows a tunable direct bandgap of 1.1–1.5 eV by varying the S/Se ratio and high absorption coefficient of over 104 cm−1 [7,8,9], is treated as a promising alternative for CdTe and CIGSSe [10,11]. However, the open-circuit voltage (VOC) deficit and low fill factor (FF) critically obstruct PCE in CZTSSe solar cells, with the highest efficiency reaching 12.6% [12], which is significantly lower than the established chalcogenide absorbers [4]. On the basis of the existing configuration of CIGSSe solar cells, molybdenum (Mo) is also used as the back contact for CZTSSe. Nevertheless, Mo induces the detrimental decomposition of CZTSSe and the formation of performance-harmful binary secondary phases, such as CuxS(e)y, SnxS(e)y and ZnxS(e)y, leading to some voids near the back sides, and MoS(e)2 [13]. Although MoS(e)2 improves the adhesion and quasiohmic contact if it is thin enough [14,15,16], thick MoS(e)2 is more likely to be formed due to a required high sulfur (S) and/or selenium (Se) pressure [17] during the formation of a CZTSSe absorber. Moreover, CZTS-based absorbers are generally chalcogen-poor, and S/Se vacancies are classified as deep-level defects [18] leading to high series resistance (Rs), thus degrading the overall solar cell performance, especially the VOC deficit and low FF [17].
In order to suppress thick MoS(e)2 formation, several different intermediate layers were introduced between the CZTSSe absorber and Mo back contact, such as TiN [17,19], TiB2 [20], carbon [21], and ZnO [22]. In this regard, introducing a very thin MoOx at the interface between CZTSSe and Mo back contact can be advantageous because it may modify the surface energy of the Mo substrate and thus further enhance CZTSSe grain growth. From a band alignment point of view, oxidized MoS2 (i.e., MoS2-xOx) layers with p-type conductivity are also preferred [23,24,25,26,27]. N-type MoS2 layers with low work function are not suitable for the hole-transporting contact with p-type CZTSSe because it may form another p–n junction. In solar cell devices with nonohmic contacts, the VOC is limited by the work function difference between the top and bottom electrodes instead of the quasi-Fermi level splitting of the absorbers [10,28]. Therefore, MoOx can be a promising sacrificial material because not only the formation of a conductive p-type MoS(e)2-xOx layer is beneficial, but also the voids and unwanted secondary phases can be inhibited by avoiding direct reaction between CZTSSe and Mo.
In this work, we utilized a thermally evaporated ultrathin MoO3 nanolayer as a sacrificial layer to optimize the back contact interface between CZTSSe and Mo. Cross-sectional transmission electron macroscopy (TEM) combined with energy-dispersive X-ray spectroscopy (EDS) results confirmed that the sacrificial nanolayers increase grain growth (from ~800 to ~1000 nm) and also greatly suppress the formation of MoS(e)2 layer (from ~350 to ~100 nm) and other harmful secondary phases. As a result, Rs was reduced from 1.83 to 1.54 Ω·cm2, and FF was significantly enhanced from 42.67% to 52.12%. Additionally, a MoO3 nanolayer improved the CZTSSe absorber quality by lowering the defect energy levels from 228 to 148 meV, which was determined by admittance spectroscopy (AS). To fundamentally investigate the role of MoO3 nanolayers at the back contact of CZTSSe solar cells, a first-principles study was conducted using the Vienna Ab initio Simulation Package (VASP) [29,30,31]. Calculations reveal that the partial sulfoselenized MoO3 nanolayers may be treated as a hole-selective contact between Mo/CZTSSe interfaces, and is beneficial for the device performance. Lastly, we demonstrate that introducing an ultrathin MoO3 sacrificial nanolayer remarkably improved the CZTSSe absorber’s quality, and optimized the back contact interface between CZTSSe and Mo, leading to the enhancement of overall solar cell performance.

2. Experimental Details

Multimetallic stacked Sn/Zn/Cu (CZT) precursor film of 600 nm thick was deposited onto a 1 μm Mo-coated (with/without ultrathin MoOx nanolayers) soda-lime glass (SLG) substrate using the RF magnetron sputtering process. To synthesize CZTSSe absorbers, the CZT precursor films were annealed in a semisealed graphite box (7 cm3 in volume) with Se pellets at a temperature of 550 °C for 7 mins in 20% H2S (99.95%) diluted with Ar (99.99%) at atmospheric pressure (Cu/Sn + Zn ~0.9 and Zn/Sn ~1.14 measured by X-ray fluorescence (XRF), XRF-1800, Shimadzu Scientific Instrument). The p–n heterojunction was formed by depositing 40 nm CdS buffer layer on top of CZTSSe film using chemical bath deposition method. A 300 nm DC-sputtered ITO/40 nm ZnO contact was then directly deposited onto CdS/CZTSSe films. Lastly, a 300 nm Ag top metal contact and 100 nm MgF2 antireflection coating were deposited on top of all devices by thermal evaporation. The device area (~0.105 cm2) was defined by shadow masks and mechanical scribing, resulting in a cell effective area of ~0.095 cm2 [10,11]. The current density–voltage (J–V) characteristics of the solar cell devices were measured using a Keithley 2400 source meter under the illumination of an AM 1.5G Solar Simulator (Newport) with light intensity of 100 mW/cm2, which was calibrated with an NREL-traceable Si reference cell. The spectral response and the external quantum efficiency (EQE) of cells were obtained using an ENLI Technology model EQE-D-3011 system. Temperature-dependent (120–300 K) capacitance–frequency (C–F) measurements were conducted by an Agilent E4980A LCR meter in the frequency range of 102–106 Hz at 0 V with an AC amplitude of 30 mV under dark conditions [32].

3. Results and Discussion

To confirm the successful deposition of an ultrathin MoOx film on the Mo-coated SLGs, X-ray photoelectron spectroscopy (XPS) measurements were performed, as shown in Figure 1. For the pristine Mo-coated SLGs (red line), the peaks of Mo3d5/2 and Mo3d3/2 were located at 228.3 and 231.5 eV, respectively, with a 3.2 eV energy difference, consistent with the literature values [33], indicating that there was almost no native metal oxide on the pristine Mo-coated SLGs. Once some MoOx had been thermally evaporated on the Mo-SLG substrates, except for the residual small peaks located at 228.3 eV (Mo3d5/2 for pristine Mo), these XPS peaks were shifted to 232.9 eV (Mo3d5/2 for MoO3) and 236.1 eV (Mo3d3/2 for MoO3), respectively, with 3.2 eV energy difference, demonstrating that 5 and 10 nm ultrathin MoO3 (not MoO2) nanolayers were successfully deposited on the Mo substrates. Moreover, the ultrathin MoO3 layers can be prepared by the reactive sputtering of 1 μm thick Mo on SLGs, which is compatible with the subsequent fabrication for CIGSSe solar cells [34].
The 600 nm thick Sn/Zn/Cu (CZT) multimetallic stacked precursors were sputtering-deposited on our Mo-SLGs with or without MoO3 layers [11,35]. The X-ray diffraction (XRD) of metal stacked precursors on Mo-SLGs (in Figure 2) showed two characteristic peaks of well-intermixed metal alloys at 30.26°, corresponding to bronze (Cu6Sn5 JCPDS no. 45–1488), and 43.26°, belonging to brass (Cu5Zn8 JCPDS no. 25–1288). Along with these two phases, the characteristic peaks of Sn were also found at 30.71°, 32.10°, 43.95°, and 44.97° (JCPDS No. 650296). The well-intermixed precursors are beneficial to the CZTSSe growth during sulfoselenization [11,35].
Raman spectroscopy with 473 nm wavelength excitation (around ~550 nm sampling depth) was conducted to examine the quality of CZTSSe films, as displayed in Figure 3. The signatures at 176, 207, and 239 cm−1 correspond to the CZTSe vibration modes, and the peak at 329 cm−1 corresponds to the CZTS vibration mode [11,35,36]. No peak shifts or secondary phases were observed between the pristine and 5 or 10 nm MoO3-modified CZTSSe films. The progressive peak sharpening suggests the better quality of CZTSSe films grown on MoO3 modified substrates compared with the pristine one. The Raman spectra can also be utilized to determine the sulfur and selenium (S/Se) ratio in the CZTSSe by taking the ratio of the peak shift for the sulfide based of the pure CZTS (338 cm−1) and selenide based on the pure CZTSe (196 cm−1) [9]. By calculating the ratio of the peak shift for sulfide-based (329 cm−1) and selenide-based (207 cm−1), an [S]/[S] + [Se] value of ~0.45 was obtained, leading to the S/Se ratio of 45:55.
To investigate the MoO3 effects on solar cell performances, the light J–V was measured on all the devices, as shown in Figure 4. For statistical study, there were 10 device cells for each condition. The average device performance and standard deviations (±) for different device conditions are summarized in Table 1. All cells with MoO3 nanolayers exhibited higher short circuit current (JSC), VOC, and FF compared to the reference cells. The significant improvements in Voc from 450 to 473 mV, JSC from 25.86 to 27.33 mA/cm2, and FF from 42.67% to 52.12% for devices inserted with a 5 nm MoO3 layer were attributed to the increased shunt resistance (RSH) from 145 to 151 Ω·cm2 and the reduced RS from 1.83 to 1.54 Ω·cm2, which may have resulted from the improvement in back contact quality. Lastly, a CZTSSe solar cell with 7.78% PCE was obtained by inserting 5 nm MoO3 layers, which was an improvement of about 26% compared with 5.76% PCE of the pristine one.
To inspect the CZTSSe absorber quality near the Mo back contacts, cross-sectional TEM and EDS mapping was executed, as presented in Figure 5 and Figure 6, respectively. MoS(e)2 thickness was reduced from ~350 to ~100 nm, and CZTSSe grain sizes were enlarged by inserting 5 nm MoO3 layers (Figure 5). In the pristine absorber, the void formation near the back contacts may have arisen from the vaporization of highly volatile secondary phases (i.e., SnxS and SnxSey) during the focused ion beam (FIB) fabrication of TEM samples. A large number of voids from secondary phases forming at the back interface limited the free-carrier transportation, leading to higher RS and thereby lower JSC [21]. EDS mapping (Figure 6) also directly illustrates the formation of unwanted secondary phases, such as CuxSy, ZnxSy, SnxSy, and SnxSey, near the untreated Mo back contacts, which further proves that the CZTSSe/Mo interface was chemically unstable [19]. However, the inserted MoO3 nanolayers could not be identified by the oxygen/molybdenum elemental mapping images in Figure 6b. MoO3 nanolayers may have been fully sulfoselenized to p-type MoS(e)2-xOx during a high-temperature synthesis process. Therefore, the increased VOC and FF, and reduced RS in the modified devices may have been due to the larger grains of the CZTSSe absorber (i.e., fewer grain boundary recombination centers), the suppression of some harmful secondary phases, and the thinner MoS(e)2 by inserting MoO3 sacrificial nanolayers.
To understand why the device performance was not monotonically improved as MoO3 thickness increased, cross-sectional TEM analysis was conducted, as shown in Figure 7. Inserting 10 nm MoO3 increased the Rs, and reduced the JSC and FF, thus decreasing the overall PCE (Table 1). The TEM results (Figure 7) show that the orientation of MoS(e)2 can be vertically or horizontally aligned with the Mo substrate, which can provide different contact resistances. On the basis of the direction of the photocarrier separation (i.e., built-in electric field direction), the vertically aligned MoS(e)2 was more conductive than the horizontally aligned one is. In fact, MoS(e)2 is a layered material with a weak van der Waals bond between each layer; thus, its interlayer (i.e., horizontally aligned) conductivity is two orders of magnitude lower than its intralayer (i.e., vertically aligned) conductivity (Figure 7) [37,38,39]. The vertically aligned MoS(e)2 came from the pristine Mo contact that reacted with S/Se vapor during the annealing process, while the horizontally aligned MoS(e)2 came from the MoO3 layers that lost their oxygen during the sulfoselenization process due to the high vapor pressure of S/Se [40]. Therefore, it is preferable that the MoO3 layers be transformed into horizontally aligned MoS(e)2 (or MoS(e)2−xOx) because it can hinder S/Se diffusion during annealing process [40]. Hence, undesirable MoS(e)2 thickness can be minimized. This shows that the higher Rs of 1.77 Ω·cm2 for 10 nm MoO3 devices compared with 1.54 Ω·cm2 for 5 nm MoO3 devices was due to a thicker MoO3 that transformed into a more horizontally aligned MoS(e)2. As a result, JSC was reduced from 27.33 to 26.30 mA/cm2 by inserting the thicker MoO3, as shown in Table 1.
To further study the effect of the MoO3 sacrificial layer on defect states in CZTSSe, AS measurement was performed [32]. Figure 8 reveals the C–F profiling for CZTSSe devices with MoO3 thickness of 0, 5, and 10 nm. Using the model in the work of Kimerling [41,42], capacitance at a high frequency represents the response of the free carrier density, while capacitance at a low frequency represents the response of the sum of free carriers and deep traps. The Arrhenius plot (Figure 9) was used to determine the activation energy of deep trap states with various MoO3 thickness levels. For each AS spectrum, the inflection point frequency (or step frequency ω 0 ) is determined by applying angular frequency at the maximum of the ω dC/d ω plot. The inflection point frequency of each AS spectrum can be used to construct the Arrhenius plot by fitting the spectra according to the following equation:
ω 0 = 2 π ν 0 T 2 e x p ( E a k T )
where ω 0 is the step frequency, E a is the energetic depth of the defect relative to the corresponding band edge, ν 0 is the pre-exponential factor comprising temperature-independent parts such as defect capture cross-sections for holes σp, thermal velocity ν t h , and the effective density of states in the valence band Nv.
Activation energy Ea ascertained with the Arrhenius plot was approximately the energy difference between the defect level and the valence band edge. Assuming that their energy levels were within a small range, the Ea value deduced from the Arrhenius plot could also represent the average value of activation energies of a band of defects in the band gap. The activation energies of CZTSSe devices with MoO3 thicknesses of 0, 5, and 10 nm were 228, 148, and 199 meV, respectively (Figure 9). Both results support the argument that, besides suppressing MoS(e)2 and secondary formation at the back interface, introducing MoO3 as a sacrificial nanolayer may help in improving CZTSSe absorber quality by not only increasing the grain growth, but also lowering the activation energy of defect states.
To fundamentally investigate the role of MoO3 nanolayers at the back contacts of CZTSSe solar cells, first-principles calculations were conducted through VASP [29,30,31] to examine the electronic properties of S-doped MoO3. The projector-augmented wave (PAW) method was adopted, and Perdew–Burke–Ernzerhof (PBE) exchange correlation functionals were used to account for electron–electron interactions in our system [43]. Monkhorst–Pack K-point grids of 2 × 2 × 2 and 400 eV energy cutoff for plane wave basis set were used for geometric optimization. The geometric structures were entirely relaxed when the total energy was converged to 10−6 eV in the 3 × 2 MoO3 supercell model. The density-of-states (DOS) calculation of S-doped MoO3 was performed using the same parameters except for the 6 × 6 × 6 Monkhorst–Pack K-point grids. Due to sulfidation, sulfur appeared in MoO2, and this could be regarded as a sulfur substitution for oxygen in which three different sites substitutive defects could be formed, denoted as SO1, SO2, and SO3. SO1 is single oxygen coordinated, while SO2 and SO3 are double and triple oxygen coordinated, respectively, as shown in Figure 10.
Figure 11 reveals the DOS of MoO3 with SO defects, which indicates that there were additional defective states (left-hand side near 0 eV) with slightly higher energy than that of the valence band maximum (VBM). There was no significant difference between sulfurization and selenization for MoO3. The small difference was that the additional defective states of selenization were a little serious compared with the sulfurization one. The calculated DOS was almost the same for these three different defect sites, which means that the DOS was not sensitive to the chemical environment of SO, and only one is shown. These defective SO states above the VBM, which contribute to the p-type carriers (holes) within the partial sulfoselenized MoO3 nanolayers, may be treated as hole-selective contacts between Mo/CZTSSe interfaces, further improving the overall PCE, as demonstrated experimentally.

4. Conclusions

While CZTSSe is a very promising alternative for CdTe and CIGSSe for thin-film solar cells, its VOC deficit and low FF critically obstruct the PCE. The formation of many harmful secondary phases at the back interface between CZTSSe and Mo leads to high series resistance (Rs), thus degrading overall solar cell performance. In this study, a MoO3 nanolayer was applied as a sacrificial layer to optimize the back contact interface between CZTSSe and Mo. Cross-sectional TEM and EDS results show that the MoO3 sacrificial nanolayers not only improved CZTSSe grain growth, but also successfully suppressed the formation of an undesirable MoS(e)2 layer (from ~350 to ~100 nm) and other harmful secondary phases. From the device point of view, the Rs was reduced from 1.83 to 1.54 Ω·cm2 and the FF was significantly increased from 42.67% to 52.12%, resulting in an increase in Voc by 23 mV. Importantly, MoO3 nanolayers also improved CZTSSe absorber quality by lowering the defect energy levels from 228 to 148 meV. In addition to the experimental results, the first-principles calculations revealed that the partial sulfoselenized MoO3 nanolayers may be treated as the (p-type) hole-selective contacts between Mo/CZTSSe interfaces, contributing to an overall improvement in device performance. Lastly, a CZTSSe solar cell with 7.78% PCE was obtained by inserting 5 nm MoO3 nanolayers, which was an improvement of about 26% compared with the 5.76% PCE of the pristine one.

Author Contributions

C.-Y.C. and S.K. contributed equally for the first authorship. C.-Y.C., S.K. and W.-C.C. conceived the original idea and research plan. C.-Y.C. and S.K. conceived the presented idea, carried out the experiment, analyzed the results, and wrote the manuscript. Y.-R.L. and S.K. conducted the AS experiments. W.-C.C. executed the XRD. C.-Y.C. and W.-C.C. analyzed the TEM characterization. J.-W.L. and P.-T.C. performed the simulation with first-principles calculations. C.-Y.C., K.-H.C. and L.-C.C. supervised this project. All authors discussed the results and contributed to the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the National Science and Technology Council (NSTC) in Taiwan under the Academic Summit Project (107-2745-M-002-001-ASP), the Science Vanguard Project (108-2119-M-002-030, 109-2123-M-002-004 and 110-2123-M-002-006), MOST 107-2112-M-131-002-MY2, and MOST 111-2112-M-131 -002 -MY3. Financial support is acknowledged from the Center of Atomic Initiative for New Materials (AI-Mat), the National Taiwan University (108L9008, 109 L9008 and 110 L9008), and the Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE) in Taiwan.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank Ms. C.-Y. Chien of the National Science and Technology Council (NSTC) (National Taiwan University) for their assistance in the FIB experiments. We also acknowledge the technical support provided by the core facilities for nanoscience and nanotechnology at Academia Sinica, National Taiwan University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Green, M.A.; Hishikawa, Y.; Dunlop, E.D.; Levi, D.H.; Hohl-Ebinger, J.; Yoshita, M.; Ho-Baillie, A.W.Y. Solar cell efficiency tables (Version 53). Prog. Photovolt. Res. Appl. 2019, 27, 3–12. [Google Scholar] [CrossRef] [Green Version]
  2. Jackson, P.; Wuerz, R.; Hariskos, D.; Lotter, E.; Witte, W.; Powalla, M. Effects of heavy alkali elements in Cu(In,Ga)Se2 solar cells with efficiencies up to 22.6%. Phys. Status Solidi (RRL) Rapid Res. Lett. 2016, 10, 583–586. [Google Scholar] [CrossRef] [Green Version]
  3. Chirila, A.; Reinhard, P.; Pianezzi, F.; Bloesch, P.; Uhl, A.R.; Fella, C.; Kranz, L.; Keller, D.; Gretener, C.; Hagendorfer, H.; et al. Potassium-induced surface modification of Cu(In,Ga)Se2 thin films for high-efficiency solar cells. Nat. Mater. 2013, 12, 1107–1111. [Google Scholar] [CrossRef] [PubMed]
  4. Nakamura, M.; Yamaguchi, K.; Kimoto, Y.; Yasaki, Y.; Kato, T.; Sugimoto, H. Cd-Free Cu(In,Ga)(Se,S)2 Thin-Film Solar Cell With Record Efficiency of 23.35%. IEEE J. Photovolt. 2019, 9, 1863–1867. [Google Scholar] [CrossRef]
  5. Ginley, D.; Green, M.A.; Collins, R. Solar Energy Conversion Toward 1 Terawatt. MRS Bull. 2011, 33, 355–364. [Google Scholar] [CrossRef] [Green Version]
  6. Butler, D. Thin Films: Ready for their close-up? Nature 2008, 454, 558–559. [Google Scholar] [CrossRef] [Green Version]
  7. Katagiri, H.; Nishimura, M.; Onozawa, T.; Maruyama, S.; Fujita, M.; Sega, T.; Watanabe, T. Rare-metal free thin film solar cell. In Proceedings of the Power Conversion Conference, Nagaoka, Japan, 6 August 1997. [Google Scholar]
  8. Katagiri, H.; Sasaguchi, N.; Hando, S.; Hoshino, S.; Ohashi, J.; Yokota, T. Preparation and evaluation of Cu2ZnSnS4 thin films by sulfurization of E-B evaporated precursors. Sol. Energy Mater. Sol. Cells 1997, 49, 407–414. [Google Scholar] [CrossRef]
  9. Mitzi, D.B.; Gunawan, O.; Todorov, T.K.; Wang, K.; Guha, S. The path towards a high-performance solution-processed kesterite solar cell. Sol. Energy Mater. Sol. Cells 2011, 95, 1421–1436. [Google Scholar] [CrossRef]
  10. Chen, C.-Y.; Aprillia, B.S.; Chen, W.-C.; Teng, Y.-C.; Chiu, C.-Y.; Chen, R.-S.; Hwang, J.-S.; Chen, K.-H.; Chen, L.-C. Above 10% efficiency earth-abundant Cu2ZnSn(S,Se)4 solar cells by introducing alkali metal fluoride nanolayers as electron-selective contacts. Nano Energy 2018, 51, 597–603. [Google Scholar] [CrossRef]
  11. Chen, W.-C.; Chen, C.-Y.; Tunuguntla, V.; Lu, S.H.; Su, C.; Lee, C.-H.; Chen, K.-H.; Chen, L.-C. Enhanced solar cell performance of Cu2ZnSn(S,Se)4 thin films through structural control by using multi-metallic stacked nanolayers and fast ramping process for sulfo-selenization. Nano Energy 2016, 30, 762–770. [Google Scholar] [CrossRef]
  12. Wang, W.; Winkler, M.T.; Gunawan, O.; Gokmen, T.; Todorov, T.K.; Zhu, Y.; Mitzi, D.B. Device Characteristics of CZTSSe Thin-Film Solar Cells with 12.6% Efficiency. Adv. Energy Mater. 2014, 4, 1301465. [Google Scholar] [CrossRef]
  13. Scragg, J.J.; Watjen, J.T.; Edoff, M.; Ericson, T.; Kubart, T.; Platzer-Bjorkman, C. A detrimental reaction at the molybdenum back contact in Cu2ZnSn(S,Se)4 thin-film solar cells. J. Am. Chem. Soc. 2012, 134, 19330–19333. [Google Scholar] [CrossRef] [PubMed]
  14. Shin, B.; Bojarczuk, N.A.; Guha, S. On the kinetics of MoSe2 interfacial layer formation in chalcogen-based thin film solar cells with a molybdenum back contact. Appl. Phys. Lett. 2013, 102, 091907. [Google Scholar] [CrossRef] [Green Version]
  15. Scragg, J.J.; Ericson, T.; Kubart, T.; Edoff, M.; Platzer-Björkman, C. Chemical Insights into the Instability of Cu2ZnSnS4 Films during Annealing. Chem. Mater. 2011, 23, 4625–4633. [Google Scholar] [CrossRef]
  16. Nishiwaki, S.; Kohara, N.; Negami, T.; Wada, T. MoSe2 layer formation at Cu(In,Ga)Se2/Mo Interfaces in High Efficiency Cu(In1-xGax)Se2 Solar Cells. Jpn. J. Appl. Phys. 1998, 37, 71–73. [Google Scholar] [CrossRef]
  17. Shin, B.; Zhu, Y.; Bojarczuk, N.A.; Jay Chey, S.; Guha, S. Control of an interfacial MoSe2 layer in Cu2ZnSnSe4 thin film solar cells: 8.9% power conversion efficiency with a TiN diffusion barrier. Appl. Phys. Lett. 2012, 101, 053903. [Google Scholar] [CrossRef]
  18. Walsh, A.; Chen, S.; Wei, S.-H.; Gong, X.-G. Kesterite Thin-Film Solar Cells: Advances in Materials Modelling of Cu2ZnSnS4. Adv. Energy Mater. 2012, 2, 400–409. [Google Scholar] [CrossRef]
  19. Scragg, J.J.; Kubart, T.; Wätjen, J.T.; Ericson, T.; Linnarsson, M.K.; Platzer-Björkman, C. Effects of Back Contact Instability on Cu2ZnSnS4 Devices and Processes. Chem. Mater. 2013, 25, 3162–3171. [Google Scholar] [CrossRef]
  20. Liu, F.; Sun, K.; Li, W.; Yan, C.; Cui, H.; Jiang, L.; Hao, X.; Green, M.A. Enhancing the Cu2ZnSnS4 solar cell efficiency by back contact modification: Inserting a thin TiB2 intermediate layer at Cu2ZnSnS4/Mo interface. Appl. Phys. Lett. 2014, 104, 051105. [Google Scholar] [CrossRef]
  21. Zhou, F.; Zeng, F.; Liu, X.; Liu, F.; Song, N.; Yan, C.; Pu, A.; Park, J.; Sun, K.; Hao, X. Improvement of Jsc in a Cu2ZnSnS4 Solar Cell by Using a Thin Carbon Intermediate Layer at the Cu2ZnSnS4/Mo Interface. ACS Appl. Mater. Interfaces 2015, 7, 22868–22873. [Google Scholar] [CrossRef]
  22. López-Marino, S.; Placidi, M.; Pérez-Tomás, A.; Llobet, J.; Izquierdo-Roca, V.; Fontané, X.; Fairbrother, A.; Espíndola-Rodríguez, M.; Sylla, D.; Pérez-Rodríguez, A.; et al. Inhibiting the absorber/Mo-back contact decomposition reaction in Cu2ZnSnSe4 solar cells: The role of a ZnO intermediate nanolayer. J. Mater. Chem. A 2013, 1, 8338–8343. [Google Scholar] [CrossRef]
  23. Yin, Z.; Zhang, X.; Cai, Y.; Chen, J.; Wong, J.I.; Tay, Y.Y.; Chai, J.; Wu, J.; Zeng, Z.; Zheng, B.; et al. Preparation of MoS2-MoO3 hybrid nanomaterials for light-emitting diodes. Angew. Chem. 2014, 53, 12560–12565. [Google Scholar] [CrossRef]
  24. Qin, P.; Fang, G.; Ke, W.; Cheng, F.; Zheng, Q.; Wan, J.; Lei, H.; Zhao, X. In situ growth of double-layer MoO3/MoS2 film from MoS2 for hole-transport layers in organic solar cell. J. Mater. Chem. A 2014, 2, 2742–2756. [Google Scholar] [CrossRef]
  25. McDonnell, S.; Azcatl, A.; Addou, R.; Gong, C.; Battaglia, C.; Chuang, S.; Cho, K.; Javey, A.; Wallace, R.M. Hole Contacts on Transition Metal Dichalcogenides: Interface Chemistry and Band Alignments. ACS Nano 2014, 8, 6265–6272. [Google Scholar] [CrossRef] [PubMed]
  26. Yun, J.M.; Noh, Y.J.; Lee, C.H.; Na, S.I.; Lee, S.; Jo, S.M.; Joh, H.I.; Kim, D.Y. Exfoliated and partially oxidized MoS2 nanosheets by one-pot reaction for efficient and stable organic solar cells. Small 2014, 10, 2319–2324. [Google Scholar] [CrossRef] [PubMed]
  27. Chuang, S.; Battaglia, C.; Azcatl, A.; McDonnell, S.; Kang, J.S.; Yin, X.; Tosun, M.; Kapadia, R.; Fang, H.; Wallace, R.M.; et al. MoS2 P-type transistors and diodes enabled by high work function MoOx contacts. Nano Lett. 2014, 14, 1337–1342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Gao, S.; Jiang, Z.; Wu, L.; Ao, J.; Zeng, Y.; Sun, Y.; Zhang, Y. Interfaces of high-efficiency kesterite Cu2ZnSnS(e)
    4 thin film solar cells. Chin. Phys. B 2018, 27, 018803. [Google Scholar] [CrossRef] [Green Version]
  29. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  30. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251–14269. [Google Scholar] [CrossRef]
  31. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  32. Lin, Y.-R.; Tunuguntla, V.; Wei, S.-Y.; Chen, W.-C.; Wong, D.; Lai, C.-H.; Liu, L.-K.; Chen, L.-C.; Chen, K.-H. Bifacial sodium-incorporated treatments: Tailoring deep traps and enhancing carrier transport properties in Cu2ZnSnS4 solar cells. Nano Energy 2015, 16, 438–445. [Google Scholar] [CrossRef]
  33. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X-ray Photoelectron Spectroscopy: A Reference Book of Standard Spectra for Identification and Interpretation of XPS Data; Physical Electronics: Chanhassen, MN, USA, 1995. [Google Scholar]
  34. Zhou, D.; Zhu, H.; Liang, X.; Zhang, C.; Li, Z.; Xu, Y.; Chen, J.; Zhang, L.; Mai, Y. Sputtered molybdenum thin films and the application in CIGS solar cells. Appl. Surf. Sci. 2016, 362, 202–209. [Google Scholar] [CrossRef]
  35. Chen, W.-C.; Tunuguntla, V.; Li, H.-W.; Chen, C.-Y.; Li, S.-S.; Hwang, J.-S.; Lee, C.-H.; Chen, L.-C.; Chen, K.-H. Fabrication of Cu2ZnSnSe4 solar cells through multi-step selenization of layered metallic precursor film. Thin Solid Film. 2016, 618, 42–49. [Google Scholar] [CrossRef]
  36. Gürel, T.; Sevik, C.; Çağın, T. Characterization of vibrational and mechanical properties of quaternary compounds Cu2ZnSnS4 and Cu2ZnSnSe4 in kesterite and stannite structures. Phys. Rev. B 2011, 84, 205201. [Google Scholar] [CrossRef]
  37. Evans, B.L.; Young, P.A. Optical absorption and dispersion in molybdenum disulphide. Proc. R. Soc. Lond. A 1965, 284, 402–422. [Google Scholar] [CrossRef]
  38. Hu, S.Y.; Liang, C.H.; Tiong, K.K.; Lee, Y.C.; Huang, Y.S. Preparation and characterization of large niobium-doped MoSe2 single crystals. J. Cryst. Growth 2005, 285, 408–414. [Google Scholar] [CrossRef]
  39. Laursen, A.B.; Kegnæs, S.; Dahl, S.; Chorkendorff, I. Molybdenum sulfides—Efficient and viable materials for electro-and photoelectrocatalytic hydrogen evolution. Energy Environ. Sci. 2012, 5, 5577. [Google Scholar] [CrossRef]
  40. Kong, D.; Wang, H.; Cha, J.J.; Pasta, M.; Koski, K.J.; Yao, J.; Cui, Y. Synthesis of MoS2 and MoSe2 films with vertically aligned layers. Nano Lett. 2013, 13, 1341–1347. [Google Scholar] [CrossRef]
  41. Kimerling, L.C. Influence of deep traps on the measurement of free-carrier distributions in semiconductors by junction capacitance techniques. J. Appl. Phys. 1974, 45, 1839–1845. [Google Scholar] [CrossRef]
  42. Duan, H.-S.; Yang, W.; Bob, B.; Hsu, C.-J.; Lei, B.; Yang, Y. The Role of Sulfur in Solution-Processed Cu2ZnSn(S,Se)4 and its Effect on Defect Properties. Adv. Funct. Mater. 2013, 23, 1466–1471. [Google Scholar] [CrossRef]
  43. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. XPS spectra of thermally evaporated 0, 5, and 10 nm MoOx on Mo-SLG substrates.
Figure 1. XPS spectra of thermally evaporated 0, 5, and 10 nm MoOx on Mo-SLG substrates.
Sustainability 14 09511 g001
Figure 2. XRD of Sn/Zn/Cu (CZT) multimetallic stacked precursors on Mo-SLG substrates.
Figure 2. XRD of Sn/Zn/Cu (CZT) multimetallic stacked precursors on Mo-SLG substrates.
Sustainability 14 09511 g002
Figure 3. Raman spectra of CZTSSe grown on the Mo-SLGs without MoO3, with 5 and 10 nm MoO3 sacrificial nanolayers.
Figure 3. Raman spectra of CZTSSe grown on the Mo-SLGs without MoO3, with 5 and 10 nm MoO3 sacrificial nanolayers.
Sustainability 14 09511 g003
Figure 4. Current density–voltage (J–V) characteristics of CZTSSe devices without MoO3, with 5 and 10 nm MoO3 sacrificial nanolayers.
Figure 4. Current density–voltage (J–V) characteristics of CZTSSe devices without MoO3, with 5 and 10 nm MoO3 sacrificial nanolayers.
Sustainability 14 09511 g004
Figure 5. Cross-sectional TEM images of CZTSSe devices (a) without MoO3 and (b) with 5 nm MoO3 sacrificial nanolayers.
Figure 5. Cross-sectional TEM images of CZTSSe devices (a) without MoO3 and (b) with 5 nm MoO3 sacrificial nanolayers.
Sustainability 14 09511 g005
Figure 6. Elemental mapping images of CZTSSe films grown on the Mo-SLGs (a) without MoO3 and (b) with 5 nm MoO3 sacrificial nanolayers.
Figure 6. Elemental mapping images of CZTSSe films grown on the Mo-SLGs (a) without MoO3 and (b) with 5 nm MoO3 sacrificial nanolayers.
Sustainability 14 09511 g006
Figure 7. Cross-sectional TEM of (a) MoO3-modified and (b) pristine Mo/CZTSSe interfaces for determining the MoS(e)2 orientations relative to the Mo substrates.
Figure 7. Cross-sectional TEM of (a) MoO3-modified and (b) pristine Mo/CZTSSe interfaces for determining the MoS(e)2 orientations relative to the Mo substrates.
Sustainability 14 09511 g007
Figure 8. Admittance spectra of CZTSSe devices with different thickness levels of MoO3.
Figure 8. Admittance spectra of CZTSSe devices with different thickness levels of MoO3.
Sustainability 14 09511 g008
Figure 9. Arrhenius plot of the characteristic frequencies to extract the activation energy of CZTSSe devices with different thickness levels of MoO3.
Figure 9. Arrhenius plot of the characteristic frequencies to extract the activation energy of CZTSSe devices with different thickness levels of MoO3.
Sustainability 14 09511 g009
Figure 10. The optimized structure of MoO3 containing (a) SO1, (b) SO2, and (c) SO3 defects.
Figure 10. The optimized structure of MoO3 containing (a) SO1, (b) SO2, and (c) SO3 defects.
Sustainability 14 09511 g010
Figure 11. DOS of MoO3 with SO defects for (a) sulfurization and (b) selenization. The energy levels of defective states from SO defects were slightly higher than those of VBM.
Figure 11. DOS of MoO3 with SO defects for (a) sulfurization and (b) selenization. The energy levels of defective states from SO defects were slightly higher than those of VBM.
Sustainability 14 09511 g011
Table 1. Solar cell parameters of CZTSSe devices with different thickness of MoO3.
Table 1. Solar cell parameters of CZTSSe devices with different thickness of MoO3.
PCE (%)JSC (mA/cm2)VOC (mV)FF (%)RS (Ω·cm2)RSH (Ω·cm2)
Pristine5.21 ± 0.6625.86 ± 2.26450 ± 1542.67 ± 1.51.83 ± 0.44133 ± 45
5 nm MoO36.80 ± 0.8527.33 ± 2.20473 ± 552.12 ± 3.01.54 ± 0.23153 ± 68
10 nm MoO36.49 ± 0.5526.30 ± 1.57474 ± 750.69 ± 1.51.77 ± 0.26252 ± 63
Best cell
(5 nm MoO3)
7.7828.8448055.831.68180
Statistics from 10 cells for each condition.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, C.-Y.; Kholimatussadiah, S.; Chen, W.-C.; Lin, Y.-R.; Lin, J.-W.; Chen, P.-T.; Chen, R.-S.; Chen, K.-H.; Chen, L.-C. Back Contact Engineering to Improve CZTSSe Solar Cell Performance by Inserting MoO3 Sacrificial Nanolayers. Sustainability 2022, 14, 9511. https://0-doi-org.brum.beds.ac.uk/10.3390/su14159511

AMA Style

Chen C-Y, Kholimatussadiah S, Chen W-C, Lin Y-R, Lin J-W, Chen P-T, Chen R-S, Chen K-H, Chen L-C. Back Contact Engineering to Improve CZTSSe Solar Cell Performance by Inserting MoO3 Sacrificial Nanolayers. Sustainability. 2022; 14(15):9511. https://0-doi-org.brum.beds.ac.uk/10.3390/su14159511

Chicago/Turabian Style

Chen, Cheng-Ying, Septia Kholimatussadiah, Wei-Chao Chen, Yi-Rung Lin, Jia-Wei Lin, Po-Tuan Chen, Ruei-San Chen, Kuei-Hsien Chen, and Li-Chyong Chen. 2022. "Back Contact Engineering to Improve CZTSSe Solar Cell Performance by Inserting MoO3 Sacrificial Nanolayers" Sustainability 14, no. 15: 9511. https://0-doi-org.brum.beds.ac.uk/10.3390/su14159511

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop