Next Article in Journal
Enhanced Strength–Ductility Synergy of Mg-Al-Sn-Ca Alloy via Composite Asymmetric Extrusion
Previous Article in Journal
Preparation of Reduced Graphene Oxide Films with High and Uniform Thickness for Electromagnetic Interference Shielding
Previous Article in Special Issue
Combustion Synthesis of Zirconium-Doped Ceria Nanocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optical and Optoelectrical Properties of Ternary Chalcogenide CuInS2/TiO2 Nanocomposite Prepared by Mechanochemical Synthesis

1
Institute of Geotechnics, Slovak Academy of Sciences, 04001 Košice, Slovakia
2
Institute of Electronics and Photonics, Slovak University of Technology, 84104 Bratislava, Slovakia
3
Jožef Stefan Institute, 1000 Ljubljana, Slovenia
4
Central European Institute of Technology, Brno University of Technology, 61200 Brno, Czech Republic
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(4), 324; https://doi.org/10.3390/cryst14040324
Submission received: 12 March 2024 / Revised: 25 March 2024 / Accepted: 26 March 2024 / Published: 30 March 2024
(This article belongs to the Special Issue Metal Oxides: Crystal Structure, Synthesis and Characterization)

Abstract

:
In this work, a nanocomposite consisting of ternary chalcogenide CuInS2 and TiO2 was prepared and its optical and optoelectrical properties were investigated. The CuInS2/TiO2 nanocomposite was produced via one-step mechanochemical synthesis and characterized from the crystal structure, microstructural, morphology, surface, optical, and optoelectrical properties viewpoints. X-ray diffraction confirmed the presence of both components, CuInS2 and TiO2, in the nanocomposite and revealed a partial transformation of anatase to rutile. The presence of both components in the samples was also proven by Raman spectroscopy. HRTEM confirmed the nanocrystalline character of the samples as crystallites ranging from around 10 nm and up to a few tens of nanometers were found. The presence of the agglomerated nanoparticles into larger grains was proven by SEM. The measured optical properties of CuInS2, TiO2, and CuInS2/TiO2 nanocomposites demonstrate optical bandgaps of ~1.62 eV for CuInS2 and 3.26 eV for TiO2. The measurement of the optoelectrical properties showed that the presence of TiO2 in the CuInS2/TiO2 nanocomposite increased its conductivity and modified the photosensitivity depending on the ratio of the components. This study has demonstrated the possibility of preparing a CuInS2/TiO2 nanocomposite material with promising applications in optoelectronics in the visible region in an eco-friendly manner.

1. Introduction

Ternary chalcogenide semiconductors of the I–III–VI group with promising applications in electronics, optics, and catalysis have been intensively studied in recent years [1]. However, the majority of the best-investigated sulfide-based semiconductors contain toxic heavy metals, which seriously limits their potential application.
The ternary chalcogenide CuInS2 as a semiconductor with direct and indirect bandgap values of 1.60 eV and 1.51 eV [2] enables the absorption of radiation in the region of the solar radiation spectrum and thus is a promising semiconductor for the development of solar cells with low toxicity and good chemical stability. Its combination with other metallic inorganic semiconductors (e.g., TiO2 and ZnO) with wider bandgaps can lead to better optical properties due to the elimination of surface non-radiative recombination defects. Nanostructured TiO2 semiconductors are extensively used in solar cells and photocatalysis. CuInS2/TiO2 nanocomposite represents a non-toxic alternative to other previously studied nanocomposites, such as CdSe/TiO2, CdS/TiO2, and PbS/TiO2, that contain toxic heavy metals [3,4,5].
Many synthetic techniques have been reported for the synthesis of TiO2-coated CuInS2 nanomaterials, including solvothermal synthesis, the ultrasonication-assisted cathodic electrodeposition strategy, the ionic layer adsorption and reaction (SILAR) method, robotic spray pyrolysis, hydrogen plasma treatment, etc. [6,7,8,9,10,11,12,13].
CuInS2/TiO2 composites with different depositing amounts were prepared by the deposition of CuInS2 on TiO2 with one-step or two-step solvothermal reactions [6]. Enesca et al. [7] prepared a CuInS2/TiO2/SnO2 heterostructure for air decontamination by deposition of layers by spray pyrolysis followed by annealing. A new type of photoelectrode has been successfully realized using an n-type CuInS2 heterostructure modified with a TiO2 array of nanotubes (NTs) by cathodic electrodeposition using ultrasound [8]. Shen et al. [9] investigated the properties of p-type CuInS2 quantum dots synthesized via the solvothermal route and bound to n-type TiO2 nanoparticles by a heat treatment process. Also, CuInS2–TiO2 heterojunction solar cells were prepared by Nanu and co-workers using the atomic layer deposition method [10]. CuInS2/TiO2 electrodes formed from nanotube heterojunction arrays were prepared by synthesis using a modified sequential ion layer adsorption and reaction (SILAR) method [11]. Chen et al. conducted an experiment on the preparation of a TiO2 NR/CuInS2 solar cell by formation in hydrogen plasma [12]. The spin coating-assisted SILAR method was used for synthesis of CuInS2 QD-sensitized TiO2 films, which were further used as promising photoelectrodes for solid-state QDSSCs [13].
The high-energy ball milling process affords all the benefits for the expanding field of mechanochemistry [14,15]. This method represents a simple, low-temperature, and environmentally friendly alternative to the traditional preparation methods. Its eco-friendly process means that synthesis takes place in the absence of solvents and toxic organometallic precursors and unwanted side-products. We have already prepared other metal sulfide-based semiconductor–TiO2 nanocomposites via mechanochemical synthesis in the past [16,17].
According to our best knowledge, the CuInS2/TiO2 nanocomposite has not been prepared via mechanochemical synthesis so far. In the present work, the CuInS2/TiO2 nanocomposite in two different CuInS2:TiO2 molar ratios was prepared via mechanochemical synthesis and its optical and optoelectrical properties were investigated. The novelty of this study is the simple and environmentally friendly mechanochemical synthesis of CuInS2/TiO2 nanocomposite in a very short time, at ambient pressure and temperature, as suitable material for solar cell applications.

2. Materials and Methods

2.1. Materials

The CuInS2/TiO2 nanocomposite was prepared from CuInS2, which was synthesized from copper (99.7%, Aldrich, Darmstadt, Germany), indium (99.99%, Aldrich, Darmstadt, Germany), sulfur (99%, Ites, Vranov n/T, Slovakia), and commercially available TiO2 Degussa P25 (Degussa, The Netherlands) (75% anatase and 25% rutile).

2.2. Mechanochemical Synthesis of CuInS2/TiO2 Nanocomposite

The mechanochemical synthesis of CuInS2 was performed in a planetary mill Pulverisette 6 (Fritsch, Idar-Oberstein, Germany) by milling of 1.31 g of elemental copper, 2.37 g of elemental indium, and 1.32 g of elemental sulfur. A tungsten carbide milling chamber (volume 250 mL) filled with 50 balls (diameter 10 mm) of the same material was used for milling, and the milling was carried out at 550 rpm for 60 min in an argon atmosphere according to Equation (1) and the procedure provided in [18]. Mechanochemically prepared CuInS2 according to Equation (1) and commercially available TiO2 were used to synthesize the CuInS2/TiO2 nanocomposite (Equation (2)) (in molar ratios of 4:1 and 1:4 chosen based on our results published in papers [16,17] for similar nanocomposites). The synthetic process of the CuInS2/TiO2 nanocomposite is displayed in Figure 1 and can be described by the following reactions:
Cu + In + 2S→CuInS2
CuInS2 + TiO2→CuInS2/TiO2
The mechanochemical synthesis of CuInS2/TiO2 nanocomposite was also carried out in a planetary ball mill Pulverisette 6 (Fritsch, Idar-Oberstein, Germany) in argon atmosphere (>99.998%, Linde Gas group, Bratislava, Slovakia). Milling took 30 min without breaking cooling due to the shorter milling time. A 250 mL tungsten carbide chamber with 50 tungsten carbide balls with a diameter of 10 mm was vented with argon for 3 min to ensure an inert atmosphere. The rotational speed of the mill was set at 500 rpm. A ball to powder ratio of 72:1 was used.

2.3. Characterization Techniques

X-ray diffraction (XRD) measurements were performed using a D8 Advance diffractometer (Bruker, Bremen, Germany) equipped with a θ-θ goniometer, in CuKα radiation (40 kV, 40 mA), a secondary graphite monochromator, and a scintillation detector. The XRD pattern was scanned from 15° to 60° with a step of 0.03° and a counting time of 12 s. The XRD pattern was analyzed using Diffracplus Eva software with the ICDD—PDF2 database. Rietveld refinement was carried out using TOPAS Academic software [19,20].
TEM analyses were carried out using a 200 kV JEM 2100 microscope (JEOL, Akishima, Japan) with a LaB6 electron source and equipped with an energy-dispersive X-ray spectrometer (EDS). A small amount of the sample was ultrasonically homogenized in absolute ethanol. A drop of the homogenized suspension was deposited on a carbon-coated copper grid and dried. Prior to TEM analyses, the sample was covered with carbon to prevent charging under the electron beam.
Morphology was studied using a field emission-scanning electron microscope (FE-SEM, Mira 3, Tescan, Czech Republic) coupled with an EDS analyzer (Oxford Instruments, Oxford, UK).
The values of the specific surface area were obtained by using a NOVA 1200e Surface Area & Pore Size Analyzer (Quantachrome Instruments, Boynton Beach, FL, USA).
X-ray photoelectron spectroscopy (XPS) measurements were performed on a Kratos Axis Supra XPS instrument (Manchester, UK) with monochromatic AlKα X-rays (hυ = 1486.69 eV). Scans were performed at an emission current of 15 mA and in hybrid lens mode. Survey XPS and high-resolution (HR) spectra were obtained using scan steps of 1.0 and 0.1 eV, respectively. Prior to data analysis, the C1s spectrum was calibrated to 284.8 eV. Low-resolution broad scan and high-resolution scan were performed for Cu, In, S, Ti, and O. The low-resolution wide survey and high-resolution scans were performed for Cu, In, S, Ti, and O, respectively. The HR spectra were used to analyze the elemental composition and chemical state of the sample. Deconvolution and fitting of the elements of interest were performed using CasaXPS software (version 2.3.17, Casa Software Ltd., Teignmouth, UK) using Spine Shirley background in the HR spectra and Gaussian/Lorentzian line shape to fit the XPS peaks.
UV–Vis spectra of individual CuInS2, commercially available TiO2 and mechanochemically synthesized CuInS2/TiO2 nanocomposite were measured using a UV–Vis light source connected to an optical fiber and to a stage for measuring reflectance using two optical fibers. The reflectivity of the samples could only be measured on samples prepared in the form of a thin slice formed by pressing the nanocrystalline powder. The spectra were measured by the OceanOptics system and processed after software correction in the range 200–1000 nm.
Micro-Raman and PL spectra measurements at room temperature were performed using a confocal Raman microscope in backscattering geometry (Spectroscopy, Imaging, Warstein, Germany) and using a focused Ar laser beam tuned to excitation wavelength of 514 nm. The calibration of the measuring system before the measurement was carried out using the frequency of the Raman line of crystalline Si at 520 cm−1.
The optoelectrical properties of individual CuInS2, commercially available TiO2, and CuInS2/TiO2 nanocomposite were realized by measuring the I–V characteristics in the dark and under focused white light with an illumination intensity of ~600 mW/cm2 using a halogen lamp and Agilent 4155C semiconductor parameter analyzer. Each sample for measurement was prepared by dropping a solution of nanocrystalline powder in isopropyl alcohol onto a Au interdigital structure with an area of 3 × 3 mm and with finger/gap dimensions of 30/12 µm. The electrical connection to the socket was completed using Au wires glued with silver paste.

3. Results and Discussion

3.1. Structural and Microstructural Characterization

For comparison, the X-ray diffraction (XRD) patterns of previously mechanochemically prepared CuInS2, commercial TiO2, and mechanochemically synthesized CuInS2/TiO2 4:1 and 1:4 nanocomposites are shown in Figure 2. The phase composition of the prepared CuInS2/TiO2 nanocomposites was carried out by Rietveld refinement (Figure 3). In the XRD pattern of the CuInS2/TiO2 4:1 (Figure 3a), the four most intensive diffraction peaks can be ascribed to the tetragonal structure of roquesite CuInS2 (JCPDS 00-027-0159). In addition to this, a small peak at 25.1°, which can be indexed to anatase (TiO2 P25) with a tetragonal structure (JCPDS 01-075-2545), was found. A very small amount of tungsten carbide wear coming from milling balls and a vial was also evidenced. The approximate crystallite size of roquesite nanocrystals in this sample was calculated to be 22 ± 4 nm, which is in accordance with the 18 nm reported in [18]. In both cited papers, the crystallite size is reported for the as-prepared pure CuInS2. In the case of the CuInS2/TiO2 1:4 sample (Figure 3b), the diffraction lines characteristic of anatase modification (TiO2 P25) with a tetragonal structure (JCPDS 01-075-2545) are much stronger; however, the diffraction lines of roquesite CuInS2 (JCPDS 00-027-0159) with a tetragonal structure are present and are still intensive. In addition, a small amount of rutile (another tetragonal crystallographic modification of TiO2 with JCPDS 01-076-9000) was found. This fact points to partial phase transition of anatase into rutile during milling [16]. A greater amount of tungsten carbide from the wear of the milling balls and the vessel has also been demonstrated. The diffraction peaks of WC are much more intense; namely, the one at 31.3° (overlapping with the weaker reflection of roquesite) is significant. In general, all the diffraction lines are broad, indicating the formation of fine crystallites, as well as structural disorder that is common during the process of milling [21].
Figure 4 shows the Raman spectra of the synthesized CuInS2, commercially available TiO2, and mechanochemically synthesized CuInS2/TiO2 4:1 and 1:4 nanocomposites. The measured spectra confirmed the appearance of characteristic peaks that correspond to the formation of crystalline nanoparticles. It is clear from the figure that the dominant feature of the measured spectrum of the CuInS2/TiO2 nanocomposite includes TiO2 peaks, which indicate the formation of the compound in different crystallographic forms—polymorphism [22]. The intense peak at 143 cm−1 and lower peaks near 397 and 514 cm−1 correspond to anatase-form TiO2 with symmetries Eg, B1g, and A1g, respectively. The broad peak around 406 cm−1 includes two mixed peaks from 398 to 437 cm−1, which may be related to the formation of anatase and rutile forms of TiO2 with A1g and Eg symmetry, respectively [22]. The lower-intensity peaks at 245, 296, and 338 cm−1 correspond to CuInS2 phase with symmetries A1, E, and B2, respectively [18,23,24]. The results of Raman measurement confirm the results obtained by means of XRD as well as the presence of WC peaks at 808, 1420, and 1600 cm−1 for the CuInS2/TiO2 1:4 sample, as shown in the inset of Figure 4, corresponding to published data [25].
Microstructural characterization of CuInS2/TiO2 nanocomposite in 4:1 and 1:4 ratios was performed by transmission electron microscopy (TEM). Both the CuInS2/TiO2 nanocomposites consist of agglomerated nanoparticles, which was confirmed by a low-magnification TEM image (Figure 5a,b).
The crystallinity, crystallite size, and chemical composition of the samples were inspected by selected-area electron diffraction (SAED), high-resolution TEM (HRTEM), and energy-dispersive X-ray spectroscopy (EDXS). The results of the analyses for the 4:1 and 1:4 nanocomposites are shown in Figure 5c,d, respectively. The SAED patterns of both samples are diffraction rings, indicating that the products are composed of randomly oriented nanoparticles. Based on the measured d-values of the rings, we confirmed that the main phases in both nanocomposites are roquesite CuInS2 and anatase TiO2. In the 4:1 sample, roquesite is the predominant phase, whereas, in the 1:4 sample, the situation is the other way around. HRTEM analyses of both samples additionally disclose that the size of the crystallites in both samples is in the nanometer range, ranging from around 10 nm up to a few tens of nanometers. The presence of CuInS2 and TiO2 in different ratios in both nanocomposites was confirmed by EDXS. The analyses recorded from larger areas of the sample clearly show the prevalence of Cu, In, and S in the 4:1 sample and a significantly higher fraction of Ti in the 1:4 sample. For evaluation of CuInS2 and TiO2 distribution in the samples, EDXS mapping at the nanoscale would be required.

3.2. Surface and Morphological Characterization

Mechanically activated or mechanochemically synthesized samples are characterized by a variety of methods, and one of the most important characteristics is the measurement of the specific surface area (SA) value [21]. The SA of pure CuInS2, from which the studied nanocomposite was synthesized, was 4 m2g−1, as was reported in paper [18]. Upon the introduction of TiO2, the SA value considerably increased (values of 6.3 m2g−1 and 9.7 m2g−1 have been evidenced for the CuInS2/TiO2 4:1 and 1:4 samples, respectively). The SA value of TiO2 P25 is 28.7 m2g−1. The SA values of the mechanochemically synthesized CuInS2/TiO2 reported here are lower compared with the samples prepared by other methods [26].
The morphology of the prepared products was investigated by scanning electron microscopy. The SEM micrographs of the synthesized CuInS2/TiO2 nanocomposite are displayed in Figure 6.
The SEM images of the CuInS2/TiO2 4:1 and 1:4 nanocomposites show that the prepared sample consists of fine nanoparticles that form densely arranged irregular aggregates with irregular size and shape. In addition to larger micrograins, a large number of nanoparticles (100–300 nm in size) can also be found. Smaller nanoparticles can be found on the surface of larger particles and are also dispersed among larger particles.
For the surface elemental analysis, EDS mapping was performed, and the results are shown in Figure 7. The layered EDS images containing all the elements are shown in Figure 7A,B for the CuInS2/TiO2 4:1 and 1:4 nanocomposites, respectively, and in Figure 7b–f (for sample A) and Figure 7b–f (for sample B). The individual energy-dispersive X-ray spectroscopy (EDS) mapping for Cu, In, S, Ti, and O can be viewed. The distribution of all the elements in the nanocomposite appears to be homogeneous. The signal of titanium is stronger in nanocomposite CuInS2/TiO2 1:4.
The surface composition and oxidation states of various elements present in the CuInS2/TiO2 nanocomposites were investigated by X-ray photoelectron spectroscopy (XPS) (Figure 8). Survey XPS spectra revealed the presence of different elements, such as copper (Cu), indium (In), sulfur (S), titanium (Ti), oxygen (O), and carbon (C). In the case of the CuInS2/TiO2 4:1 nanocomposite (Figure 8A), the peaks of Cu, O, In, and S appear on the wide spectra of the sample. The HR low-intensity spectrum of Cu2p has a doublet of Cu2p3/2 and Cu2p1/2. However, the intensity was not high enough to analyze their chemical states. Moreover, the intensity of the S2p spectrum was not sufficiently high to clearly deconvolute the spectrum; however, it can be presumed that the spectrum is represented by the doublet S2p3/2 and S2p1/2 at 162.44 eV and 163.60 eV, respectively. The more intensive peaks at 170.30 and 169.16 eV may correspond to metal sulfate (SO4)2− or a S–O bond. In the case of the mechanochemically synthesized samples, the presence of some metal sulfates due to mechanochemical surface oxidation can be expected [27]. The intensity of the Ti HR XPS was not high enough to analyze their chemical states. However, deconvolution of the O1s spectrum revealed the presence of three components. The peaks at 532.25 eV and at 533.0 eV are related to the capping agents and the contamination as a result of the sample’s exposure to the atmosphere. The presence of titanium dioxide (TiO2) is confirmed by a slight signal at 530.07 eV. The core-level region of In3d has well-separated spin-orbit components at 445.99 eV (In3d5/2) and at 453.52 eV (In3d3/2), with a charge separation of 7.53 eV, which matches well with that of In3+ in CuInS2 [28]. The peaks’ shapes and positions indicate the potential presence of indium oxides.
In the case of nanocomposite CuInS2/TiO2 1:4 (Figure 8B), the presence of Cu, In, S, Ti, and O elements was discovered in the survey spectrum. The intensities of the Cu HR XPS were not high enough to analyze their chemical states. Similar to the CuInS2/TiO2 4:1 nanocomposite, the deconvolution of the core-level O1s spectrum demonstrates three components, with a much stronger peak due to a higher amount of TiO2, as was expected. The Ti2p region develops typical Ti2p3/2—Ti2p1/2 splitting, with the components appearing at 458.71 eV (Ti2p3/2) and 464.47 eV (Ti2p1/2). Similar to the O1s spectrum, the components’ positions confirm the presence of TiO2. Moreover, a charge separation ΔE of 5.76 eV between the two components also indicates metal oxide presence. Typical In3d5/2—In3d3/2 splitting is found again in the In3d core-level region. The components appear to be at 445.81 eV (In3d5/2) and 453.36 eV (In3d3/2), with a charge separation of 7.55 eV. The results are in line with those of the In3d region in the sample CuInS2–TiO2 (4:1) (A).

3.3. Optical Properties

Mechanochemically synthesized CuInS2, CuInS2/TiO2 4:1, and 1:4 nanocomposites and commercially available TiO2 were examined using UV–Vis reflectivity spectroscopy to investigate their optical properties (Figure 9a). Even with relatively low reflectance of the samples, it was possible to determine the change in reflectance and recalculate the relative absorption of individual samples. The diffuse reflectance data (shown in Figure 9a) were transformed to pseudoabsorption data using the Kubelka−Munk function (F(R)). This expresses the absorbance as a function of reflectance: F(R) = α/S = (1 − R)2/(2R), where R is the reflectance, α is the absorption coefficient, and S is the scattering coefficient [29]. The dependence of (F(R) hν)2 for a direct semiconductor and (F(R) hν)1/2 for an indirect semiconductor as a function of photon energy hν was used to determine the optical bandgap energy (Eg) by extrapolating the straight part to the zero line as shown in Figure 9b. As is clear from the measured reflectivity spectra of CuInS2, TiO2, and CuInS2/TiO2 with different ratios, the measured spectra are composed of the absorption properties of the individual components. Thus, from the measurement of the reflectance CuInS2/TiO2 with a molar ratio of 1:4, it was possible to determine both bandgaps of the individual components, which were for CuInS2 (1.62 eV) as a direct semiconductor (Figure 9b-inset) and for TiO2 (3.26 eV) as an indirect semiconductor. The determination of the Eg of the chalcogenide CuInS2 is in good agreement with the published data for thin films [2] and nanoparticles [30]. Similarly, the determination of Eg for TiO2 as an indirect semiconductor is in good agreement with the published data for P25 nanopowder [31].
The micro-PL spectra of the CuInS2, CuInS2/TiO2 4:1, and 1:4 nanocomposites and commercially available TiO2 when the samples are excited at a wavelength of 514 nm are shown in Figure 10. Broad emission PL spectra are typical in semiconductor nanocrystals based on chalcopyrite I–III–VI2. The size of the CuInS2 nanoparticles can cause a blue shift in the emission, and the measured PL spectrum thus exhibits size-dependent optical properties, indicating quantum confinement effects [32,33]. As can be observed in Figure 10a, the emission spectra, especially for the CuInS2/TiO2 1:4 nanocomposite, are significantly influenced by the mutual interaction of the individual components by the increased separation and transfer of the photogenerated electrons trapped in TiO2 compared to pure TiO2 and CuInS2 [34]. This is especially evident in the 500–600 nm region, where the intensity of the emission spectrum gradually increases and significantly changes the edge of the TiO2 emission spectrum. Its expected maximum is in the region of 430 nm [35] and is not shown due to the excitation wavelength of 514 nm.
Figure 10b shows the emission spectrum of the ternary chalcogenite CuInS2 as a function of energy. To determine the recombination processes in CuInS2, the measured PL spectrum can be qualitatively decomposed as the sum of two Gaussian functions with maxima of 1.66 and 1.37 eV. The emission spectrum in the region of 1.66 eV is close to the determined bandgap from the absorption measurement and corresponds to interband radiative recombination. The difference from the bulk bandgap of 1.53 eV [36] can be caused by size-dependent optical bandgap properties. The emission peak of 1.37 eV can be assigned to the D–A transition from the donor state of the indium atom, which occupies the copper vacancy (InCu) below the edge of the conduction band to the acceptor level of the copper vacancy (VCu). This corresponds well to the model of energy levels in undoped CuInS2 described and published by Ueng [36].

3.4. Optoelectrical Properties

As can be seen in Figure 11, the current–voltage (I–V) characteristics for the measured samples show an almost linear dependence, which confirms the formation of an ohmic contact on the prepared element.
As can be seen from the measured I–V characteristics, the CuInS2 sample shows the lowest conductivity. By increasing the amount of TiO2 in the ratios of CuInS2/TiO2 4:1 and 1:4, the conductivity of the samples increases proportionally to the amount of TiO2. Overall, TiO2 shows an order of magnitude higher conductivity compared to CuInS2. Intense illumination causes an increase in the number of generated charge carriers in all the samples, which results in increases in conductivity and photoresponsive current. When comparing the measurement results, the samples show 41%, 61%, 8%, and 28% increases in photoresponsive current under illumination at an applied voltage of 2 V compared to the current in the dark for CuInS2, the CuInS2/TiO2 4:1 and CuInS2/TiO2 1:4 nanocomposites, and TiO2, respectively.

4. Conclusions

In this study, CuInS2/TiO2 nanocomposites in 4:1 and 1:4 molar ratios were prepared by a facile, low-cost mechanochemical method. X-ray diffraction confirmed the presence of both components in the composites and revealed partial anatase-to-rutile phase transformation and contamination from milling media as a result of the mechanochemical treatment. The micro-Raman measurements confirmed the results from the X-ray diffraction measurements. The nanocrystalline character of the samples was confirmed by TEM as crystallites ranging from around 10 nm up to a few tens of nanometers in size were found. SEM showed that the nanoparticles agglomerated into larger grains. High-resolution XPS analysis confirmed the presence of all the elements with their expected oxidation states and documented the surface oxidation of the produced nanocomposites. Mechanochemically synthesized CuInS2 and CuInS2/TiO2 1:4 nanocomposite exhibits reflectivity measurements regarding a bandgap near 1.62 eV, while TiO2 demonstrates a bandgap of 3.26 eV. Measurement of the I–V characteristics showed that the presence of TiO2 in CuInS2/TiO2 nanocomposites increases their conductivity and modifies their sensitivity to light compared to the starting materials, CuInS2 and TiO2. The measured optoelectrical properties of the CuInS2/TiO2 nanocomposite confirmed its potential for photovoltaic applications.

Author Contributions

E.D.: Data curation, Conceptualization, Investigation, Visualization, Writing—original draft. M.B.: Investigation, Visualization, Writing—review and editing. J.K.: Investigation, Visualization, Writing—review and editing. N.D.: Investigation, Visualization, Writing—review and editing. A.K.: Investigation, Visualization, Writing—review and editing. J.B.: Investigation, Visualization. J.K.J.: Investigation, Visualization. S.K.: Investigation, Visualization. L.Č.: Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Slovak Research and Development Agency under the contracts No. APVV-18-0357, APVV-20-0437, and by the Slovak Grant Agency VEGA (projects 2/0112/22 and 2/0084/23). The support of COST Action CA18112 MechSustInd (www.mechsustind.eu, accessed on 4 March 2024), supported by the COST Association (European Cooperation in Science and Technology, www.cost.eu, accessed on 4 March 2024), is also acknowledged.

Data Availability Statement

The data presented in this study are available on request from the corresponding author, due to privacy.

Conflicts of Interest

The authors declare that they have no competing conflicts 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. Klenk, R.; Klaer, J.; Koble, C.; Mainz, R.; Merdes, S.; Rodriguez-Alvarez, H.; Scheer, R.; Schock, H.W. Development of CuInS2-based solar cells and modules. Sol. Energy Mater. Sol. Cells 2011, 95, 1441–1445. [Google Scholar] [CrossRef]
  2. Giri, R.K.; Chaki, S.H.; Dave, M.S.; Bharucha, S.R.; Khimani, A.J.; Kannaujiya, R.M.; Deshpande, M.P.; Solanki, M.B. First principle insights and experimental investigations of the electronic and optical properties of CuInS2 single crystals. Mater. Adv. 2023, 4, 3246–3256. [Google Scholar] [CrossRef]
  3. Sun, W.T.; Yu, Y.; Pan, H.Y.; Gao, X.F.; Chen, Q.; Peng, L.M. CdS quantum dots sensitized TiO2 nanotube-array photoelectrodes. J. Am. Chem. Soc. 2008, 130, 1124–1125. [Google Scholar] [CrossRef]
  4. Kongkanand, A.; Tvrdy, K.; Takechi, K.; Kuno, M.; Kamat, P.V. Quantum dot solar cells. Tuning photoresponse through size and shape control of CdSe-TiO2 architecture. J. Am. Chem. Soc. 2008, 130, 4007–4015. [Google Scholar] [CrossRef]
  5. Kang, Q.; Liu, S.H.; Yang, L.X.; Cai, Q.Y.; Grimes, C.A. Fabrication of PbS Nanoparticle-Sensitized TiO2 Nanotube Arrays and Their Photoelectrochemical Properties. ACS Appl. Mater. Interfaces 2011, 3, 746–749. [Google Scholar] [CrossRef] [PubMed]
  6. Yue, W.J.; Pan, Y.W.; Xie, Z.W.; Yang, X.; Hu, L.L.; Hong, L.R.; Tong, Y.F.; Cheng, Q.Q. Different depositing amount of CuInS2 on TiO2 nanoarrays for polymer/CuInS2-TiO2 solar cells. Mater. Sci. Semicond. Process. 2015, 40, 257–261. [Google Scholar] [CrossRef]
  7. Enesca, A.; Yamaguchi, Y.; Terashima, C.; Fujishima, A.; Nakata, K.; Duta, A. Enhanced UV-Vis photocatalytic performance of the CuInS2/TiO2/SnO2 hetero-structure for air decontamination. J. Catal. 2017, 350, 174–181. [Google Scholar] [CrossRef]
  8. Li, T.T.; Li, X.Y.; Zhao, Q.D.; Shi, Y.; Teng, W. Fabrication of n-type CuInS2 modified TiO2 nanotube arrays heterostructure photoelectrode with enhanced photoelectrocatalytic properties. Appl. Catal. B-Environ. 2014, 156, 362–370. [Google Scholar] [CrossRef]
  9. Shen, F.Y.; Que, W.X.; Liao, Y.L.; Yin, X.T. Photocatalytic Activity of TiO2 Nanoparticles Sensitized by CuInS2 Quantum Dots. Ind. Eng. Chem. Res. 2011, 50, 9131–9137. [Google Scholar] [CrossRef]
  10. Nanu, M.; Reijnen, L.; Meester, B.; Goossens, A.; Schoonman, J. CuInS2-TiO2 heterojunctions solar cells obtained by atomic layer deposition. Thin Solid Film. 2003, 431, 492–496. [Google Scholar] [CrossRef]
  11. Li, T.T.; Li, X.Y.; Zhao, Q.D.; Teng, W. Preparation of CuInS2/TiO2 nanotube heterojunction arrays electrode and investigation of its photoelectrochemical properties. Mater. Res. Bull. 2014, 59, 227–233. [Google Scholar] [CrossRef]
  12. Chen, B.F.; Niu, W.Z.; Lou, Z.R.; Ye, Z.Z.; Zhu, L.P. Improving the photovoltaic performance of the all-solid-state TiO2 NR/CulnS2 solar cell by hydrogen plasma treatment. Nanotechnology 2018, 29, 275402. [Google Scholar] [CrossRef]
  13. Fu, B.W.; Deng, C.; Yang, L. Efficiency Enhancement of Solid-State CuInS2 Quantum Dot-Sensitized Solar Cells by Improving the Charge Recombination. Nanoscale Res. Lett. 2019, 14, 198. [Google Scholar] [CrossRef]
  14. Jones, W.; Eddleston, M.D. Introductory Lecture: Mechanochemistry, a versatile synthesis strategy for new materials. Faraday Discuss. 2014, 170, 9–34. [Google Scholar] [CrossRef]
  15. Colacino, E.; Isoni, V.; Crawford, D.; Garcia, F. Upscaling Mechanochemistry: Challenges and Opportunities for Sustainable Industry. Trends Chem. 2021, 3, 335–339. [Google Scholar] [CrossRef]
  16. Kostova, N.; Dutkova, E. Mechanochemical synthesis and properties of ZnS/TiO2 composites. Bulg. Chem. Commun. 2016, 48, 161–166. [Google Scholar]
  17. Dutkova, E.; Balaz, M.; Daneu, N.; Tatykayev, B.; Karakirova, Y.; Velinov, N.; Kostova, N.; Briancin, J.; Balaz, P. Properties of CuFeS2/TiO2 Nanocomposite Prepared by Mechanochemical Synthesis. Materials 2022, 15, 6913. [Google Scholar] [CrossRef]
  18. Dutková, E.; Sayagues, M.J.; Briančin, J.; Zorkovská, A.; Bujňáková, Z.; Kováč, J.; Kováč, J.J.; Baláž, P.; Ficeriová, J. Synthesis and characterization of CuInS2 nanocrystalline semiconductor prepared by high-energy milling. J. Mater. Sci. 2016, 51, 1978–1984. [Google Scholar] [CrossRef]
  19. Coelho, A.A. TOPAS and TOPAS-Academic: An optimization program integrating computer algebra and crystallographic objects written in C plus. J. Appl. Crystallogr. 2018, 51, 210–218. [Google Scholar] [CrossRef]
  20. Evans, J.S.O. Advanced Input Files & Parametric Quantitative Analysis Using Topas. Mater. Sci. Forum 2010, 651, 1–9. [Google Scholar] [CrossRef]
  21. Baláž, P. Mechanochemistry in Nanoscience and Minerals Engineering; Springer: Berlin/Heidelberg, Germany, 2008; 413p. [Google Scholar]
  22. Tuschel, D. Molecular Spectroscopy Workbench Raman Spectroscopy and Polymorphism. Spectroscopy 2019, 34, 10–21. [Google Scholar]
  23. Gorai, S.; Bhattacharya, S.; Liarokapis, E.; Lampakis, D.; Chaudhuri, S. Morphology controlled solvothermal synthesis of copper indium sulphide powder and its characterization. Mater. Lett. 2005, 59, 3535–3538. [Google Scholar] [CrossRef]
  24. Alvarez-Garcia, J.; Barcones, B.; Perez-Rodriguez, A.; Romano-Rodriguez, A.; Morante, J.R.; Janotti, A.; Wei, S.H.; Scheer, R. Vibrational and crystalline properties of polymorphic CuInC2 (C = Se,S) chalcogenides. Phys. Rev. B 2005, 71, 054303. [Google Scholar] [CrossRef]
  25. Pawbake, A.; Waykar, R.; Jadhavar, A.; Kulkarni, R.; Waman, V.; Date, A.; Late, D.; Pathan, H.; Jadkar, S. Wide band gap and conducting tungsten carbide(WC) thin films prepared by hot wire chemical vapor deposition(HW-CVD) method. Mater. Lett. 2016, 183, 315–317. [Google Scholar] [CrossRef]
  26. Xu, F.Y.; Zhang, J.J.; Zhu, B.C.; Yu, J.G.; Xu, J.S. CuInS2 sensitized TiO2 hybrid nanofibers for improved photocatalytic CO2 reduction. Appl. Catal. B-Environ. 2018, 230, 194–202. [Google Scholar] [CrossRef]
  27. Balaz, M.; Dutkova, E.; Bujnakova, Z.; Tothova, E.; Kostova, N.G.; Karakirova, Y.; Briancin, J.; Kanuchova, M. Mechanochemistry of copper sulfides: Characterization, surface oxidation and photocatalytic activity. J. Alloys Compd. 2018, 746, 576–582. [Google Scholar] [CrossRef]
  28. Xie, B.B.; Hu, B.B.; Jiang, L.F.; Li, G.; Du, Z.L. The phase transformation of CuInS2 from chalcopyrite to wurtzite. Nanoscale Res. Lett. 2015, 10, 86. [Google Scholar] [CrossRef]
  29. Makula, P.; Pacia, M.; Macyk, W. How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV-Vis Spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef]
  30. Peng, Z.Y.; Liu, Y.L.; Shu, W.; Chen, K.Q.; Chen, W. Synthesis of Various Sized CuInS2 Quantum Dots and Their Photovoltaic Properties as Sensitizers for TiO2 Photoanodes. Eur. J. Inorg. Chem. 2012, 2012, 5239–5244. [Google Scholar] [CrossRef]
  31. Zhou, X.F.; Lu, J.; Jiang, J.J.; Li, X.B.; Lu, M.N.; Yuan, G.T.; Wang, Z.S.; Zheng, M.; Seo, H.J. Simple fabrication of N-doped mesoporous TiO2 nanorods with the enhanced visible light photocatalytic activity. Nanoscale Res. Lett. 2014, 9, 34. [Google Scholar] [CrossRef]
  32. Leach, A.D.P.; Macdonald, J.E. Optoelectronic Properties of CuInS2 Nanocrystals and Their Origin. J. Phys. Chem. Lett. 2016, 7, 572–583. [Google Scholar] [CrossRef]
  33. Omata, T.; Nose, K.; Otsuka-Yao-Matsuo, S. Size dependent optical band gap of ternary I-III-VI2 semiconductor nanocrystals. J. Appl. Phys. 2009, 105, 073106. [Google Scholar] [CrossRef]
  34. Ye, M.D.; Gong, J.J.; Lai, Y.K.; Lin, C.J.; Lin, Z.Q. High-Efficiency Photoelectrocatalytic Hydrogen Generation Enabled by Palladium Quantum Dots-Sensitized TiO2 Nanotube Arrays. J. Am. Chem. Soc. 2012, 134, 15720–15723. [Google Scholar] [CrossRef]
  35. Pandikumar, A.; Ramaraj, R. TiO2-Au nanocomposite materials modified photoanode with dual sensitizer for solid-state dye-sensitized solar cell. J. Renew. Sustain. Energy 2013, 5, 043101. [Google Scholar] [CrossRef]
  36. Ueng, H.Y.; Hwang, H.L. The Defect Structure of CuInS2. part I: Intrinsic Defects. J. Phys. Chem. Solids 1989, 50, 1297–1305. [Google Scholar] [CrossRef]
Figure 1. Scheme of the CuInS2/TiO2 nanocomposite preparation.
Figure 1. Scheme of the CuInS2/TiO2 nanocomposite preparation.
Crystals 14 00324 g001
Figure 2. XRD patterns of (a) CuInS2, (b) TiO2, (c) CuInS2/TiO2 4:1, and (d) CuInS2/TiO2 1:4 nanocomposites. The card numbers in the legend are taken from ICDD-PDF2 database.
Figure 2. XRD patterns of (a) CuInS2, (b) TiO2, (c) CuInS2/TiO2 4:1, and (d) CuInS2/TiO2 1:4 nanocomposites. The card numbers in the legend are taken from ICDD-PDF2 database.
Crystals 14 00324 g002
Figure 3. XRD patterns and the results of Rietveld refinement of CuInS2/TiO2 nanocomposite with CuInS2:TiO2 ratios (a) 4:1 and (b) 1:4.
Figure 3. XRD patterns and the results of Rietveld refinement of CuInS2/TiO2 nanocomposite with CuInS2:TiO2 ratios (a) 4:1 and (b) 1:4.
Crystals 14 00324 g003
Figure 4. Micro-Raman spectra of CuInS2, CuInS2/TiO2 4:1, 1:4 nanocomposites, and TiO2 (inset shows CuInS2/TiO2 1:4 long-range Raman spectrum).
Figure 4. Micro-Raman spectra of CuInS2, CuInS2/TiO2 4:1, 1:4 nanocomposites, and TiO2 (inset shows CuInS2/TiO2 1:4 long-range Raman spectrum).
Crystals 14 00324 g004
Figure 5. Low-magnification TEM images of CuInS2/TiO2 nanocomposites mixed in (a) 4:1 and (b) 1:4 ratios. Indexed SAED patterns, energy-dispersive X-ray spectra (EDXS), and high-resolution TEM (HRTEM) images of 4:1 and 1:4 nanocomposites are shown in (c,d), respectively.
Figure 5. Low-magnification TEM images of CuInS2/TiO2 nanocomposites mixed in (a) 4:1 and (b) 1:4 ratios. Indexed SAED patterns, energy-dispersive X-ray spectra (EDXS), and high-resolution TEM (HRTEM) images of 4:1 and 1:4 nanocomposites are shown in (c,d), respectively.
Crystals 14 00324 g005
Figure 6. SEM of CuInS2/TiO2 4:1 (a) and 1:4 (b) nanocomposites.
Figure 6. SEM of CuInS2/TiO2 4:1 (a) and 1:4 (b) nanocomposites.
Crystals 14 00324 g006
Figure 7. EDS mapping of CuInS2/TiO2 4:1 (A) and 1:4 (B) nanocomposites. (a) EDS layered image including all elements and elemental maps of (b) Cu, (c) In, (d) S, (e) Ti, and (f) O.
Figure 7. EDS mapping of CuInS2/TiO2 4:1 (A) and 1:4 (B) nanocomposites. (a) EDS layered image including all elements and elemental maps of (b) Cu, (c) In, (d) S, (e) Ti, and (f) O.
Crystals 14 00324 g007
Figure 8. XPS survey spectrum (a) and high-resolution XPS spectra of mechanochemically synthesized CuInS2/TiO2 4:1 (A) and 1:4 (B) nanocomposite: (b)—Cu 2p core level, (c)—In 3d core level, (d)—S 2p core level, (e) Ti—2p core level, and (f)—O 1s core level.
Figure 8. XPS survey spectrum (a) and high-resolution XPS spectra of mechanochemically synthesized CuInS2/TiO2 4:1 (A) and 1:4 (B) nanocomposite: (b)—Cu 2p core level, (c)—In 3d core level, (d)—S 2p core level, (e) Ti—2p core level, and (f)—O 1s core level.
Crystals 14 00324 g008
Figure 9. (a) UV–Vis reflectivity spectra of all samples (inset 9a shows details of the measurement of samples with low reflectivity) and (b) calculated (F(R) hν)2 and (F(R) hν)1/2 dependence for bandgap determination of CuInS2/TiO2 1:4 and TiO2 (inset 9b is the details of Eg determination for the CuInS2 sample).
Figure 9. (a) UV–Vis reflectivity spectra of all samples (inset 9a shows details of the measurement of samples with low reflectivity) and (b) calculated (F(R) hν)2 and (F(R) hν)1/2 dependence for bandgap determination of CuInS2/TiO2 1:4 and TiO2 (inset 9b is the details of Eg determination for the CuInS2 sample).
Crystals 14 00324 g009
Figure 10. (a) PL spectra of CuInS2, CuInS2/TiO2 4:1, 1:4 nanocomposites, and TiO2 and (b) PL spectra of CuInS2 vs. energy.
Figure 10. (a) PL spectra of CuInS2, CuInS2/TiO2 4:1, 1:4 nanocomposites, and TiO2 and (b) PL spectra of CuInS2 vs. energy.
Crystals 14 00324 g010
Figure 11. (a) Measured I–V characteristics of CuInS2, CuInS2/TiO2 4:1, and 1:4 nanocomposites and TiO2 and (b) details of CuInS2, CuInS2/TiO2 4:1, and 1:4 nanocomposites in the dark state and under light illumination.
Figure 11. (a) Measured I–V characteristics of CuInS2, CuInS2/TiO2 4:1, and 1:4 nanocomposites and TiO2 and (b) details of CuInS2, CuInS2/TiO2 4:1, and 1:4 nanocomposites in the dark state and under light illumination.
Crystals 14 00324 g011
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dutkova, E.; Baláž, M.; Kováč, J.; Daneu, N.; Kashimbetova, A.; Briančin, J.; Kováč, J., Jr.; Kováčová, S.; Čelko, L. Optical and Optoelectrical Properties of Ternary Chalcogenide CuInS2/TiO2 Nanocomposite Prepared by Mechanochemical Synthesis. Crystals 2024, 14, 324. https://doi.org/10.3390/cryst14040324

AMA Style

Dutkova E, Baláž M, Kováč J, Daneu N, Kashimbetova A, Briančin J, Kováč J Jr., Kováčová S, Čelko L. Optical and Optoelectrical Properties of Ternary Chalcogenide CuInS2/TiO2 Nanocomposite Prepared by Mechanochemical Synthesis. Crystals. 2024; 14(4):324. https://doi.org/10.3390/cryst14040324

Chicago/Turabian Style

Dutkova, Erika, Matej Baláž, Jaroslav Kováč, Nina Daneu, Adelia Kashimbetova, Jaroslav Briančin, Jaroslav Kováč, Jr., Soňa Kováčová, and Ladislav Čelko. 2024. "Optical and Optoelectrical Properties of Ternary Chalcogenide CuInS2/TiO2 Nanocomposite Prepared by Mechanochemical Synthesis" Crystals 14, no. 4: 324. https://doi.org/10.3390/cryst14040324

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