Next Article in Journal
Fiber Optic Sensor for Real-time Monitoring of Freezing–Thawing Cycle in Cryosurgery
Previous Article in Journal
Portable Sensor System for Registration, Processing and Mathematical Analysis of PPG Signals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Substrate Temperature during InxSy Sputtering on Cu(In,Ga)Se2/Buffer Interface Properties and Solar Cell Performance

Zentrum für Sonnenenergie- und Wasserstoff-Forschung Baden-Württemberg (ZSW), 70563 Stuttgart, Germany
*
Author to whom correspondence should be addressed.
Appl. Sci. 2020, 10(3), 1052; https://doi.org/10.3390/app10031052
Submission received: 20 December 2019 / Revised: 23 January 2020 / Accepted: 31 January 2020 / Published: 5 February 2020
(This article belongs to the Special Issue Selected Papers from NEXTGEN’19: Next Generation High-Efficiency)

Abstract

:
Indium sulfide (InxSy)—besides CdS and Zn(O,S)—is already used as a buffer layer in chalcopyrite-type thin-film solar cells and modules. We discuss the influence of the substrate temperature during very fast magnetron sputtering of InxSy buffer layers on the interface formation and the performance of Cu(In,Ga)Se2 solar cells. The substrate temperature was increased from room temperature up to 240 °C, and the highest power conversion efficiencies were obtained at a temperature plateau around 200 °C, with the best values around 15.3%. Industrially relevant in-line co-evaporated polycrystalline Cu(In,Ga)Se2 absorber layers were used, which yield solar cell efficiencies of up to 17.1% in combination with a solution-grown CdS buffer. The chemical composition of the InxSy buffer as well as of the Cu(In,Ga)Se2/InxSy interface was analyzed by time-of-flight secondary ion mass spectrometry. Changes from homogenous and stoichiometric In2S3 layers deposited at RT to inhomogenous and more sulfur-rich and indium-deficient compositions for higher temperatures were observed. This finding is accompanied with a pronounced copper depletion at the Cu(In,Ga)Se2 absorber surface, and a sodium accumulation in the InxSy buffer and at the absorber/buffer interface. These last two features seem to be the origin for achieving the highest conversion efficiencies at substrate temperatures around 200 °C.

1. Introduction

Indium sulfide (In2S3) is one of the materials, besides CdS and Zn(O,S), which is already used as a buffer layer in chalcopyrite-type solar modules [1,2]. Due to environmental concerns and the political decision to avoid Cd in solar modules, In2S3 could be a candidate of choice to apply as a buffer layer for Cu(In,Ga)Se2 (CIGS) thin-film solar cells instead of the established CdS. InxSy layers for the application as CIGS buffers can be grown by almost all available deposition techniques, including atomic layer deposition (ALD), thermal evaporation, chemical bath deposition (CBD), ion layer gas reaction (ILGAR), spray pyrolysis, and sputtering [3,4,5]. A value of 18.2% (with anti-reflective coating) was reported by Spiering et al. for a CIGS cell with thermal evaporated InxSy buffer [6], 17.9% (certified) was achieved with a 30 cm × 30 cm module with thermal evaporated InxSy:Na buffer by the company Avancis [7], 16.4% were published for a cell with an ALD InxSy buffer [4], 16.8% (in-house) and 16.1% (certified) were achieved by Saez-Araoz et al. with an ILGAR InxSy:Cl buffer [8], and 14.3% for a 225 cm2 cell with a reactively sputtered Inx(O,S)y buffer fabricated by the company Midsummer [2].
The sputtering approach can be easily implemented in an industrial environment, similar to the sputtering of the molybdenum back contact, high-resistive layers and transparent conductive oxides used as front contact. Furthermore, the method would be favorable compared to ALD, thermal evaporation, or growth by CBD, due to the very fast deposition times of approximately one minute or less; compared to ten or even more minutes to deposit a 20–60 nm thick InxSy buffer layer. This “dry” sputtering process could be a candidate for an application in a roll-to-roll coater with a flexible substrate or the implementation in an in-line deposition machine without breaking the vacuum after CIGS absorber deposition.
We present our development of very fast rf-magnetron sputtered InxSy buffer layers deposited at different substrate temperatures, show the efficiency results of InxSy-buffered cells with open-circuit voltage (VOC) values close to reference cells with CBD-CdS buffer, and discuss the influence of substrate temperature during InxSy sputtering. In addition, a detailed materials characterization of our various InxSy buffer layers including chemical depth profiles for the analysis of the InxSy buffer and CIGS/InxSy interface reveal a sodium accumulation in the InxSy buffer and at the CIGS/buffer interface and a copper depletion of the CIGS surface with higher substrate temperatures.

2. Materials and Methods

Polycrystalline CIGS absorber layers with a thickness of 2.3 µm, a Ga/(Ga+In) ratio of 0.3 and a Cu/(Ga+In) of 0.8 were grown in an industrially relevant in-line multi-stage co-evaporation process on Mo-coated soda-lime glass [9]. No alkali post-deposition treatment was applied prior to the buffer deposition, which can significantly improve the device performance mainly due to an increased VOC.
In the next step, InxSy as the buffer layer was rf-magnetron sputtered in a VON ARDENNE (Dresden, Germany) high-vacuum sputtering system of type CS 730S from a commercially available In2S3 target with a diameter of 20 cm in a pure argon atmosphere (for further details see [10]). The substrate temperature during sputtering of the InxSy layers was increased from room temperature (RT) up to 240 °C at low power densities of approximately 1 W/cm2 compared to high power densities around 10 W/cm2 typically used for ZnO:Al sputtering [11]. Thicknesses of InxSy buffer layers for solar cells were in the range of 30–60 nm for deposition times of 40–80 s, resulting in a sputter deposition rate of approximately 0.8 nm/s. In addition, around 500 nm thick InxSy layers were deposited on highly transparent suprasil quartz glass substrates for transmittance and gracing incident X-ray diffraction (GIXRD) measurements. The thicknesses of the thin InxSy layers were determined on CIGS/InxSy cross-sections with a scanning electron microscope and for the thick layers by evaluation of the transmittance data. The commonly used CdS buffer with approximately 50 nm thickness was grown by CBD with a thiourea-based process as reference material.
Subsequently, rf-sputtered undoped ZnO (i-ZnO) was applied as high-resistive layer and dc-sputtered ZnO:Al served as front contact with Ni/Al/Ni grid fingers deposited by e-beam evaporation on top. The total cell area was 0.5 cm2. Figure 1a illustrates the stacking sequence as used in this work with a sputtered InxSy buffer in combination with i-ZnO. Alternatively we tried different stacking sequences with rf-sputtered Zn0.75Mg0.25O as high-resistive layer instead of i-ZnO and without any high-resistive layer with ZnO:Al deposited directly on top of InxSy. Both approaches turned out fine but could not significantly improve the cell efficiencies compared to our standard stacking sequence with i-ZnO, which we used for the present work.
Current–voltage (IV) curves were measured with a WACOM (Saitama, Japan) AM1.5G solar simulator with four-point geometry at standard testing conditions with a silicon reference solar cell for calibration. External quantum efficiency (EQE) measurements were performed with a BENTHAM (Berkshire, UK) PVE 300 setup with an additional light bias. For IV and EQE measurements the CIGS cells were in the as-grown state (no further light-soaking or post-annealing procedures were applied). The transmittance of 500 nm thick InxSy layers on suprasil quartz glass substrates was measured with a Perkin Elmer (Waltham, MA, USA) UV/VIS/NIR Lambda 900 spectrometer and GIXRD patterns for structural characterization were obtained with a PANalytical (Almelo, The Netherlands) Empyrian diffractometer with Cu Kα radiation with an incident angle of 0.5°.
We performed time-of-flight secondary ion mass spectroscopy (ToF-SIMS) depth profiles with a ToF5-SIMS instrument from IONTOF (Münster, Germany). A Bi+ liquid metal gun running at 30 keV generates the secondary ions, which are recorded by a time-of-flight detector. We choose an oxygen gas gun as sputtering source to detect sodium in a good sensibility. To acquire quantified depth profiles of sulfur, oxygen, and CIGS matrix elements, we selected a cesium metal gun. Both sputter guns worked at an energy of 2 keV. The size of the sputter area with 130 × 130 µm2 was larger than the analysis area of 80 × 80 µm2 to get rid of crater edge information. For the quantified measurement, we chose an analysis area of 80 × 80 µm2 and a sputter area of 500 × 500 µm2 to increase the number of data points.

3. Results

3.1. Solar Cell Characteristics

3.1.1. Comparison of Sputtered InxSy- and CBD CdS-Buffered Solar Cells

Figure 1 depicts the IV and EQE curves of the best InxSy- and CdS-buffered CIGS solar cells fabricated within this experimental campaign. The corresponding solar cell parameters are listed in the inset of Figure 1b. With 15.3%, the efficiency of the InxSy-buffered cell is reduced compared to the reference cell with CdS buffer with 17.1% efficiency—mainly due to a lower fill factor (FF). Nevertheless, the VOC is only slightly lower with 661 mV compared to 680 mV of the reference cell. There is no gain in short-circuit current density (JSC) for the InxSy-buffered cell as there is an absorption in the short wavelength region in the EQE (see Figure 1c). This feature is the result of a too-low bandgap energy and parasitic absorption due to the pronounced thickness of the InxSy layer. In addition, the EQE values of the InxSy-buffered cell are slightly below the CdS reference in the region between 500 and 1100 nm. The gradual decrease in EQE between 700 and 400 nm would tend to an indirect bandgap energy for InxSy. We observed no significant effect, neither positive nor negative, with cold-light soaking under AM1.5G conditions on solar cell parameters of CIGS solar cells with sputtered InxSy buffer.

3.1.2. Dependency of Solar Cell Efficiency on Substrate Temperature During InxSy Sputtering

Figure 2 shows the dependency of solar cell parameters on substrate temperature during sputtering of the InxSy buffer layer. Relatively low efficiency values resulted for cells with InxSy sputtered at RT whereas there is an efficiency plateau for a substrate temperature around 200 °C. For significantly higher substrate temperatures such as 240 °C there is a decrease in the efficiency values again. The difference in efficiency is mainly affected by the VOC whereas FF and JSC have minor influence. The increase in the scattering of efficiency data for solar cells fabricated with InxSy sputtered at temperatures lower and higher than 220 °C is mainly driven by the FF.

3.2. Analysis of InxSy Layers and CIGS/Buffer Interface

3.2.1. Crystal Structure and Bandgap Energy of Sputtered InxSy

Figure 3a depicts the GIXRD patterns obtained on 500 nm thick InxSy layers on glass substrates. Both layers deposited at RT and 220 °C are crystalline and exhibit the tetragonal β-In2S3 structure [PDF no. 96-400-0814]. It should be noted that the reflexes of the cubic high-temperature structure [PDF no. 96-151-8188] overlap with most of the tetragonal reflexes, and Soni et al. reported recently the formation of cubic In2S3 by sputtering [12]. Therefore, a coexistence of this cubic α-phase could not be totally excluded in our case and probably this phase could occur in case there are locally very high temperatures above 717 K [13]. Even the sputtering power (density) [14] and thickness of the InxSy layers [15] could influence the structural properties of the deposited InxSy films. The high-temperature modification γ-In2S3 which typically forms for temperatures above 1049 K, as reported by Pistor et al. [13] is very unlikely to occur during our sputtering conditions.
Figure 3b illustrates the difference in transmittance of InxSy films deposited at RT and 220 °C under otherwise identical conditions on suprasil quartz glass substrates. The evaluated thicknesses from these transmittance measurements are 490 and 480 nm for RT and 220 °C, respectively. Thus, there is only a slight reduction in the sputter deposition rate with temperature. The main finding is that the optical absorption edge slightly shifts to higher energies for the sample with the InxSy buffer deposited at 220 °C. According to a model with an indirect bandgap energy (Eg) for In2S3 the values of Eg = 2.02 eV (RT) and Eg = 2.08 eV (220 °C), respectively, were extracted by extrapolation of the linear region in the sqrt(α E)-E-plot, as illustrated in Figure 3c. In case a model with a direct bandgap Eg is used for the evaluation (not shown here), higher Eg values would result but with the same trend—a slight increase in Eg with elevated temperature. However, the corresponding (α E)²-E-plots show no distinct linear shape as expected for a semiconductor with a direct bandgap. Generally, an increase in bandgap energy might be expected as a result of the incorporation of oxygen in In2S3 [16]. In our case, this oxygen could probably stem from oxygen/water residual from the sputtering system, which could be more easily incorporated in the growing InxSy film deposited at higher substrate temperatures. Nevertheless, we could not clearly detect a higher amount of oxygen inside the InxSy layer grown at higher temperatures as reported in Section 3.2.2. A more likely explanation could be changes in chemical composition, i.e., in the sulfur/indium ratio as also shown Section 3.2.2.
As a conclusion, only slight differences in structural and optical material properties between both In2S3 layers sputtered at RT and 220 °C were observed, which is an indication that pronounced changes in solar cell performance could be more affected by the interface formation between CIGS absorber and InxSy buffer layer.

3.2.2. Chemical Composition of InxSy/CIGS Interface Region

Depth profiles were performed by ToF-SIMS in order to get an idea what influences the efficiency and VOC and to reveal changes in the chemical composition of the InxSy buffer and CIGS/buffer interface due to the different substrate temperatures during InxSy sputtering.
The quantified ToF-SIMS depth profiles shown in Figure 4 highlight changes in the chemical composition of the InxSy buffer layer deposited on CIGS. The chemical composition of InxSy grown at RT is stoichiometric, i.e., the layer has a homogeneous In2S3 composition over the whole 60 nm thickness; whereas for substrate temperatures higher than 200 °C, the InxSy composition becomes more sulfur-rich and indium-deficient, and the layers tend to be more inhomogenous. ToF-SIMS depth profiles obtained on samples sputtered at 160 and 180 °C (omitted here for clarity) reflect the same trend as the RT sample.
Since pure argon without any addition of oxygen was used as sputtering gas, the oxygen concentration in all layers is expectedly very low, i.e., the S/(S+O) ratio is close to 1, indicating that no oxygen is incorporated during the sputtering process even at the higher substrate temperatures of 220 and 240 °C. Small amounts of oxygen are only detected at the InxSy surface.
In addition to changes in the composition of the InxSy buffer, there is a pronounced copper depletion of the CIGS surface, which is more obvious for layers deposited at higher substrate temperatures of 220 and 240 °C compared to RT, 160, 180, and 200 °C. In a more detailed consideration, the copper depletion at the CIGS surface was estimated to rise from around 10 nm at RT, up to approximately 70 nm at 240 °C.
ToF-SIMS depth profiles were also performed on complete CIGS solar cells from ZnO:Al front contact down to the soda-lime glass substrate to reveal differences in chemical composition between CIGS solar cells with efficiencies of 7.1% and 14.6% with InxSy buffer layers sputtered at RT and at 220 °C, respectively (Figure 5). The depth profile for the sample fabricated at higher substrate temperature exhibits a strong increase in the sodium signal inside the InxSy buffer layer and at the CIGS/buffer interface (note the logarithmic scale here). The sodium is originally diffusing from the soda-lime glass substrate through the molybdenum back contact into the CIGS layer at elevated deposition temperatures during the CIGS co-evaporation process but was not intentionally incorporated into the CIGS absorber or InxSy buffer layer.
The observed sodium accumulation combined with a moderate copper depletion at the CIGS/InxSy interface region could be the most significant effect for an efficiency improvement of In2S3 buffered CIGS-based solar cells.

4. Discussion

We have demonstrated that we could achieve suitable film quality for InxSy buffer layers with the very fast deposition method of rf-magnetron sputtering applied to CIGS thin-film solar cells to realize efficiencies above 15%. For our sputtered buffer layers, we applied high sputtering deposition rates around 0.8 nm/s; compared to slower deposition processes such as ALD, CBD, ILGAR, and thermal evaporation. Nevertheless, there is still an efficiency gap compared to the CdS-buffered reference cells, which was also observed in our previous works with sputtered InxSy buffers [10,17] and reported by other groups for CIGS cells with sputtered InxSy buffers [12,18,19], evaporated InxSy buffers [20,21], and our own cells with InxSy layers grown by metal-organic chemical vapor deposition [22].
For both 500 nm thick InxSy layers deposited at RT and 220 °C substrate temperature on glass substrates, we obtained crystalline structures without any huge difference between both layers. We indicated the tetragonal β-phase for In2S3. This result is in accordance with structural investigations on rf-sputtered InxSy layers reported by other groups [14,15,23], but also for structural investigations on evaporated InxSy layers [20,24]. In contrast, high-resolution transmission electron microscopy analyses revealed nanocrystalline structures for our oxygen-containing CBD-Inx(O,S)y [25] and sputtered Inx(O,S)y buffer layers deposited from an In2(O0.25S0.75)3 target [26] at the CIGS/buffer interface in the complete device.
The extracted bandgap energy (indirect) for our InxSy is in the range of 2.0–2.1 eV, thus not perfectly ideal for an application as a buffer in CIGS thin-film solar cells. With the addition of dopants like chlorine [8], oxygen [16], or sodium [1], the bandgap energy can be increased to suitable values to reduce parasitic absorption of the buffer and obtain high JSC. In addition, a beneficial band alignment with chalcopyrite-based absorber is necessary to obtain high VOC [27] and close the efficiency gap to the CdS-buffered reference cells. In this sense, the surface termination of the CIGS absorber could play an important role. The parasitic absorption can be also reduced by using thinner InxSy buffers with thicknesses below 20 nm. In this case the subsequent high-resistive layer becomes more important in terms of band alignment. It should be noted that the power density during sputtering could influence the bandgap value of InxSy layers as well [14]. Our extracted bandgap energy increases slightly around 0.06 eV with the substrate temperature during InxSy sputtering. This trend with increasing temperature was also observed by other groups for sputtered InxSy layers [17,18,28]. In contrast, Karthikeyan et al. reported an opposite trend as found for their InxSy deposited by pulsed direct magnetron sputtering [23]. From theoretical density calculations there is only a very small difference between direct and indirect transitions in β-In2S3 [29,30].
The substrate temperature during InxSy sputtering has a major influence on the performance of the device. Solar cells with InxSy sputtered at RT result in poor efficiencies whereas a maximum seems to be around 200 °C. For higher temperatures >240 °C, there is a decrease in efficiency and VOC again. The strong dependency of solar cell efficiency on substrate temperature during sputtering and sample pre-conditioning was also observed in previous experiments at ZSW [10,17] and by other groups [31]. For higher substrate temperatures above 220 °C there seems to be a destruction of the important pn-junction at the CIGS/buffer interface resulting in deteriorated solar cell parameters.
In addition, the substrate temperature during InxSy sputtering has an influence on the chemical composition of the InxSy buffer and CIGS/buffer interface. In this work, we observed an increasing sulfur/indium ratio starting from a value of 3:2 (In2S3) grown at RT with increasing substrate temperature during sputtering, similar to the trend in our previous experiments, but with a different starting ratio of around 1:1 [17], and similar to results obtained by D. Abou-Ras et al. [31].
The sodium accumulation in the InxSy buffer and at the CIGS/buffer interface and the copper-depleted CIGS surface seem to be the most likely candidates to increase the VOC of the 220 °C cell compared to the RT cell. Sodium addition to CIGS in general increases the VOC, e.g., in case a sodium-free substrate is used and sodium is intentionally incorporated with a precursor [32] or post-deposition process [33]. Sodium accumulation in the buffer and at the CIGS/buffer interface was also found after post-annealing processes for CBD Zn(O,S)-buffered devices [34]. A sodium accumulation on the nano scale was observed by atom probe tomography measurements at the CIGS/InxSy interface for samples annealed at temperatures above 225 °C compared to the RT samples without sodium accumulation [12]. The observed copper depletion at the CIGS surface could result in an ordered vacancy compound with a slightly higher bandgap due to a lowered valence band maximum. This feature could promote an improved band alignment in combination with the InxSy buffer and could be also a reason for improved VOC values as reported for evaporated InxSy:Na buffer on Cu(In,Ga)(S,Se)2 [27]. The formation of such a Cu-poor CuIn5S8 composition at the CIGS/sputtered InxSy interface was observed at relatively high temperatures around 340 °C [31]. In general, the integral copper content of the CIGS absorber can significantly influence the efficiency and the dominating recombination mechanism of InxSy-buffered devices [35].
The phenomena of copper diffusion at the CIGS/In2S3 interface, when investigated, was mostly isolated from alkali elements in the past. Pistor et al. clearly showed with high kinetic energy X-ray photoelectron spectroscopy that above 200 °C, the copper diffusion from CIGS into In2S3 is enhanced [36]. Recent calculations with density functional theory also took sodium into account and revealed an interplay between sodium and copper diffusion at the CIGS/In2S3 interface, and the formation of Na- and Cu-containing phases is very likely due to strong negative driving force energies (energy releasing processes). The even lower driving force energy value for sodium implies that sodium diffusion is more favored than copper diffusion and would therefore be actuated earlier [37]. This difference could be one explanation for our observed results from the depth profiles: First, there is a sodium diffusion up to 200 °C with only small changes in copper. For higher temperatures >220 °C the copper diffusion is no longer hindered by the sodium and copper diffusion takes place. Following this argumentation, the important parameter for an increased efficiency is the sodium accumulation at the CIGS/buffer interface and a moderate copper depletion at the CIGS surface. An enhanced copper enrichment of the In2S3 layer would be detrimental. It should be noted that, besides the copper depletion of the CIGS surface, an intermixing between InxSy buffer and CIGS absorber could also take place. A possible sulfur incorporation into the CIGS absorber, which we did not observe clearly in the ToF-SIMS profiles, would lead to a lowered valence band maximum and a different band alignment to InxSy.
Future research directions could address one or several of the following approaches: Sodium can be intentionally inserted in the In2S3 target to increase the bandgap energy, similar to the sodium-containing InxSy buffer used by the company Avancis deposited by thermal evaporation [7]. In addition, post-annealing procedures or heat-light soaking could be an adequate approach to diffuse sodium into the InxSy buffer and to the CIGS/buffer interface, which was already successful applied to CIGS cells with CBD-Zn(O,S)/(Zn,Mg)O buffer system. Oxygen can be intentionally introduced via reactive sputtering with oxygen in the sputtering gas [2], via variation of the base pressure with water vapor as oxygen source [18], the use of mixed In2(O,S)3 targets, or co-sputtering from pure In2S3 and In2O3 targets. Another approach could be the use of a very thin InxSy layer with a thickness of 20 nm or even less. In this case the subsequent high-resistive layer should be optimized, e.g., (Zn,Mg)O or ZnTiOx [38] could be candidates to replace i-ZnO. Furthermore, different window layers—such as InZnOx or InSnOx instead of ZnO:Al—can influence the device performance of CIGS solar cells in combination with sputtered Inx(O,S)y buffers [2].

5. Conclusions

The substrate temperature during the sputtering of InxSy buffer layers significantly influences the performance of CIGS thin-film solar cells. The best power conversion efficiencies in the as-grown state were obtained at a temperature plateau around 200 °C with the best value of 15.3% compared to the reference cell with solution-grown CdS buffer with 17.1% efficiency. The difference in performance compared to cells with buffers deposited below 200 °C is mainly due to a reduced VOC. The chemical composition and homogeneity of InxSy changes from stoichiometric and homogenous In2S3 for layers deposited RT to inhomogenous and more sulfur-rich and indium-deficient composition for higher temperatures in the range of 220–240 °C. In addition, there is a pronounced copper depletion at the CIGS absorber surface and a sodium accumulation in the InxSy buffer and at the CIGS/buffer interface for higher temperatures compared to InxSy layers deposited at RT. These two features could explain the higher VOC values of cells with InxSy buffers grown at higher substrate temperatures compared to RT. However, InxSy layers with a non-ideal bandgap energy of 2.0–2.1 eV (indirect) induce parasitic losses and a limited JSC. In future research, appropriate quantities of sodium and/or oxygen could be intentionally incorporated in the InxSy layer and—with very thin InxSy layers, in combination with an optimized high-resistive layer—the parasitic absorption could be further reduced and the JSC increased.

Author Contributions

Conceptualization, D.H. and W.W.; validation, D.H., W.H., R.M. and W.W.; investigation, D.H., W.H., R.M., and W.W; writing—original draft preparation, W.W.; writing—review and editing, D.H., W.H., R.M., and W.W.; visualization, D.H.; project administration, W.W.; funding acquisition, W.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the German Federal Ministry for Economic Affairs and Energy (BMWi) within the “EFFCIS” project under contract number 0324076A.

Acknowledgments

We thank the CIGS team at ZSW for solar cell preparation, especially Stefan Paetel for CIGS deposition.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Palm, J.; Dalibor, T.; Lechner, R.; Pohlner, S.; Verma, R.; Dietmüller, R.; Heiß, A.; Vogt, H.; Karg, F. Cd-free CIS thin film solar modules at 17% efficiency. In Proceedings of the 29th European Photovoltaic Solar Energy Conference, Amsterdam, The Netherlands, 22–26 September 2014; pp. 1433–1438. [Google Scholar]
  2. Niemi, E.; Sterner, J.; Carlsson, P.; Oliv, J.; Jaremalm, E.; Lindström, S. All-sputtered flexible CIGS cells at high speed. In Proceedings of the 31st European Photovoltaic Solar Energy Conference, Hamburg, Germany, 14–18 September 2015; pp. 1010–1013. [Google Scholar]
  3. Barreau, N. Indium sulfide and relatives in the world of photovoltaics. Sol. Energy 2009, 83, 363–371. [Google Scholar] [CrossRef]
  4. Naghavi, N.; Abou-Ras, D.; Allsop, N.; Barreau, N.; Bücheler, S.; Ennaoui, A.; Fischer, C.-H.; Guillen, C.; Hariskos, D.; Herrero, J.; et al. Buffer layers and transparent conducting oxides for chalcopyrite Cu(In,Ga)(S,Se)2 based thin film photovoltaics: Present status and current developments. Prog. Photovolt. Res. Appl. 2010, 18, 411–433. [Google Scholar] [CrossRef]
  5. Mughal, M.A.; Engelken, R.; Sharma, R. Progress in indium (III) sulfide (In2S3) buffer layer deposition techniques for CIS, CIGS, and CdTe-based thin film solar cells. Sol. Energy 2015, 120, 131–146. [Google Scholar] [CrossRef]
  6. Spiering, S.; Nowitzki, A.; Kessler, F.; Igalson, M.; Maksoud, H.A. Optimization of buffer-window layer system for CIGS thin film devices with indium sulphide buffer by in-line evaporation. Sol. Energy Mater. Sol. Cells 2016, 144, 544–550. [Google Scholar] [CrossRef]
  7. Dalibor, T.; Eraerds, P.; Grave, M.; Algasinger, M.; Visbeck, S.; Niesen, T.; Palm, J. Advanced PVD buffers on the road to GW-scale CIGSSe production. In Proceedings of the 43rd IEEE Photovoltaics Specialists Conference, Portland, OR, USA, 5–10 June 2016; pp. 1433–1437. [Google Scholar]
  8. Sáez-Araoz, R.; Krammer, J.; Harndt, S.; Koehler, T.; Krueger, M.; Pistor, P.; Jasenek, A.; Hergert, F.; Lux-Steiner, M.C.; Fischer, C.-H. ILGAR In2S3 buffer layers for Cd-free Cu(In,Ga)(S,Se)2 solar cells with certified efficiencies above 16%. Prog. Photovolt. Res. Appl. 2012, 20, 855–861. [Google Scholar] [CrossRef] [Green Version]
  9. Voorwinden, G.; Kniese, R.; Jackson, P.; Powalla, M. In-Line Cu(In,Ga)Se2 co-evaporation process on 30 cm × 30 cm substrates with multiple deposition stages. In Proceedings of the 22nd European Photovoltaic Solar Energy Conference, Milan, Italy, 3–7 September 2007; pp. 2115–2118. [Google Scholar]
  10. Hariskos, D.; Menner, R.; Spiering, S.; Eicke, A.; Powalla, M.; Ellmer, K.; Oertel, M.; Dimmler, B. In2S3 Buffer layer deposited by magnetron sputtering for Cu(In,Ga)Se2 solar cells. In Proceedings of the 19th European Photovoltaic Solar Energy Conference, Paris, France, 7–11 June 2004; pp. 1894–1897. [Google Scholar]
  11. Kluth, O.; Schöpe, G.; Rech, B.; Menner, R.; Oertel, M.; Orgassa, K.; Schock, H.W. Comparative material study on RF and DC magnetron sputtered ZnO:Al films. Thin Solid Films 2006, 502, 311–316. [Google Scholar] [CrossRef]
  12. Soni, P.; Raghuwanshi, M.; Wuerz, R.; Berghoff, B.; Knoch, J.; Raabe, D.; Cojocaru-Mirédin, O. Role of elemental intermixing at the In2S3/CIGSe heterojunction deposited using reactive RF magnetron sputtering. Sol. Energy Mater. Sol. Cells 2019, 195, 367–375. [Google Scholar] [CrossRef]
  13. Pistor, P.; Merino Álvarez, J.M.; León, M.; di Michiel, M.; Schorr, S.; Klenk, R.; Lehmann, S. Structure reinvestigation of α-, β- and γ-In2S3. Acta Crystallogr. B Struct. Sci. Cryst. Eng. Mater. 2016, B72, 410–415. [Google Scholar] [CrossRef] [Green Version]
  14. Hwang, D.H.; Cho, S.; Hui, K.N.; Son, Y.G. Effect of sputtering power on structural and optical properties of radio frequency-sputtered in2S3 thin films. J. Nanosci. Nanotechnol. 2014, 14, 8978–8981. [Google Scholar] [CrossRef]
  15. Ji, Y.; Ou, Y.; Yu, Z.; Yan, Y.; Wang, D.; Yan, C.; Liu, L.; Zhang, Y.; Zhao, Y. Effect of film thickness on physical properties of RF sputtered In2S3 layers. Surf. Coat. Technol. 2015, 276, 587–594. [Google Scholar] [CrossRef]
  16. Barreau, N.; Bernède, J.C.; Marsillac, S.; Mokrani, A. Study of low temperature elaborated tailored optical band gap β-In2S3−3xO3x thin films. J. Cryst. Growth 2002, 235, 439–449. [Google Scholar] [CrossRef]
  17. Hariskos, D.; Menner, R.; Lotter, E.; Spiering, S.; Powalla, M. Magnetron sputtering of indium sulphide as the buffer layer in Cu(In,Ga)Se2-based solar cells. In Proceedings of the 20th European Photovoltaic Solar Energy Conference, Barcelona, Spain, 6–10 June 2005; pp. 1713–1716. [Google Scholar]
  18. Ho, W.-H.; Hsu, C.-H.; Wei, S.-Y.; Cai, C.-H.; Huang, W.-C.; Lai, C.-H. Sputtered Inx(O,S)y buffer layers for Cu(In,Ga)Se2 thin-film solar cells: Engineering of band alignment and interface properties. ACS Appl. Mater. Interfaces 2017, 9, 17586–17594. [Google Scholar] [CrossRef] [PubMed]
  19. Soni, P.; Raghuwanshi, M.; Wuerz, R.; Berghoff, B.; Knoch, J.; Raabe, D.; Cojocaru-Mirédin, O. Sputtering as a viable route for In2S3 buffer layer deposition in high efficiency Cu(In,Ga)Se2 solar cells. Energy Sci. Eng. 2019, 7, 478–487. [Google Scholar] [CrossRef] [Green Version]
  20. Pistor, P.; Caballero, R.; Hariskos, D.; Izquierdo-Roca, V.; Wächter, R.; Schorr, S.; Klenk, R. Quality and stability of compound indium sulphide as source material for buffer layers in Cu(In,Ga)Se2 solar cells. Sol. Energy Mater. Sol. Cells 2009, 93, 148–152. [Google Scholar] [CrossRef]
  21. Verma, R.; Datta, D.; Chirila, A.; Güttler, D.; Perrenoud, J.; Pianezzi, F.; Müller, U.; Kumar, S.; Tiwari, A.N. Optical, structural, and chemical properties of flash evaporated In2S3 buffer layer Cu(In, Ga)Se2 for solar cells. J. Appl. Phys. 2010, 108, 074904. [Google Scholar] [CrossRef] [Green Version]
  22. Spiering, S.; Bürkert, L.; Hariskos, D.; Powalla, M.; Dimmler, B.; Giesen, C.; Heuken, M. MOCVD indium sulphide for application as a buffer layer in CIGS solar cells. Thin Solid Films 2009, 517, 2328–2331. [Google Scholar] [CrossRef]
  23. Karthikeyan, S.; Hill, A.E.; Pilkington, R.D. Low temperature pulsed direct current magnetron sputtering technique for single phase β-In2S3 buffer layers for solar cell applications. Appl. Surf. Sci. 2017, 418, 199–206. [Google Scholar] [CrossRef]
  24. Barreau, N.; Marsillac, S.; Albertini, D.; Bernede, J.C. Structural, optical and electrical properties of β-In2S3−3xO3x thin films obtained by PVD. Thin Solid Films 2002, 403–404, 331–334. [Google Scholar] [CrossRef]
  25. Jin, X.; Popescu, R.; Pasha, A.; Schneider, R.; Hariskos, D.; Witte, W.; Powalla, M.; Gerthsen, D. Structural and microchemical characterization of Cu(In,Ga)Se2 solar cells with solution-grown CdS, Zn(O,S), and Inx(O,S)y buffers. Thin Solid Films 2019, 671, 133–138. [Google Scholar] [CrossRef]
  26. Jin, X.; Schneider, R.; Popescu, R.; Hariskos, D.; Witte, W.; Powalla, M.; Gerthsen, D. Characterization of solution-grown and sputtered Inx(O,S)y buffer layers in Cu(In,Ga)Se2 solar cells by analytical TEM. J. Semicond. Sci. Technol. 2020, 35, 034001. [Google Scholar] [CrossRef]
  27. Hauschild, D.; Meyer, F.; Benkert, A.; Kreikemeyer-Lorenzo, D.; Dalibor, T.; Palm, J.; Blum, M.; Yang, W.; Wilks, R.G.; Bär, M.; et al. Improving performance by Na doping of a buffer layer—Chemical and electronic structure of the InxSy:Na/CuIn(S,Se)2 thin-film solar cell interface. Prog. Photovolt. Res. Appl. 2018, 26, 359–366. [Google Scholar] [CrossRef]
  28. Siol, S.; Dhakal, T.P.; Gudavalli, G.S.; Rajbhandari, P.P.; DeHart, C.; Baranowski, L.L.; Zakutayev, A. Combinatorial reactive sputtering of In2S3 as an alternative contact layer for thin film solar cells. ACS Appl. Mater. Interfaces 2016, 8, 14004–14011. [Google Scholar] [CrossRef] [PubMed]
  29. Sharma, Y.; Srivastava, P. Electronic, optical and transport properties of α-, β- and γ-phases of spinel indium sulphide: An ab initio study. Mater. Chem. Phys. 2012, 135, 385–394. [Google Scholar] [CrossRef]
  30. Ghorbani, E.; Erhart, P.; Albe, K. Energy level alignment of Cu(In,Ga)(S,Se)2 absorber compounds with In2S3, NaIn5S8, and CuIn5S8 Cd-free buffer materials. Phys. Rev. Mater. 2019, 3, 075401. [Google Scholar] [CrossRef]
  31. Abou-Ras, D.; Kostorz, G.; Hariskos, D.; Menner, R.; Powalla, M.; Schorr, S.; Tiwari, A.N. Structural and chemical analyses of sputtered InxSy buffer layers in Cu(In,Ga)Se2 thin-film solar cells. Thin Solid Films 2009, 517, 2792–2798. [Google Scholar] [CrossRef]
  32. Caballero, R.; Kaufmann, C.A.; Eisenbarth, T.; Unold, T.; Klenk, R.; Schock, H.-W. High efficiency low temperature grown Cu(In,Ga)Se2 thin film solar cells on flexible substrates using NaF precursor layers. Prog. Photovolt. Res. Appl. 2011, 19, 547–551. [Google Scholar] [CrossRef]
  33. Rudmann, D.; da Cunha, A.F.; Kaelin, M.; Kurdesau, F.; Zogg, H.; Tiwari, A.N.; Bilger, G. Efficiency enhancement of Cu(In,Ga)Se2 solar cells due to post-deposition Na incorporation. Appl. Phys. Lett. 2004, 84, 1129–1131. [Google Scholar] [CrossRef]
  34. Witte, W.; Hariskos, D.; Eicke, A.; Menner, R.; Kiowski, O.; Powalla, M. Impact of annealing on Cu(In,Ga)Se2 solar cells with Zn(O,S)/(Zn,Mg)O buffers. Thin Solid Films 2013, 535, 180–183. [Google Scholar] [CrossRef]
  35. Spiering, S.; Paetel, S.; Kessler, F.; Igalson, M.; Maksoud, H.A. Copper variation in Cu(In,Ga)Se2 solar cells with indium sulphide buffer layer. Thin Solid Films 2015, 582, 328–331. [Google Scholar] [CrossRef]
  36. Pistor, P.; Allsop, N.; Braun, W.; Caballero, R.; Camus, C.; Fischer, C.-H.; Gorgoi, M.; Grimm, A.; Johnson, B.; Kropp, T.; et al. Cu in In2S3: Interdiffusion phenomena analysed by high kinetic energy X-ray photoelectron spectroscopy. Phys. Status Solidi A 2009, 5, 1059–1062. [Google Scholar] [CrossRef]
  37. Ghorbani, E.; Albe, K. Influence of Cu and Na incorporation on the thermodynamic stability and electronic properties of β-In2S3. J. Mater. Chem. C 2018, 6, 7226–7231. [Google Scholar] [CrossRef]
  38. Löckinger, J.; Nishiwaki, S.; Andres, C.; Erni, R.; Rossell, M.D.; Romanyuk, Y.E.; Buecheler, S.; Tiwari, A.N. ALD-ZnxTiyO as window layer in Cu(In,Ga)Se2 solar cells. ACS Appl. Mater. Interfaces 2018, 10, 43603–43609. [Google Scholar]
Figure 1. (a) Scheme of Cu(In,Ga)Se2 (CIGS) thin-film solar cell with rf-sputtered InxSy buffer layer (not to scale); (b) IV curves of CIGS solar cells either with rf-mangetron sputtered InxSy buffer at 220 °C substrate temperature (red) or with a CdS buffer layer(black) grown by chemical bath deposition (CBD), both with i-ZnO/ZnO:Al on top. Corresponding solar cell parameters are listed in the inset. JSC values are corrected with EQE; (c) External quantum efficiency (EQE) curves of the same cells.
Figure 1. (a) Scheme of Cu(In,Ga)Se2 (CIGS) thin-film solar cell with rf-sputtered InxSy buffer layer (not to scale); (b) IV curves of CIGS solar cells either with rf-mangetron sputtered InxSy buffer at 220 °C substrate temperature (red) or with a CdS buffer layer(black) grown by chemical bath deposition (CBD), both with i-ZnO/ZnO:Al on top. Corresponding solar cell parameters are listed in the inset. JSC values are corrected with EQE; (c) External quantum efficiency (EQE) curves of the same cells.
Applsci 10 01052 g001
Figure 2. Solar cell parameters of CIGS thin-film solar cells with 60 nm thick InxSy buffer layers sputtered at different substrate temperatures. For each substrate temperature 20 solar cells were measured: (a) Power conversion efficiency η; (b) Open circuit voltage VOC; (c) Fill factor FF; (d) Short-circuit current density JSC.
Figure 2. Solar cell parameters of CIGS thin-film solar cells with 60 nm thick InxSy buffer layers sputtered at different substrate temperatures. For each substrate temperature 20 solar cells were measured: (a) Power conversion efficiency η; (b) Open circuit voltage VOC; (c) Fill factor FF; (d) Short-circuit current density JSC.
Applsci 10 01052 g002
Figure 3. (a) GIXRD patterns measured with an incident angle of 0.5° on approximately 500 nm thick InxSy layers on suprasil quartz glass; (b) Transmittance measurement of about 500 nm thick InxSy layers on suprasil quartz glass; (c) Corresponding sqrt(α E)-E-plot (indirect bandgap energy for extraction of Eg value was used).
Figure 3. (a) GIXRD patterns measured with an incident angle of 0.5° on approximately 500 nm thick InxSy layers on suprasil quartz glass; (b) Transmittance measurement of about 500 nm thick InxSy layers on suprasil quartz glass; (c) Corresponding sqrt(α E)-E-plot (indirect bandgap energy for extraction of Eg value was used).
Applsci 10 01052 g003
Figure 4. Quantified ToF-SIMS depth profiles performed with cesium cluster source for sputtering on InxSy/CIGS samples deposited at different substrate temperatures (Tsubstrate) during InxSy sputtering. The x-axes are normalized for clarity. Nominal thickness of InxSy layers is 60 nm and sputtering inside CIGS was carried out within the first 100 nm. Changes in InxSy chemical composition are highlighted with red arrows and copper depletion of CIGS absorber is marked with blue circles.
Figure 4. Quantified ToF-SIMS depth profiles performed with cesium cluster source for sputtering on InxSy/CIGS samples deposited at different substrate temperatures (Tsubstrate) during InxSy sputtering. The x-axes are normalized for clarity. Nominal thickness of InxSy layers is 60 nm and sputtering inside CIGS was carried out within the first 100 nm. Changes in InxSy chemical composition are highlighted with red arrows and copper depletion of CIGS absorber is marked with blue circles.
Applsci 10 01052 g004
Figure 5. ToF-SIMS depth profiles performed with oxygen sputtering source on complete CIGS solar cells with stacking sequence ZnO:Al/i-ZnO/InxSy/CIGS/Mo/glass. (a) Solar cell with 7.1% efficiency and InxSy buffer deposited at RT; (b) Solar cell with 14.6% efficiency and InxSy sputtered at 220 °C substrate temperature. The enhanced sodium signal is highlighted with a dashed red circle.
Figure 5. ToF-SIMS depth profiles performed with oxygen sputtering source on complete CIGS solar cells with stacking sequence ZnO:Al/i-ZnO/InxSy/CIGS/Mo/glass. (a) Solar cell with 7.1% efficiency and InxSy buffer deposited at RT; (b) Solar cell with 14.6% efficiency and InxSy sputtered at 220 °C substrate temperature. The enhanced sodium signal is highlighted with a dashed red circle.
Applsci 10 01052 g005

Share and Cite

MDPI and ACS Style

Hariskos, D.; Hempel, W.; Menner, R.; Witte, W. Influence of Substrate Temperature during InxSy Sputtering on Cu(In,Ga)Se2/Buffer Interface Properties and Solar Cell Performance. Appl. Sci. 2020, 10, 1052. https://doi.org/10.3390/app10031052

AMA Style

Hariskos D, Hempel W, Menner R, Witte W. Influence of Substrate Temperature during InxSy Sputtering on Cu(In,Ga)Se2/Buffer Interface Properties and Solar Cell Performance. Applied Sciences. 2020; 10(3):1052. https://doi.org/10.3390/app10031052

Chicago/Turabian Style

Hariskos, Dimitrios, Wolfram Hempel, Richard Menner, and Wolfram Witte. 2020. "Influence of Substrate Temperature during InxSy Sputtering on Cu(In,Ga)Se2/Buffer Interface Properties and Solar Cell Performance" Applied Sciences 10, no. 3: 1052. https://doi.org/10.3390/app10031052

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