research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206
Volume 72| Part 3| June 2016| Pages 410-415

Structure reinvestigation of α-, β- and γ-In2S3

CROSSMARK_Color_square_no_text.svg

aHelmholtz-Zentrum Berlin, Germany, bApplied Physics Department, Universidad Autónoma de Madrid, Spain, cESRF – The European Synchrotron, Grenoble, France, dFreie Universität Berlin, Berlin, Germany, and eSolid State Physics, Lund University, Sweden
*Correspondence e-mail: ppistor@irec.cat

Edited by P. Bordet, Institut Néel, France (Received 27 January 2016; accepted 26 April 2016; online 26 May 2016)

Semiconducting indium sulfide (In2S3) has recently attracted considerable attention as a buffer material in the field of thin film photovoltaics. Compared with this growing interest, however, detailed characterizations of the crystal structure of this material are rather scarce and controversial. In order to close this gap, we have carried out a reinvestigation of the crystal structure of this material with an in situ X-ray diffraction study as a function of temperature using monochromatic synchrotron radiation. For the purpose of this study, high quality polycrystalline In2S3 material with nominally stoichiometric composition was synthesized at high temperatures. We found three modifications of In2S3 in the temperature range between 300 and 1300 K, with structural phase transitions at temperatures of 717 K and above 1049 K. By Rietveld refinement we extracted the crystal structure data and the temperature coefficients of the lattice constants for all three phases, including a high-temperature trigonal γ-In2S3 modification.

1. Introduction

In2S3 is a widegap semiconductor with high photoconductive and photoluminescent properties, which makes it a promising material for optoelectronic applications (Shazly et al., 1998[Shazly, A. A. E., Elhady, D., Metwally, H. & Seyam, M. (1998). J. Phys. Condens. Matter, 10, 5943-5954.]). Most prominently, its potential application as a buffer layer in chalcopyrite solar cells has triggered an increased research effort in its fundamental materials properties (e.g. crystal structure, optical properties, electronic bandstructure etc.) as well as in deposition technology. The compatibility with various thin film deposition methods make it a versatile alternative to the commonly applied CdS buffer layer. Among the compatible deposition methods, atomic layer deposition (Naghavi et al., 2003[Naghavi, N., Spiering, S., Powalla, M., Cavana, B. & Lincot, D. (2003). Prog. Photovolt. Res. Appl. 11, 437-443.]), the ion layer gas reaction (ILGAR) method (Sáez-Araoz et al., 2012[Sáez-Araoz, R., Krammer, J., Harndt, S., Koehler, T., Krueger, M., Pistor, P., Jasenek, A., Hergert, F., Lux-Steiner, M. & Fischer, C. (2012). Prog. Photovolt. Res. Appl. 20, 855-861.]), spray pyrolisis (John et al., 2005[John, T. T., Mathew, M., Sudhakartha, C., Vijayakumar, K. P., Abe, T. & Kashiwaba, Y. (2005). Solar Energy Mater. Solar Cells, 89, 27-36.]), sputtering (Hariskos et al., 2005[Hariskos, D., Menner, R., Lotter, E., Spiering, S. & Powalla, M. (2005). 20th European Photovoltaic Solar Energy Conference, pp. 1713-1716. Barcelona, Spain.]) and evaporation (Strohm et al., 2005[Strohm, A., Eisenmann, L., Gebhardt, R., Harding, A., Schlötzer, T., Abou-Ras, D. & Schock, H. (2005). Thin Solid Films, 480-481, 162-167.]) have been successfully applied. The interested reader is referred here to the excellent review by Barreau (2009[Barreau, N. (2009). Solar Energy, 83, 363-371.]) on the role of In2S3  in the world of photovoltaics. Various reports on In2S3 buffer layers correlate deposition process parameters with crystallographic properties (Rao & Kumar, 2012[Rao, P. & Kumar, S. (2012). Thin Solid Films, 524, 93-99.]; Larina et al., 2004[Larina, L., Kim, K. H., Yoon, K. H., Konagai, M. & Ahn, B. T. (2004). J. Electrochem. Soc. 151, C789-C792.]; Yoosuf & Jayaraj, 2005[Yoosuf, R. & Jayaraj, M. (2005). Solar Energy Mater. Solar Cells, 89, 85-94.]) and ultimately with final solar cell device parameters (Naghavi et al., 2003[Naghavi, N., Spiering, S., Powalla, M., Cavana, B. & Lincot, D. (2003). Prog. Photovolt. Res. Appl. 11, 437-443.]; Pistor, Caballero et al., 2009[Pistor, P., Caballero, R., Hariskos, D., Izquierdo-Roca, V., Wächter, R., Schorr, S. & Klenk, R. (2009). Solar Energy Mater. Solar Cells, 93, 148-152.]).

For a correct interpretation of the crystallographic data, a comprehensive understanding of the relevant crystal structure modifications of In2S3  is mandatory. Data reported on the phase labelling and temperature sequence in the In–S system is contradictory. In view of recent technological and scientific interest in In2S3 and the commonly drawn connection to its crystal structure properties, in this contribution we therefore report about a crystal structure reinvestigation of the In2S3 system in the temperature range from room temperature up to 1322 K, close to the melting point at 1363 K (Diehl et al., 1976[Diehl, R., Carpentier, C.-D. & Nitsche, R. (1976). Acta Cryst. B32, 1257-1260.]). The present study results in high quality reference powder diffraction data sets and enhanced knowledge on the different structure modifications of In2S3, which is expected to have a direct impact on the technological understanding in terms of e.g. material quality or diffusion parameters.

The first to describe the crystal structure of In2S3 were Hahn & Klingler (1949[Hahn, H. & Klingler, W. (1949). Z. Anorg. Chem. 260, 97-109.]). They reported a cubic phase at temperatures below 600 K, which they called α-In2S3 and a transition to a tetragonal spinel-like high-temperature modification which they called β-In2S3. They already stressed the similarity between both modifications and suggested an ordering of the In atoms to be the main difference between the two phases. Later studies revealed that the cubic phase is in fact the higher-temperature phase and the stoichiometric phase existing at room temperature is tetragonal. However, for sulfur-deficient In2S3 the temperature range for the cubic phase extends down to room temperature. So Hahn probably measured a sulfur-deficient cubic α-In2S3 at room temperature and called it the lower temperature modification and from here most confusion about the labelling and temperature sequence of phases arises.

While the majority of authors follow Hahn in their nomenclature resulting in an ordering of the phases from low to high temperature as βαγ, some have relabeled the cubic and tetragonal phase with a resulting order of αβγ. Depending on the author, β-In2S3 in recent publications may therefore refer either to the low-temperature tetragonal phase or the cubic high-temperature phase. Although αβγ would be the logical order, we will follow the majority in the literature and assign β-In2S3 to the low-temperature, tetragonal phase.

There are some more and sometimes contradictory publications in the literature on the crystal structure of the different In2S3 phases, of which the most relevant ones will be shortly introduced in the following. Gödecke & Schubert (1985[Gödecke, T. & Schubert, K. (1985). Z. Metallkdd. 76, 358-364.]) suggest a phase diagram in which for the In:S ratio of 2:3 three modifications (β-In2S3, α-In2S3 and γ-In2S3) of indium sulfide exist in three different temperature regimes. King (1962[King, G. S. D. (1962). Acta Cryst. 15, 512.]) determined the space group of the tetragonal β-In2S3 as I41/amd (space group No. 141) and the lattice parameters to a0 = 7.61 Å and c0 = 32.24 Å using Weissenberg photographs. The space group was later confirmed by Goodyear & Steigmann (1961[Goodyear, J. & Steigmann, G. (1961). Proc. Phys. Soc. 78, 491-495.]) and Steigmann et al. (1965[Steigmann, G. A., Sutherland, H. H. & Goodyear, J. (1965). Acta Cryst. 19, 967-971.]) who reported the lattice parameters as a0 = 7.623 Å and c0 = 32.36 Å . Hahn described the space group of the cubic α-In2S3 as [Fd\overline{3}m], with a lattice parameter of 10.72 Å (Hahn & Klingler, 1949[Hahn, H. & Klingler, W. (1949). Z. Anorg. Chem. 260, 97-109.]). The high-temperature γ-In2S3 modification has not been characterized in detail yet since a quenching to room temperature conditions was not successful (Diehl et al., 1976[Diehl, R., Carpentier, C.-D. & Nitsche, R. (1976). Acta Cryst. B32, 1257-1260.]). However, Diehl et al. succeeded in synthesizing a modified γ-In2S3 phase with an additional 5 at. % of As or Sb stabilizing it at room temperature. For this modified trigonal γ-In2S3 they suggested a space group of [P\overline{3}m1] with lattice parameters a0 between 3.806 and 3.831 Å and c0 between 9.044 and 9.049 Å.

Some recent publications refer to the thin film application of In2S3 and cite X-ray diffraction (XRD) database information. An assignment to the tetragonal β-In2S3 is often made although the quality of the diffraction data for In2S3 is generally poor and does not allow for a definite distinction between the tetragonal β-In2S3 and the cubic α-In2S3. A correct distinction between the two phases would however allow for a verification of the film stoichiometry, as the β-In2S3 only exists in a very small stoichiometry range (Gödecke & Schubert, 1985[Gödecke, T. & Schubert, K. (1985). Z. Metallkdd. 76, 358-364.]). It is the scope of this work to reinvestigate the In2S3 in view of these aspects and to present a clear and detailed description of its crystal structure over the temperature range from room temperature to above 1300 K.

2. Experimental methods

2.1. Sample preparation

Indium sulfide was synthesized by heating weighted stoichiometric amounts of sulfur and indium in the ratio 3:2 in evacuated quartz ampoules. Source materials were indium granulate (> 99.999% purity) and sulfur flakes (> 99.999% purity). An excess of 4 atomic percent of S was provided for two reasons: (i) to account for sulfur losses during the preparation; (ii) the excess sulfur is not incorporated into the crystal structure according to the phase diagram (Gödecke & Schubert, 1985[Gödecke, T. & Schubert, K. (1985). Z. Metallkdd. 76, 358-364.]). The weighted source materials were filled in a graphite boat and placed in a quartz glass ampoule, evacuated (< 10−3 mbar) and sealed. The ampoule was placed in a two-zone oven and heated above the melting point to 1373 K for 24 h to enable the complete sulfurization of the indium. To assure that the synthesis was completed, the ampoule was then kept at 1073 K for 2 d, at 873 K for 2 d and at 673 K for 4 d. The synthesized indium sulfide was manually ground in a mortar before XRD measurements and had a brick red appearance.

2.2. Diffraction measurement

For the diffraction measurement, ground indium sulfide powder was sealed in a quartz glass ampoule. The XRD measurements were performed at the ESRF, the European Synchrotron, Grenoble, France, at the beamline ID15B using monochromatic high-energy synchrotron light with a wavelength of 0.14276 Å. XRD patterns were recorded every 5 K, while heating the sample from 304 to 1322 K with a constant heating rate of 300 K h−1. The detector type used was a Pixium 4700, the detector-to-sample distance was 1037.132 mm, and the furnace was a resistively heated tube furnace.

2.3. Details of the Rietveld refinement

The software FullProf Suite (February 2016 version) was used for the Rietveld refinement of the recorded diffractograms (Rodriguez-Carvajal, 2001[Rodriguez-Carvajal, J. (2001). IUCr Newsl. 26, 12-19.]). Apart from 7 profile parameters, the lattice parameters were fitted, as well as the atomic position coordinate parameters where appropriate, the isotropic temperature factor for all atomic positions, and the occupational factors of the indium positions. In addition we fitted the background by a list of manually inserted points which will add to the list of refined parameters.

3. Results

3.1. Temperature-dependent phase analysis

In2S3 powder was prepared from the elements via a high-temperature route as described in §2.1[link]. The mortared In2S3 powder was filled into quartz glass ampoules which were instantanously evacuated and sealed. The X-ray diffractograms were recorded in the temperature range from 304 to 1322 K, and details on the diffraction measurements can be found in §2.2[link]. A colorscale map of all diffractograms is depicted in Fig. 1[link]. In this graph, color indicates the counts and the three temperature ranges corresponding to the three different phases with distinct diffraction patterns can be well separated. We find a sharp structural phase transition between the first two phases at a temperature of 717 ± 5 K. This transition is characterized by the disappearance of several minor intensity diffraction peaks, while the main diffraction peak positions and intensities remain appoximately constant for both phases (see Figs. 1[link] and 2[link]). A second transition was observed in the temperature range between 1049 and 1084 K. Here, all peaks of the α-In2S3 disappear and are replaced by the diffraction peaks of γ-In2S3 indicating a complete reordering of the atoms in the crystal structure.

[Figure 1]
Figure 1
Map of temperature-dependent X-ray diffractograms (XRD) for the In2S3 powder. Each diffractogram is measured at a specific temperature which corresponds to one column in the graph with the y-direction displaying the 2θ diffraction angle. The position of the column in the x-direction corresponds to the temperature at which the diffractogram was recorded, while the colour indicates the XRD intensity in logarhythmic scale (dark blue = low intensity, red = high intensity). Three structural modifications of In2S3 in different temperature regimes can be distinguished according to appearing/disappearing diffraction peaks as indicated in the figure. The wavelength of the incident X-ray photons is 0.14276 Å.
[Figure 2]
Figure 2
Diffraction data and Rietveld refinement of the three In2S3 modifications fully refined in this study. The displayed residua have been vertically shifted for better comparison. (a) Tetragonal modification measured at 309 K; (b) cubic modification measurement at 749 K; (c) trigonal modification measured at 1099 K.

3.2. Rietveld refinement of the three In2S3 modifications

We carried out full Rietveld refinements of the three diffractograms recorded at temperatures of 309, 749 and 1099 K. The refined lattice parameters are listed in Table 1[link]. The three diffractograms, refinements and residues are displayed in Fig. 2[link]. The calculated figures of merit and atomic positions can be found in Tables 2[link] and 3–5[link][link][link], respectively. For all three phases we obtained a good agreement between measurement and simulation with χ2 values below 11 and Bragg RI-factors below 0.03. More detailed information on the refinement parameters are included in the CIF file in the supporting information.

Table 1
Lattice parameters as extracted from the Rietveld refinement for the three In2S3 modifications

    Temperature (K) Space group Number a0 = b0 (Å) c0 (Å) [{\alpha = \beta}] (°) γ (°)
#1 β-In2S3 309 I41/amd 141 7.6231 (4) 32.358 (3) 90 90
#2 α-In2S3 749 [Fd\overline{3}m] 227 10.8315 (2) 10.8315 (2) 90 90
#3 γ-In2S3 1099 [P\overline{3}m1] 164 3.8656 (2) 9.1569 (5) 90 120

Table 2
Refinement parameters for the Rietveld refinements performed for the three modifications of In2S3

RB: Bragg RI factor, [\chi _{{\nu}}^{{2}}]: [\chi _{{\nu}}^{{2}}] for points with Bragg contribution, Rwp: weighted profile R-factor (not corrected for background), Rexp: expected R-factor (not corrected for background), cRwp: weighted profile R factor (corrected for background), cRexp: expected R-factor (corrected for background), Ratioh/p: ratio between effective number of reflections and intensity parameters.

    Temperature (K) [\chi _{{\nu}}^{{2}}] RB Rwp Rexp cRwp cRexp Ratioh/p
#1 β-In2S3 309 10.4 0.014 0.031 0.01 0.049 0.015 5.9
#2 α-In2S3 749 7.9 0.017 0.028 0.01 0.055 0.020 7.3
#3 γ-In2S3 1099 10.8 0.028 0.030 0.01 0.072 0.024 3.8

Table 3
Atomic sites for the tetragonal β-In2S3, (space group 141, origin choice No. 2)

Atom Wickoff x y z Uiso Occ.
S1 16h 0 −0.005 (2) 0.2513 (7) 0.013 (4) 1.0
S2 16h 0 0.008 (2) 0.0777 (7) 0.016 (4) 1.0
S3 16h 0 0.020 (2) 0.4133 (7) 0.015 (4) 1.0
In1 8e 0 1/4 0.2046 (2) 0.0097 (8) 0.973 (6)
In2 8c 0 0 0 0.0143 (15) 0.972 (7)
In3 16h 0 −0.0196 (3) 0.3327 (2) 0.0111 (10) 0.974 (6)

Table 4
Atomic sites for the cubic α-In2S3 (space group 227, origin choice No. 2)

Atom Wickoff x y z Uiso Occ.
S1 32e 0.2564 (2) 0.2564 (2) 0.2564 (2) 0.0347 (11) 1.0
In1 8a 1/8 1/8 1/8 0.0306 (9) 0.64 (4)
In2 16d 1/2 1/2 1/2 0.0445 (6) 0.978 (6)

Table 5
Atomic sites for the trigonal γ-In2S3

Atom Wickoff x y z Uiso Occ.
S1 2d 1/3 2/3 0.3359 (7) 0.054 (4) 1.0
S2 1a 0 0 0 0.091 (5) 1.0
In1 2d 1/3 2/3 0.8085 (3) 0.0510 (9) 0.829 (10)
In2 2d 1/3 2/3 0.6485 (12) 0.064 (6) 0.144 (3)

3.3. Temperature dependence of the lattice parameters

Finally, the XRD data measured at different temperatures have been processed in batch-mode Rietveld refinements to extract the temperature dependence of the relevant lattice parameters. An example is shown in Fig. 3[link] for the lattice parameter a0 of the cubic α-In2S3 phase in the temperature range between 749 and 1044 K. The data are well fitted with the linear fit function

[a(T) = 10.7480\,{\rm \AA} + 1.121\, (2)T \times 10^{-4}\,{\rm \AA}\, {\rm K}^{-1}.]

From the temperature dependence of the lattice constants, we obtain the linear thermal expansion coeffient α ([\alpha _{\rm T} = (da/dT)_{\rm T}/a_{\rm T}]; Kundra & Ali, 1976[Kundra, K. & Ali, S. (1976). Phys. Status Solidi. (A), 36, 517-525.]) for the cubic phase. In this temperature range, we calculate an average α of 10.3 × 10−6 K, in relatively good agreement with Kundra & Ali (1976[Kundra, K. & Ali, S. (1976). Phys. Status Solidi. (A), 36, 517-525.]), 10.8 × 10−6–10.9 × 10−6 K). The temperature dependence of the remaining lattice parameters of the tetragonal and trigonal phases are obtained accordingly and the fit functions are summarized in Table 6[link]. For the tetragonal β-In2S3 we find an average linear expansion coefficient α in the a direction of 11.7 × 10−6 K and in the c direction of 6.7 × 10−6 K. For the trigonal phase, an average linear expansion coefficient of 14.1 × 10−6 K in the a direction and 26.7 × 10−6 K in the c direction was determined.

Table 6
Linear fit function of the lattice parameters of In2S3 from the Rietveld refinements

    Temperature range (K) Fit function
#1 β-In2S3 309–704 a = 7.5949 (2) Å + 8.967 (31)T × 10−5 Å K−1
      c = 32.307 (2) Å + 1.607 (4)T × 10−4 Å K−1
#2 α-In2S3 749–1044 a = 10.7480 (2) Å + 1.121 (2)T × 10−4 Å K−1
#3 γ-In2S3 1099–1322 a = 3.8044 (2) Å + 5.566 (15)T × 10−5 Å K−1
      c = 8.877 (3) Å + 2.52 (2)T × 10−4 Å K−1
[Figure 3]
Figure 3
Temperature dependence of the lattice parameter for the cubic α-In2S3. The resulting fit parameters of a linear fit are listed in the inset.

4. Discussion

In this section we will briefly discuss the crystal structure of the three modifications and how the different modifications may impact the application of In2S3 in thin film solar cells.

The low-temperature modification β-In2S3 is best described with a defect spinel-type structure. The S atoms form a distorted cubic closed-spaced sublattice, in which the In atoms occupy the tetrahedral and octahedral interstitials the same way cations do in a regular spinel-like MgAl2O4 (Kleber et al., 2002[Kleber, W., Bautsch, H.-J., Bohm, J. & Klimm, D. (2002). Einführung in die Kristallographie. Oldenbourg Wissenschaftsverlag.]). While all the octahedral cation sites are occupied in β-In2S3, one third of the tetrahedal sites remain unoccupied. For that reason the In2S3 structure is sometimes described in a quasi-ternary compound formula: [In2/3(Vac)1/3]tet[In]oct2S4, where []tet and []oct denote tetrahedral and octahedral sites and (Vac) the vacancies. In β-In2S3, the vacancies are ordered along the 41 screw axis which is by definition parallel to the c-axis. The ordering of the vacancies gives rise to a small distortion of the cubic symmetry of the regular spinel structure. This small distortion is the origin of the tetragonal structure of the β-In2S3 with lower symmetry and leads to the additional peaks observed in the X-ray diffraction. Fig. 4[link] shows a plot of the β-In2S3 crystal structure based on the results obtained by the Rietveld refinement.

[Figure 4]
Figure 4
Structure model of a β-In2S3 unit cell. The tetrahedral bonds are drawn thicker for better identification. The indium vacancies are marked as grey spheres. In the tetragonal β-In2S3 configuration, the vacancies are ordered on a 41 screw axis parallel to the c-axis of the crystal. In the α-In2S3 configuration, the vacancies are randomly distributed over all tetrahedral indium sites. The edges of the unit cell of the tetragonal β-In2S3 structure (cubic α-In2S3) structure are indicated by black (blue) lines.

The transition from β-In2S3 to α-In2S3 is an order–disorder transition. In α-In2S3, the indium vacancies are randomly distributed over all tetrahedral sites, in contrast to the ordered configuration of vacancies in the β-In2S3. As a result of the disordering, α-In2S3 adopts a cubic crystal structure. The resulting higher crystal symmetry explains the observed disappearance of some of the minor intensity peaks in the diffractograms at the transition from β-In2S3 to α-In2S3 at 717 K.

Finally, the γ-In2S3 can be described as a layered structure as suggested by Diehl et al. and Bartzokas et al. (Diehl et al., 1976[Diehl, R., Carpentier, C.-D. & Nitsche, R. (1976). Acta Cryst. B32, 1257-1260.]; Bartzokas et al., 1978[Bartzokas, D., Manolikas, C. & Spyridelis, J. (1978). Phys. Status Solidi. (A), 47, 459-467.]). Here, the S atoms remain in a nearly closed-packed sublattice while the In atoms are exclusively found on octahedral sites forming a layered structure of subsequent S—In—S—In—S slabs.

The defect spinel-type structure of the β-In2S3 and α-In2S3 has a direct phenomenological and technological impact. Because of the large number of natural vacancies in the structure, In2S3 can host various other atoms such as Na or Cu within its original lattice configuration (Barreau et al., 2006[Barreau, N., Deudon, C., Lafond, A., Gall, S. & Kessler, J. (2006). Solar Energy Mater. Solar Cells, 90, 1840-1848.]). Both have been found to diffuse efficiently through In2S3 thin films (Pistor, Allsop et al., 2009[Pistor, P., Allsop, N., Braun, W., Caballero, R., Camus, C., Fischer, C.-H., Gorgoi, M., Grimm, A., Johnson, B., Kropp, T., Lauermann, I., Lehmann, S., Mönig, H., Schorr, S., Weber, A. & Klenk, R. (2009). Phys. Status Solidi A, 206, 1059-1062.]; Juma, Pistor et al., 2012[Juma, A., Pistor, P., Fengler, S., Dittrich, T. & Wendler, E. (2012). Thin Solid Films. In the press.]; Juma, Kavalakkatt et al., 2012[Juma, A., Kavalakkatt, J., Pistor, P., Latzel, B., Schwarzburg, K. & Dittrich, T. (2012). Phys. Status Solidi. (A), 209, 663-668.]). The diffusion phenomena at interfaces in thin film solar cells containing In2S3 have not yet been fully understood but might benefit from an in-depth knowledge of the crystal (vacancy) structure of In2S3. Where XRD data of good quality exist, it is an easy task to distinguish between the tetragonal and cubic phase of In2S3 by an examination of the characteristic additional peaks only present for the β-In2S3. The differentiation between α-In2S3 and β-In2S3 allows testing for stoichiometry, since the tetragonal phase only exists in a very small compositional range. According to Gödecke & Schubert (1985[Gödecke, T. & Schubert, K. (1985). Z. Metallkdd. 76, 358-364.]) and Diehl & Nitsche (1975[Diehl, R. & Nitsche, R. (1975). J. Cryst. Growth, 28, 306-310.]), the compositional range for the β-In2S3 is less than 1 at %. The addition of a very small amount of surplus indium effectively surpresses the ordering of the In vacancies and therefore the formation of the tetragonal β-In2S3 phase. As a consequence, the crystal structure of off-stoichiometric In2+xS3 remains in the cubic α-In2S3 modification down to room temperature. This specific feature is used for example to check if In2S3 source material for an evaporation process is still within the described stoichiometry range (Pistor, Caballero et al., 2009[Pistor, P., Caballero, R., Hariskos, D., Izquierdo-Roca, V., Wächter, R., Schorr, S. & Klenk, R. (2009). Solar Energy Mater. Solar Cells, 93, 148-152.]). Alike the distinction of polycrystalline powder material, this type of analysis tool would be rather desirable for the evalution of In2S3 thin film material as well. However, reasonable X-ray diffraction data are necessary to distinguish between the two very similar spectra in order to resolve the fine additional peaks, a criterion often not met for XRD patterns available on In2S3 thin films.

5. Conclusions

We provide a detailed crystal structure analysis of In2S3 over the entire temperature range from room temperature up to close to the melting point covering the three modifications β-In2S3, α-In2S3 and γ-In2S3. With this, we contribute to the comprehensive understanding of the different phases existent and their interdependence. The high-temperature phase γ-In2S3 has been analysed and refined for the first time in the pure phase. Finally we show how the detailed knowledge of the phase diagram and the different In2S3 modifications might have a direct impact on the technological use of In2S3 in applications such as buffer layer deposition in thin film solar cell production.

Supporting information


Computing details top

For all compounds, program(s) used to refine structure: FULLPROF.

(TetragonalIn2S3) top
Crystal data top
In2S3V = 1880.4 (2) Å3
Mr = 5115.31Z = 1
Tetragonal, I41/amdMelting point: 1323 K
Hall symbol: -I 4bd 2Synchrotron radiation
a = 7.6231 (4) ÅT = 309 K
c = 32.358 (3) ÅSpecimen preparation: Prepared at 1373 K
Data collection top
Beamline ID 15B at ESRF
diffractometer
2θmin = 0.060°, 2θmax = 13.598°, 2θstep = 0.008°
Radiation source: synchrotron radiation
Refinement top
Rp = 0.022Excluded region(s): 2
Rwp = 0.03150 parameters
Rexp = 0.0101 restraint
RBragg = 0.014(Δ/σ)max = 0.002
1624 data points
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
S10.000000.005 (2)0.2513 (7)0.013 (4)
S20.000000.008 (2)0.0777 (7)0.016 (4)
S30.000000.020 (2)0.4133 (7)0.015 (3)
In10.000000.250000.20459 (11)0.0097 (8)0.973 (6)
In20.000000.000000.000000.0143 (15)0.972 (7)
In30.000000.0196 (3)0.33273 (12)0.0111 (10)0.974 (6)
Geometric parameters (Å, º) top
S1—In12.462 (19)In1—S3iv2.454 (12)
S1—In2i2.669 (11)In2—S1iii2.669 (11)
S1—In2ii2.669 (11)In2—S1iv2.669 (11)
S1—In32.64 (2)In2—S1viii2.669 (11)
S2—In22.51 (2)In2—S1ix2.669 (11)
S2—In3iii2.552 (11)In2—S22.51 (2)
S2—In3iv2.552 (11)In2—S2x2.51 (2)
S3—In1i2.454 (18)In3—S12.64 (2)
S3—In32.62 (2)In3—S2i2.552 (11)
S3—In3v2.705 (10)In3—S2ii2.552 (11)
S3—In3vi2.705 (10)In3—S32.62 (2)
In1—S12.462 (19)In3—S3v2.705 (12)
In1—S1vii2.462 (19)In3—S3vi2.705 (12)
In1—S3iii2.454 (12)
In1—S1—In2i122.9 (5)S1iv—In2—S1viii91.2 (7)
In1—S1—In2ii122.9 (5)S1iv—In2—S1ix180.0 (7)
In1—S1—In3130.3 (7)S1iv—In2—S288.1 (9)
In2i—S1—In2ii91.1 (3)S1iv—In2—S2x91.9 (10)
In2i—S1—In389.2 (6)S1viii—In2—S1ix88.8 (3)
In2ii—S1—In389.2 (6)S1viii—In2—S291.9 (10)
In2—S2—In3iii94.7 (6)S1viii—In2—S2x88.1 (9)
In2—S2—In3iv94.7 (6)S1ix—In2—S291.9 (10)
In3iii—S2—In3iv87.0 (4)S1ix—In2—S2x88.1 (9)
In1i—S3—In3116.4 (7)S2—In2—S2x180.0 (16)
In1i—S3—In3v121.2 (5)S1—In3—S2i88.0 (9)
In1i—S3—In3vi121.2 (5)S1—In3—S2ii88.0 (9)
In3—S3—In3v97.0 (6)S1—In3—S3171.0 (16)
In3—S3—In3vi97.0 (6)S1—In3—S3v90.9 (10)
In3v—S3—In3vi98.9 (3)S1—In3—S3vi90.9 (10)
S1—In1—S1vii104.3 (10)S2i—In3—S2ii92.6 (3)
S1—In1—S3iii109.5 (11)S2i—In3—S398.2 (10)
S1—In1—S3iv109.5 (11)S2i—In3—S3v174.0 (8)
S1vii—In1—S3iii109.5 (11)S2i—In3—S3vi93.3 (8)
S1vii—In1—S3iv109.5 (11)S2ii—In3—S398.2 (10)
S3iii—In1—S3iv114.0 (4)S2ii—In3—S3v93.3 (8)
S1iii—In2—S1iv88.8 (3)S2ii—In3—S3vi174.0 (8)
S1iii—In2—S1viii180.0 (7)S3—In3—S3v82.3 (9)
S1iii—In2—S1ix91.2 (7)S3—In3—S3vi82.3 (9)
S1iii—In2—S288.1 (9)S3v—In3—S3vi80.8 (4)
S1iii—In2—S2x91.9 (10)
Symmetry codes: (i) y+1/4, x1/4, z+1/4; (ii) y1/4, x1/4, z+1/4; (iii) y+1/4, x+1/4, z1/4; (iv) y1/4, x+1/4, z1/4; (v) y1/4, x+1/4, z+3/4; (vi) y+1/4, x+1/4, z+3/4; (vii) x, y+1/2, z; (viii) y1/4, x1/4, z+1/4; (ix) y+1/4, x1/4, z+1/4; (x) x, y, z.
(CubicIn2S3) top
Crystal data top
In2S3V = 1270.76 (4) Å3
Mr = 3408.76Z = 1
Cubic, Fd3mMelting point: 1323 K
Hall symbol: -F 4vw 2vw 3Synchrotron radiation
a = 10.8315 (2) ÅT = 749 K
Data collection top
Beamline ID 15B at ESRF
diffractometer
2θmin = 0.060°, 2θmax = 13.606°, 2θstep = 0.008°
Radiation source: synchrotron radiation
Refinement top
Rp = 0.019Excluded region(s): 2
Rwp = 0.02868 parameters
Rexp = 0.0100 restraints
RBragg = 0.017(Δ/σ)max = 0.002
1625 data points
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
s10.25638 (14)0.25638 (14)0.25638 (14)0.0347 (12)
in10.125000.125000.125000.0306 (9)0.640 (4)
in20.500000.500000.500000.0444 (6)0.978 (6)
Geometric parameters (Å, º) top
s1—in12.4648 (15)in1—s1vi2.4648 (15)
s1—in2i2.6406 (15)in2—s1vii2.6406 (15)
s1—in2ii2.6406 (15)in2—s1viii2.6406 (15)
s1—in2iii2.6406 (15)in2—s1ix2.6406 (15)
in1—s12.4648 (15)in2—s1x2.6406 (15)
in1—s1iv2.4648 (15)in2—s1xi2.6406 (15)
in1—s1v2.4648 (15)in2—s1xii2.6406 (15)
in1—s1—in2i123.14 (5)s1vii—in2—s1x93.04 (8)
in1—s1—in2ii123.14 (5)s1vii—in2—s1xi86.96 (8)
in1—s1—in2iii123.14 (5)s1vii—in2—s1xii93.04 (8)
in2i—s1—in2ii92.96 (5)s1viii—in2—s1ix93.04 (8)
in2i—s1—in2iii92.96 (5)s1viii—in2—s1x86.96 (8)
in2ii—s1—in2iii92.96 (5)s1viii—in2—s1xi93.04 (8)
s1—in1—s1iv109.47 (10)s1viii—in2—s1xii86.96 (8)
s1—in1—s1v109.47 (10)s1ix—in2—s1x180.00 (10)
s1—in1—s1vi109.47 (10)s1ix—in2—s1xi86.96 (8)
s1iv—in1—s1v109.47 (10)s1ix—in2—s1xii93.04 (8)
s1iv—in1—s1vi109.47 (10)s1x—in2—s1xi93.04 (8)
s1v—in1—s1vi109.47 (10)s1x—in2—s1xii86.96 (8)
s1vii—in2—s1viii180.00 (10)s1xi—in2—s1xii180.00 (10)
s1vii—in2—s1ix86.96 (8)
Symmetry codes: (i) x+1, y1/4, z1/4; (ii) x1/4, y+1, z1/4; (iii) x1/4, y1/4, z+1; (iv) x, y+1/4, z+1/4; (v) x+1/4, y, z+1/4; (vi) x+1/4, y+1/4, z; (vii) x+3/4, y+3/4, z; (viii) x+1/4, y+1/4, z+1; (ix) x+3/4, y, z+3/4; (x) x+1/4, y+1, z+1/4; (xi) x, y+3/4, z+3/4; (xii) x+1, y+1/4, z+1/4.
(TrigonalIn2S3) top
Crystal data top
In2S3V = 118.50 (1) Å3
Mr = 319.85Z = 1
Trigonal, P3m1Melting point: 1323 K
Hall symbol: -P 3 2"Synchrotron radiation
a = 3.86564 (14) ÅT = 1099 K
c = 9.1569 (5) Å
Data collection top
Beamline ID 15B at ESRF
diffractometer
2θmin = 0.058°, 2θmax = 13.598°, 2θstep = 0.008°
Radiation source: synchrotron radiation
Refinement top
Rp = 0.020Excluded region(s): 2
Rwp = 0.03060 parameters
Rexp = 0.0100 restraints
RBragg = 0.028(Δ/σ)max = 0.002
1624 data points
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
S10.333330.666660.3359 (7)0.054 (3)
S20.000000.000000.000000.091 (5)
In10.333330.666660.8085 (3)0.0510 (9)0.829 (10)
In20.333330.666660.6485 (12)0.064 (6)0.144 (3)
Geometric parameters (Å, º) top
S1—In1i2.594 (4)In1—S1i2.594 (4)
S1—In1ii2.594 (4)In1—S1ii2.594 (4)
S1—In1iii2.594 (4)In1—S1iii2.594 (4)
S1—In22.862 (13)In1—S2viii2.8383 (17)
S1—In2i2.2364 (8)In1—S2ix2.8383 (17)
S1—In2ii2.2364 (8)In1—S2x2.8383 (17)
S1—In2iii2.2364 (8)In1—In21.465 (11)
S2—In1iv2.8383 (17)In2—S12.862 (13)
S2—In1v2.8383 (17)In2—S1i2.2364 (8)
S2—In1vi2.8383 (17)In2—S1ii2.2364 (8)
S2—In1vii2.8383 (17)In2—S1iii2.2364 (8)
S2—In1i2.8383 (17)In2—In11.465 (11)
S2—In1ii2.8383 (17)
In1i—S1—In1ii96.33 (11)In1i—S2—In1ii85.84 (5)
In1i—S1—In1iii96.33 (11)S1i—In1—S1ii96.33 (11)
In1i—S1—In2120.6 (5)S1i—In1—S1iii96.33 (11)
In1i—S1—In2ii117.51 (18)S1i—In1—S2viii88.66 (18)
In1i—S1—In2iii117.51 (18)S1i—In1—S2ix88.66 (18)
In1ii—S1—In1iii96.33 (11)S1i—In1—S2x172.49 (18)
In1ii—S1—In2120.6 (5)S1i—In1—In259.4 (5)
In1ii—S1—In2i117.51 (18)S1ii—In1—S1iii96.33 (11)
In1ii—S1—In2iii117.51 (18)S1ii—In1—S2viii88.66 (18)
In1iii—S1—In2120.6 (5)S1ii—In1—S2ix172.49 (18)
In1iii—S1—In2i117.51 (18)S1ii—In1—S2x88.66 (18)
In1iii—S1—In2ii117.51 (18)S1ii—In1—In259.4 (5)
In2—S1—In2i86.3 (5)S1iii—In1—S2viii172.49 (18)
In2—S1—In2ii86.3 (5)S1iii—In1—S2ix88.66 (18)
In2—S1—In2iii86.3 (5)S1iii—In1—S2x88.66 (18)
In2i—S1—In2ii119.60 (3)S1iii—In1—In259.4 (5)
In2i—S1—In2iii119.59 (3)S2viii—In1—S2ix85.84 (5)
In2ii—S1—In2iii119.59 (3)S2viii—In1—S2x85.84 (5)
In1iv—S2—In1v85.84 (5)S2viii—In1—In2128.2 (7)
In1iv—S2—In1vi85.84 (5)S2ix—In1—S2x85.84 (5)
In1iv—S2—In1vii94.16 (11)S2ix—In1—In2128.2 (7)
In1iv—S2—In1i94.16 (11)S2x—In1—In2128.2 (7)
In1iv—S2—In1ii179.97 (11)S1—In2—S1i93.7 (4)
In1v—S2—In1vi85.84 (5)S1—In2—S1ii93.7 (4)
In1v—S2—In1vii94.16 (11)S1—In2—S1iii93.7 (4)
In1v—S2—In1i179.97 (11)S1—In2—In1180.0 (7)
In1v—S2—In1ii94.16 (11)S1i—In2—S1ii119.60 (3)
In1vi—S2—In1vii179.97 (11)S1i—In2—S1iii119.59 (3)
In1vi—S2—In1i94.16 (11)S1i—In2—In186.3 (5)
In1vi—S2—In1ii94.16 (11)S1ii—In2—S1iii119.59 (3)
In1vii—S2—In1i85.84 (5)S1ii—In2—In186.3 (5)
In1vii—S2—In1ii85.84 (5)S1iii—In2—In186.3 (5)
Symmetry codes: (i) x, y+1, z+1; (ii) x+1, y+1, z+1; (iii) x+1, y+2, z+1; (iv) x1, y1, z1; (v) x, y1, z1; (vi) x, y, z1; (vii) x, y, z+1; (viii) x, y, z+1; (ix) x, y+1, z+1; (x) x+1, y+1, z+1.
 

Footnotes

Current address: IREC – Catalonia Institute for Energy Research, Sant Adriá de Besós, Spain.

Acknowledgements

Partial financial support for this work by the European Commission under contract number FP-6-019757 (LARCIS) and grant agreement number GA 625840 (JumpKEST) as well as by the DAAD within the PPP-program Acciones Integradas Hispano-Alemanas under the Contract No. 314-Al-e-dr (HA2006-0025 spanish reference) is gratefully acknowledged. The authors are also grateful to ESRF – The European Synchrotron, Grenoble, France, for granted beamtime. We are grateful to Juan Rodriguez-Carvajal for developing the FullProf program and his advice on the refinement.

References

First citationBarreau, N. (2009). Solar Energy, 83, 363–371.  CrossRef CAS Google Scholar
First citationBarreau, N., Deudon, C., Lafond, A., Gall, S. & Kessler, J. (2006). Solar Energy Mater. Solar Cells, 90, 1840–1848.  CrossRef CAS Google Scholar
First citationBartzokas, D., Manolikas, C. & Spyridelis, J. (1978). Phys. Status Solidi. (A), 47, 459–467.  CrossRef CAS Google Scholar
First citationDiehl, R., Carpentier, C.-D. & Nitsche, R. (1976). Acta Cryst. B32, 1257–1260.  CrossRef CAS IUCr Journals Google Scholar
First citationDiehl, R. & Nitsche, R. (1975). J. Cryst. Growth, 28, 306–310.  CrossRef CAS Google Scholar
First citationGödecke, T. & Schubert, K. (1985). Z. Metallkdd. 76, 358–364.  Google Scholar
First citationGoodyear, J. & Steigmann, G. (1961). Proc. Phys. Soc. 78, 491–495.  CrossRef CAS Google Scholar
First citationHahn, H. & Klingler, W. (1949). Z. Anorg. Chem. 260, 97–109.  CrossRef CAS Google Scholar
First citationHariskos, D., Menner, R., Lotter, E., Spiering, S. & Powalla, M. (2005). 20th European Photovoltaic Solar Energy Conference, pp. 1713–1716. Barcelona, Spain.  Google Scholar
First citationJuma, A., Kavalakkatt, J., Pistor, P., Latzel, B., Schwarzburg, K. & Dittrich, T. (2012). Phys. Status Solidi. (A), 209, 663–668.  CrossRef CAS Google Scholar
First citationJuma, A., Pistor, P., Fengler, S., Dittrich, T. & Wendler, E. (2012). Thin Solid Films. In the press.  Google Scholar
First citationKing, G. S. D. (1962). Acta Cryst. 15, 512.  CrossRef IUCr Journals Google Scholar
First citationKleber, W., Bautsch, H.-J., Bohm, J. & Klimm, D. (2002). Einführung in die Kristallographie. Oldenbourg Wissenschaftsverlag.  Google Scholar
First citationKundra, K. & Ali, S. (1976). Phys. Status Solidi. (A), 36, 517–525.  CrossRef CAS Google Scholar
First citationLarina, L., Kim, K. H., Yoon, K. H., Konagai, M. & Ahn, B. T. (2004). J. Electrochem. Soc. 151, C789–C792.  CrossRef CAS Google Scholar
First citationNaghavi, N., Spiering, S., Powalla, M., Cavana, B. & Lincot, D. (2003). Prog. Photovolt. Res. Appl. 11, 437–443.  CrossRef CAS Google Scholar
First citationPistor, P., Allsop, N., Braun, W., Caballero, R., Camus, C., Fischer, C.-H., Gorgoi, M., Grimm, A., Johnson, B., Kropp, T., Lauermann, I., Lehmann, S., Mönig, H., Schorr, S., Weber, A. & Klenk, R. (2009). Phys. Status Solidi A, 206, 1059–1062.  CrossRef CAS Google Scholar
First citationPistor, P., Caballero, R., Hariskos, D., Izquierdo-Roca, V., Wächter, R., Schorr, S. & Klenk, R. (2009). Solar Energy Mater. Solar Cells, 93, 148–152.  CrossRef CAS Google Scholar
First citationRao, P. & Kumar, S. (2012). Thin Solid Films, 524, 93–99.  CrossRef CAS Google Scholar
First citationRodriguez-Carvajal, J. (2001). IUCr Newsl. 26, 12–19.  Google Scholar
First citationSáez-Araoz, R., Krammer, J., Harndt, S., Koehler, T., Krueger, M., Pistor, P., Jasenek, A., Hergert, F., Lux-Steiner, M. & Fischer, C. (2012). Prog. Photovolt. Res. Appl. 20, 855–861.  Google Scholar
First citationShazly, A. A. E., Elhady, D., Metwally, H. & Seyam, M. (1998). J. Phys. Condens. Matter, 10, 5943–5954.  CrossRef Google Scholar
First citationSteigmann, G. A., Sutherland, H. H. & Goodyear, J. (1965). Acta Cryst. 19, 967–971.  CrossRef IUCr Journals Google Scholar
First citationStrohm, A., Eisenmann, L., Gebhardt, R., Harding, A., Schlötzer, T., Abou-Ras, D. & Schock, H. (2005). Thin Solid Films, 480–481, 162–167.  CrossRef CAS Google Scholar
First citationJohn, T. T., Mathew, M., Sudhakartha, C., Vijayakumar, K. P., Abe, T. & Kashiwaba, Y. (2005). Solar Energy Mater. Solar Cells, 89, 27–36.  CAS Google Scholar
First citationYoosuf, R. & Jayaraj, M. (2005). Solar Energy Mater. Solar Cells, 89, 85–94.  CrossRef CAS Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206
Volume 72| Part 3| June 2016| Pages 410-415
Follow Acta Cryst. B
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds