Next Article in Journal
3D Printed and Embedded Strain Sensors in Structural Composites for Loading Monitoring and Damage Diagnostics
Next Article in Special Issue
Synthesis and Properties of Rubidium Salts of Phosphotungstic Acid
Previous Article in Journal
Enhancing Stiffness, Toughness, and Creep in a 3D-Printed Bio-Based Photopolymer Using Ultra-Low Contents of Nanofibrillated Cellulose
Previous Article in Special Issue
Study on the Impact of a Combination of Synthetic Wollastonite and 2-Mercaptobenzothiazole-Based Fillers on UHMWPE Polymeric Matrix
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermoelectric and Magnetic Properties and Electronic Structure of Solid Solutions CuCr1-xLaxS2

by
Evgeniy V. Korotaev
*,
Mikhail M. Syrokvashin
and
Irina Yu. Filatova
Nikolaev Institute of Inorganic Chemistry, Siberian Branch, Russian Academy of Sciences, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2023, 7(10), 436; https://doi.org/10.3390/jcs7100436
Submission received: 15 September 2023 / Revised: 29 September 2023 / Accepted: 11 October 2023 / Published: 13 October 2023

Abstract

:
The oxidation states of atoms in CuCr1-xLaxS2 (x = 0–0.03) solid solutions were determined using the analysis of Cu2p, Cr2p, S2p, and La3d core level binding energy. The cationic substitution did not significantly affect the charge distribution on matrix elements (Cu, Cr, and S). The oxidation states of the atoms were identified as S2− for sulfur, Cu+ for copper, and Cr3+ for chromium. The cationic substitution in CuCr1-xLaxS2 was found to occur via the isovalent principle. The cationic substitution of CuCrS2 matrix with lanthanum ions led to the enhancement of the Seebeck coefficient comparing CuCr1-xLaxS2 to the initial matrix. The observed enhancement was attributed to the reconstruction of the valence band electronic structure after the cationic substitution. The maximum Seebeck coefficient value of 412 μV/K was measured for CuCr0.985La0.015S2 at 420 K. An increase in the lanthanum concentration to x = 0.03 caused a suppression of the Seebeck coefficient. The synthetic route was found to significantly affect both the magnetic properties and charge carrier concentration. The magnetic properties of CuCr1-xLaxS2 synthesized using metal sulfide reagents cannot be interpreted using the simple isovalent Cr3+ to La3+ cationic substitution model. The defectiveness of the samples and the formation of the impurity CuLaS2 phase could be additional factors that affect the magnetic properties of CuCr1-xLaxS2.

1. Introduction

Rapid technical progress demands the development of advanced functional materials and the improvement of the properties of existing materials through a series of research progress. In this context, multipurpose functional materials such as graphene, carbon nanotubes, and layered compounds based on transition metal chalcogenides are of special interest. Layered copper-chromium disulfide CuCrS2- and CuCrS2-based cation-substituted solid solutions can be considered as multipurpose advanced functional materials. These compounds exhibit various promising functional properties and effects, including thermoelectricity [1,2,3,4,5,6,7], ionic conductivity [8,9,10] and the order-to-disorder transition (ODT) [9,11,12,13], magnetoresistance [14,15] and the metal-to-insulator transition (MIT) [14,16], helimagnetic arrangement [17,18,19], light-absorbing [20] and luminescence properties [21], and ferroelectricity [22,23]. The combination of functional properties in CuCrS2-based solid solutions makes them promising materials for the fabrication of various electronic devices, including temperature sensors, thermoelectric generators (TEGs) [2,5,6,7,24], solid-state current sources [8,9], magnetic field sensors [14], solid-state memory [18,22,23], and electric or magnetic-field-driven valves [14,22,23].
The thermoelectric properties of CuCrS2-based compounds are of special interest due to the presence of ionic conductivity. The combination of thermoelectric and superionic properties allows one to consider these compounds as phonon liquid electronic crystal (PLEC) materials. PLEC materials, such as sulfide- or selenide-based compounds, possess a high Seebeck coefficient. This is due to the fact that these compounds have a fixed crystal matrix, resulting in low thermal conductivity related to the presence of mobile cations effectively scattering the phonons [25,26,27,28,29,30,31]. In the case of CuCrS2-based solid solutions, the chromium sublattice formed by the -S-Cr-S- layers can be considered as the “fixed” crystal matrix. At high temperatures, the mobile copper cations can migrate between different crystallographic sites in the copper sublattice and, thereby, can be considered as a phonon liquid [9,12,17]. The cationic substitution of the matrix elements (Cu, Cr or S) affects the Seebeck coefficient and ionic conductivity of CuCrS2-based compounds [3,9,10,32,33,34,35,36]. The cationic substitution of Cr atoms in the low-dopant concentration range (x < 0.03) increases both the Seebeck coefficient value and the ionic conductivity of CuCr1-xMxS2 solid solutions (M—transition metal or lanthanide atom) [3,33,35,36]. A systematic study focused on the thermoelectric properties and electronic structure of lanthanide-doped CuCr0.99Ln0.01S2 solid solutions revealed a notable enhancement in the Seebeck coefficient value after cationic substitution [32,33,35,36]. The most significant enhancement of the thermoelectric properties was reported for the La-doped CuCr1-xLaxS2 solid solutions [33,35,36]. Note that the Seebeck coefficient can be considered as one of the key parameters of the efficient thermoelectric materials used in temperature sensors and TEGs [37]. The critical analysis of the reported data concerning the study of functional properties of CuCrS2-based solid solutions indicates that the synthesis route and sample treatment significantly impact the thermoelectric, electrophysical, and magnetic properties of these materials [1,3,4,14,16,21,38,39,40,41]. For instance, the overwhelming majority of the reported studies were carried out on samples synthesized using commercial metal oxides as the initial reagents to obtain CuCrS2-matrix and CuCrS2-based solid solutions [3,32,33,34,39]. In particular, the series of the lanthanum-substituted solid solutions CuCr1-xLaxS2 (x = 0–0.03) was previously obtained and characterized using powder X-ray diffraction, static magnetochemistry, Seebeck coefficient, and Hall voltage measurements [42]. However, the final synthesis product had an additional impurity CuLaS2 phase in the high-dopant-concentration region for CuCr1-xLaxS2 (x > 0.01) solid solutions. The presence of the impurity phase affected both the thermoelectric and magnetic properties of CuCr1-xLaxS2 solid solutions. In this regard, it was suggested to vary the initial reagent compounds and study their influence on the formation of CuLaS2 phase in the final synthesis products and, therein, the thermoelectric and magnetic properties of CuCr1-xLaxS2 (x = 0–0.03) solid solutions.

2. Experimental

The initial sulfide reagents Cr2S3 and La2S3 were synthesized using commercial metal oxides Cr2O3 and La2O3 with a purity of 99.99%. The metal oxide powder in a glassy carbon boat was placed in a horizontal high-temperature quartz tube furnace. The sulfidation of the oxide powder was carried out with the gaseous decomposition products of NH4SCN at a temperature of 1050°C. High-purity argon gas was used as the carrier gas. During the sulfidation procedure, the product was ground for several times. The completeness of sulfidation was controlled at each grinding stage through powder X-ray diffraction (XRD) using a non-monochromatic CuKα X-ray radiation source (λ = 1.5406 Å)on a Bruker D8 Advance diffractometer (Bruker, Berlin, Germany). Then, using the synthesized sulfide powders of Cr2S3, La2S3, and the commercial Cu2S with a purity of 99.99%, the final CuCr1-xLaxS2 (x = 0, 0.005, 0.01, 0.015, and 0.03) products were synthesized. The synthesis route was the same as for the preparation of the initial sulfide powders. The phase composition and the formation of the final CuCr1-xLaxS2 product were analyzed using XRD.
The X-ray photoelectron spectroscopy (XPS) study of CuCr1-xLaxS2 was conducted using an ESCALAB 220i spectrometer (Thermo Fisher Scientific, Kingfisher, UK). The XPS lines of copper (Cu2p), lanthanum (La3d), and sulfur (S2p) were recorded at room temperature with a non-monochromatic AlKα X-ray source (hν = 1486.6 eV). The chromium (Cr2p) lines were recorded using a non-monochromatic MgKα X-ray source (hν = 1253.6 eV). The sample was fixed on a substrate using double-sided adhesive conductive carbon tape. The energy scale of the spectrometer was calibrated based on the line positions of metallic gold Au4f7/2 (84.0 eV) and copper Cu2p3/2 (932.6 eV). The measured binding energy (BE) values were corrected using the carbon C1s-line (284.8 eV) corresponding to adventitious carbon atoms in the near-surface layers of the samples. The XPS data were processed using a CasaXPS 2.3.15 software [43]. The measurement accuracy of BE values was of 0.2 eV.
The magnetic properties of CuCr1-xLaxS2 were investigated over an extended temperature range of 80–750 K using a Faraday balance-type experimental setup. A powder sample of ~20 mg in an open-type quartz cucurbit was placed in the measurement cell volume. Subsequently, the system was evacuated to a pressure of 0.01 Torr. Following this, the measurement cell was filled with helium to a pressure of 5 Torr. The temperature within the measurement cell volume was maintained using a Delta DTB9696 temperature controller. The voltage generated by the torsion-type quartz microbalance was measured using a high-precision digital 6½ Keysight 34465 A voltmeter (Keysight Technologies, Santa-Rosa, CA, USA). The magnetic susceptibility data were treated using the additive Pascal scheme. The potential presence of ferromagnetic impurities was controlled by the inverse field dependence of the magnetic susceptibility χ (1/H). The effective magnetic moment was calculated as follows:
μ e f f ( T ) 8 · χ T
The temperature dependence of the Seebeck coefficient for CuCr1−xLaxS2 was measured using a self-designed experimental setup. The ceramic samples were prepared by compressing the synthesized powder under a uniaxial pressure of 70 MPa for 2 h in a vacuum of 5 × 10−5 Torr at 923 K. The ceramic sample was positioned within the measurement cell, precisely situated between copper contact pads equipped with integrated 50 W nichrome wire heaters. Then, the measurement cell was evacuated and subsequently filled with helium with a pressure of 5 Torr. The thermoelectric power generated by the sample was measured using a 6½ Keysight 34465 A voltmeter (Keysight Technologies, Santa-Rosa, CA, USA). The experimental setup was validated using a constantan reference sample of thermocouple grade. The temperature of the contact pads was maintained by a Thermodat-13K5 temperature controller (LLC RPE Control Systems, Perm, Russia) using a platinum RTD sensor. During the measurements, the direction of the temperature gradient was systematically altered as +5, 0, and −5 K. Thus, the total Seebeck coefficient value was calculated as a slope of the voltage generated by the sample as a function of the temperature gradient.
Hall voltage measurements were conducted at ambient room temperature using the Van der Pauw technique. A direct current (DC) magnetic field of 1 T was applied perpendicularly to both the direction of the electrical current and the sample plane. A current of 10 mA was passed through the sample. The Hall voltage was measured using a 6½ Keysight 34465 A voltmeter. Throughout the measurement procedure, a systematic reversal of magnetic field polarity and current direction was carried out, followed by a subsequent interchange of the current and potential measurement probes. The Hall voltage value was determined through eight independent measurements. The polarity of the Hall voltage was calibrated using reference samples of n- and p-type silicon wafers.

3. Results and Discussion

The XRD data for the powder samples studied are plotted in Figure 1. The solid solutions CuCr1-xLaxS2 synthesized using the initial sulfide reagents and the initial CuCrS2 matrix are isostructural (R3m space group). The XRD data are consistent with the reference data for CuCrS2 of the ICSD database [44]. However, the additional weak diffraction peak at 23.6° could be observed for the CuCr0.97La0.03S2 solid solution (marked with * symbol in Figure 1). The corresponding diffraction peak was attributed previously to CuLaS2 impurity phase for the samples obtained using initial metal oxide reagents [42]. Note that the corresponding peak in [42] occurred in the solid solutions with a lower lanthanum concentration range (x > 0.01). Thus, one can conclude that the solubility limit of La in CuCrS2-matrix is 1.5 at.% greater than it was previously suggested in [42]. However, the lattice parameters increased for solid solutions with x ≤ 0.01 and decreased at a higher lanthanum concentration range (Table 1). Thus, one can conclude that the first trend is related to the isovalent cationic substitution of Cr3+ with La3+, while the second trend corresponds to the formation of CuLaS2 phase [42]. Consequently, the formation of CuLaS2 phase also occurred at x ≥ 0.01. Note that the absence of the diffraction peak of LaCrS2 for CuCr0.985La0.015S2 solid solution could be due to the low concentration of the impurity phase. The pure CuLaS2 phase was additionally synthesized as the reference sample for XPS measurements. The XRD data for CuLaS2 synthesized powder sample is depicted in Figure 1. The obtained CuLaS2 sample consisted of particles with different space groups (P21/c, P1121/b, and P63) [42]. However, the P21/c phase was prevalent.
The oxidation state of atoms in complex chemical compounds is an important parameter of particular interest. This is due to the fact that besides the electronic density distribution, the oxidation state is related to magnetic properties, magnetoresistance, and luminescence [14,45,46]. Furthermore, the oxidation state of dopant atoms affects the carrier concentration. For instance, if the oxidation state of dopant atoms is lower (or higher) compared to matrix atoms, thereby, the concentration of electrons (or holes) could be increased. X-ray photoelectron spectroscopy (XPS) is one of the most effective physical methods providing data on the oxidation state of atoms. The binding energy (BE) of core levels is related to the charge localized on the investigated atom [47]. Therefore, XPS allows one to selectively study the oxidation state of elements, even in the case of complex multi-element compounds such as CuCrS2-based solid solutions [41].
The XPS core level regions of copper, chromium, lanthanum, and sulfur in CuCr1-xLaxS2 solid solutions and CuLaS2 impurity phase are plotted in Figure 2 and Figure 3. The measured BEs of Cu2p-, Cr2p-, S2p-, and La3d-lines are listed in Table 2.
The Cu2p-region is represented by two intense lines related to the spin-orbit splitting of the Cu2p-level to Cu2p3/2- and Cu2p1/2-sublevels (Figure 2a). Note that the intense satellite lines at ~940 eV, which were previously observed for CuCr0.99Ln0.01S2 synthesized from the oxide reagents [33,35,36], are absent in the Cu2p-region of the samples studied. The presence of the satellites is the characteristic feature that indicates the presence of the surface Cu2+ species in copper compounds and CuCrS2-based solid solutions [33,35,36,48]. The sharp symmetric line-shape observed in the measured Cu2p-lines additionally confirms the absence of the surface Cu2+ species. However, the reported data concerning the copper oxidation state in the initial CuCrS2-matrix also indicated the absence of the Cu2+ state in CuCrS2 synthesized using pure elements and metal-organic reagents [1,21,40]. A systematic study of the charge distribution in CuCrS2-based solid solutions using both the surface-sensitive XPS technique and bulk-sensitive techniques (X-ray spectroscopy and magnetochemistry) showed that Cu2+ was localized on the surface [32,33,35,36,39,49]. Thus, varying the synthesis route allows one to obtain samples with or without Cu2+ on the sample surface. The measured BE values of the main Cu2p3/2-line were in the range of 931.6 to 932.0 eV and correspond to Cu+ in sulfide and selenide compounds of copper (Cu2S, BE ≈ 932.0 ÷ 932.2 eV; Cu2Se, BE ≈ 931.0 ÷ 932.2 eV [50,51]). The lanthanum concentration increase did not significantly affect the BE of Cu2p3/2-lines in CuCr1-xLaxS2 solid solutions. Thus, one can conclude that the cationic substitution in CuCr1-xLaxS2 did not significantly affect the electronic density localized on the copper atoms. The measured BE of Cu2p3/2-lines in CuLaS2 was in the same energy range as that for CuCr1-xLaxS2. This fact indicates that CuLaS2 impurity phase did not significantly affect the Cu2p-region in CuCr1-xLaxS2. Note that the Cu2p3/2 BE values measured in the present study were shifted to the low-energy region compared to those for CuCrS2-based solid solutions obtained from the metal oxide initial reagents reported previously (CuCrS2, CuCr1-xMxS2, BE ≈ 932.2 ÷ 932.6 eV) [33,36]. This could be due to the significant covalence contribution of the M-S bonding compared to the M-O. For instance, the M-S bonding leads to greater localization of the electron density on copper atoms in the sulfides compared to oxide compounds [52]. Note that the BE values of Cu2p3/2-lines for CuCrS2 samples obtained using pure elements or metal-organic compounds were within the range of 932.0 to 932.1 eV [1,21,40]. Thus, one can conclude that the BE of Cu2p-lines could vary depending on the synthesis route.
The Cr2p-region for CuCr1-xLaxS2 is depicted in Figure 3a. The spin-orbit doublet of Cr2p1/2- and Cr2p3/2-lines can be observed, similar to the case of Cu2p-lines. However, the line-shape of the Cr2p-lines is asymmetric with a shoulder in the high-energy region. Thus, one can conclude that the Cr2p-region represents a superposition of two components. The low-energy line (marked as I in Figure 3a) with a BE of 574.2 to 574.5 eV could be associated with the Cr3+ in sulfide (Cr2S3 BE ≈ 574.4–575.4 eV) or selenide (Cr2Se3 BE ≈ 574.4–575.4 eV, CuCrSe2, BE ≈ 574.7 eV; CuCr2Se4, BE ≈ 574.5 eV) compounds of chromium [40,50,51]. As it was previously reported in [33,36], the high-energy line (marked as II in Figure 3a) with a BE of 576.6 to 577.2 eV is attributed to the oxide compounds of chromium on the sample surface (Cr2O3, BE ≈ 576.5 eV; CuCrO2, BE ≈ 576.0 eV [50,51]). Hence, the low-energy components are considered to correspond to the chromium atoms in CuCr1-xLaxS2 solid solutions. Thus, one can conclude that cationic substitution did not affect the BE of the Cr2p lines and, thereby, the electron density localized on the chromium atoms. Note that measured BE values were shifted to the low-energy region compared to those for CuCrS2-based solid solutions obtained from the metal oxide initial reagents reported previously (BE ≈ 574.6–574.8 [33,36]). In the case of CuCrS2 obtained using pure elements, the BE of the Cr2p3/2-line was within the range of 574.6 to 574.8 eV [1,40]. Of particular note is that in the case of CuCrS2 synthesized using metal-organic reagents, the BE of 576.6 eV, corresponding to the Cr2p3/2-line, was significantly shifted to the high-energy region [21]. The observed shift could indicate a significant oxidation of chromium atoms in the near-surface layers of the sample. However, the contribution of the II component prevailed. Thus, one can conclude that the Cr2p-region, as well as the Cu2p-region, was significantly affected by the synthesis route. Note that the integral intensity ratio of the components in the Cr2p-region of CuCr1-xLaxS2 was more variable than it was observed previously for CuCr0.99Ln0.01S2 (Ln = La…Lu) solid solutions synthesized using metal oxide reagents [33,35,36]. The most intense component II was observed for CuCr0.97Ln0.03S2 solid solution. This could be due to the highest concentration of CuLaS2 impurity phase, which results in increased defectiveness of the chromium sublattice and, consequently, the lighter oxidation of chromium atoms in the near-surface layers.
The S2p-region for CuCr1-xLaxS2 is represented by an unresolved line consisting of the S2p1/2- and S2p3/2-components of the spin-orbit doublet (Figure 2b). Note that the S2p-region exhibits two groups of lines. The first group of lines (denoted as I in Figure 2b) with a BE of 160.6–161.0 eV arises from the sulfur atoms in the composition of CuCr1-xLaxS2 and corresponds to the S2− state [33,36,50]. Thus, taking into account the measured BE values of the corresponding components, one can conclude that cationic substitution of the initial CuCrS2-matrix did not affect electron density localized on the sulfur atoms. The measured BE values correlate with those for transition metal sulfides (FeS, BE ≈ 160.8–161.4 eV; TiS2, BE ≈ 160.9 eV; CuFeS2, BE ≈ 161.5 eV [50,51]). The second group (denoted as II in Figure 2b) with a BE of 162.5–162.7 eV is associated with the sulfur atoms of polysulfide groups and elemental sulfur in the defective near-surface layers on the samples. Note that the presence of the corresponding sulfur surface species is typical for inorganic sulfur compounds [53,54]. The measured S2p-lines for CuCr1-xLaxS2 were shifted to the low-energy region compared to the samples synthesized using the metal oxide reagents (BE ≈ 160.9–161.5 eV [33,36,50]). Meanwhile, CuCrS2 samples synthesized using meta-organic reagents exhibited an intermediate S2p-line BE of 161.0 eV [21]. The S2p-region for the impurity CuLaS2 phase is represented with the same line-shape as the solid solutions studied (Figure 2b). The measured BE for CuLaS2 was slightly shifted in the low-energy spectral region. This fact allows one to conclude that the presence of the impurity CuLaS2 phase in the composition of the final CuCr1-xLaxS2 products could lead to a slight shift of the S2p-line to the low-energy region if the CuLaS2 concentration is high.
The La3d-region for CuCr1-xLaxS2 solid solutions is presented in Figure 3b. Note that the spin-orbit splitting value for the La3d-level is ~17 eV. Thus, the La3d5/2-line could be analyzed separately from La3d3/2. The La3d5/2-line exhibits two intense components (denoted as 1 and 2) arising due to the many-electron process [47]. The main low-energy line with a BE of 834.7–834.9 eV correlates well with the La3+ oxidation state (La2O3, BE ≈ 834.0–835.2 eV [50,51]). The absence of significant shifts of the La3d-line indicated the preservation of electron density localized on the lanthanum atoms in CuCr1-xLaxS2 with an increase in lanthanum concentration. However, the line-shape for CuCr1-xLaxS2 (x > 0.01) solid solutions changed. This resulted in the evolution of the line-shape, and for x = 0.03, the La3d5/2-line was similar to that for CuLaS2 and corresponded to the impurity phase concentration increase. Taking into account the relatively low total lanthanum concentration in comparison to other elements in CuCr1-xLaxS2, CuLaS2 could noticeably impact the line-shape even when the impurity phase concentration was low. The measured BE of La3d5/2-lines was shifted to the low-energy region compared to the previously investigated samples synthesized using the metal oxide reagents (CuCr0.99La0.01S2, BE ≈ 835.4 eV [35]).
Static magnetic measurements can provide data concerning the effective magnetic moment, the magnetic phase transition temperature, the presence of specific magnetic impurities, and the exchange interactions between the magnetic centers in the sample studied. Since the μeff of transition elements depends on their oxidation state, one could obtain data concerning the electronic density localized on the magnetic centers. However, in the case of complex chemical compounds containing a few different types of magnetic centers, the interpretation of μeff values could be significantly complicated due to the macroscopic behavior of the method [55,56]. Magnetochemistry data are used as a method to study charge distribution in CuCrS2-based solid solutions [10,11,14,34,38,39,42,57]. Note that the obtained data led to contradictory conclusions regarding the oxidation state of copper and chromium ions, even in the initial CuCuS2-matrix. For instance, studies have shown Cu+ and Cr2+ [9], the coexistence of Cu+/Cr3+ and Cu2+/Cr2+ distributions [14,58], and Cu+ and Cr3+ [57]. The systematic studies involving both the macroscopic static magnetochemistry measurements and element-selective XPS and X-ray spectroscopy techniques allow one to conclude that deviations from the simple charge balance model of Cu+Cr3+(S2−)2 could be attributed to the presence of magnetic impurities and surface oxidation state species [32,33,34,35,36,39]. However, the reported data could be attributed to the sample series obtained by using the same synthesis route using metal oxide initial reagents. Here, we report a study of magnetic properties for the samples synthesized using metal sulfide initial reagents and attempt to assess the potential influence of the synthesis route on the obtained data.
The magnetic properties of CuCr1-xLaxS2 solid solutions synthesized using the metal sulfide initial reagents are presented in Figure 4. It was previously reported that the inverse magnetic susceptibility (1/χ) of similar samples synthesized using metal oxide initial reagents showed a linear dependence on temperature within the temperature range of 80 to 600 K [42]. However, the solid solutions CuCr0.995La0.005S2 and CuCr0.985La0.015S2 demonstrated deviations from linear dependence. The observed deviation consequently affected the behavior of the χ(T) and μeff(T) dependencies compared to the other samples. The most significant deviation was observed at temperatures below 400K. The presence of magnetic impurities leads to the overestimation of μeff [55]. The positive slope of χ as the function of the inverse value of the applied magnetic field strength for CuCr0.995La0.005S2 and CuCr0.985La0.015S2 indicated the influence of ferromagnetic impurities at T < 400 K (as shown in the inset in Figure 4a) [55]. The behavior of χ(T), 1/χ(T), and μeff(T) dependencies (marked as 0.005* and 0.015* in Figure 4) was similar to that for the other samples after the correction procedure. The corrected and non-corrected data correlate well at T > 400 K. Note that other samples did not contain ferromagnetic impurities. The magnetic susceptibility of the impurity CuLaS2 phase was negative. This fact indicated that CuLaS2 is diamagnetic. The absolute value of χ for CuLaS2 was two orders of magnitude lower compared to the main CuCr1-xLaxS2 phase. Thus, one can conclude that the presence of the diamagnetic impurity did not significantly affect the magnetic properties of the main phase. That fact correlates well with the previously reported data [42].
A slight decrease in the χ and μeff values is observed (Figure 4a and inset in Figure 4c, respectively) within the temperature range of 600–750 K. It was previously reported that the observed behavior is associated with the order-disorder phase transition (ODT) [34,42]. Hence, one can conclude that the data on magnetic properties can provide additional information on the order–disorder phase transition temperature for the samples studied. The concentration dependency of the μeff minimum temperature for CuCr1-xLaxS2 is shown in Figure 5. A general trend of the μeff minimum temperature decreasing as a function of lanthanum concentration is observed. It was previously suggested that the observed behavior could be associated with a decrease in the defect formation energy [42]. Thus, the dependencies of the μeff minimum temperature for the samples synthesized from metal sulfide reagents correlate with those reported for the samples obtained using metal oxide reagents.
The temperature dependencies of χ after the correction procedure were approximated according to the Curie–Weiss law (solid lines in Figure 4) [55]. The concentration dependencies of the μeff and Weiss constant (Θ) are presented in Figure 6a and Figure 6b, respectively. The μeff value of 3.77 μB measured for the initial CuCrS2-matrix correlates well with the experimental values of the μeff for Cr3+ compounds. However, the obtained value is slightly less than the theoretical value of 3.87 μΒ [55,56]. The obtained value is less compared to the sample synthesized using metal oxide reagents [42]. Note that the concentration dependency of the μeff for CuCr1-xLaxS2 shown in Figure 6 deviates from that reported in [42]. The μeff increased for CuCr1-xLaxS2 (x ≤ 0.01), while in the case of samples obtained using metal oxide regents, it decreased. This fact contradicts the simple model of the cationic substitution of paramagnetic Cr3+ with diamagnetic La3+ ions, which assumes a μeff decrease. Note that the oxidation states of metal atoms determined using XPS were Cu+, Cr3+, and La 3+. Thus, only the Cr3+ ions carry the magnetic moment due to the presence of unpaired electrons on the 3d-shell. Hence, the slight variation of the μeff value is attributed to the fluctuation of the electronic density localized on chromium atoms. This could be due to the sample defectiveness resulting from the incorporation of La3+ ions during the cationic substitution, as suggested in previous studies [42]. The μeff value decrease for x > 0.01 could be related to the diamagnetic impurity CuLaS2 phase formation [42]. The concentration dependencies of the Weiss constant Θ can be represented by two characteristic ranges of x ≤ 0.01 and x > 0.01, respectively. The first one corresponds to a notable decrease in the absolute value of Θ, and the second one is associated with minor variations in the absolute value of Θ. In terms of the molecular field theory, the Θ value is associated with the spin localized on the magnetic center and the absolute value of the total magnetic exchange interaction |∑Jizi| [55,56]. Since the La3+ ions are diamagnetic, the |∑Jizi| could be simply calculated (Figure 6c) as was described in [39,42]. The |∑Jizi| value increased for CuCr1-xLaxS2 (x ≤ 0.01). This could be due to the increase in electronic density localized on the chromium atoms and, consequently, the overlap of the electron cloud of the chromium atoms with the electron densities of the neighboring atoms. The second concentration range x > 0.01 corresponds to the formation of the impurity CuLaS2 phase.
In addition to the data on the atom oxidation state, XPS spectroscopy allows one to obtain experimental data concerning the valence band (VB) structure [47]. The VB structure and the partial density of states (DOS) for the initial CuCrS2-matrix and CuCr0.99La0.01S2 solid solutions obtained using oxygen metal initial reagents were discussed previously [3,35]. An example of the theoretical VB calculation based on the DOS calculations for the initial CuCrS2-matrix is presented in Figure 7a. It was previously demonstrated that the simple broadening of the total density of states (denoted “DOS” in Figure 7a) using the AlKα-line width (“DOS br.” in Figure 7a) did not yield a good agreement between the theoretical and experimental data [35]. Taking into account the photoionization cross-section value (σph), one could recalculate the partial contribution in the total DOS (“DOS x σph” in Figure 7a) and broaden it using the AlKα-line width (“DOS x σph br.” in Figure 7a). The line-shape of the calculated VB correlates well with the experimental VB. Since the copper atoms have a higher σph value, the experimental VB is mainly determined by the copper states, resulting in a triangle-like line-shape. For the calculation of the lanthanide-doped CuCr1-xLaxS2, the variation in lanthanum concentration was taken into account by a simple recalculation of the partial contribution of the lanthanum DOS reported previously [35]. The theoretical VB spectra for CuCr1-xLaxS2(x = 0.005–0.03) solid solutions are presented in Figure 7b. The similarity of the line-shape of both theoretical and experimental VB (Figure 7c) allows one to conclude that cationic substitution did not significantly affect the partial DOS distribution in the VB of CuCr1-xLaxS2. This could be due to the low lanthanum concentration and, therein, low contributions of the lanthanum states in the VB structure. Furthermore, the lanthanum has an unfilled 4f-shell. Thus, the contribution of the lanthanum states is low, especially when compared to the contribution of the copper states. The experimental VB of the impurity CuLaS2 phase has a similar shape to the one for CuCr1-xLaxS2 solid solutions (Figure 7c). Thus, one can conclude that the presence of - CuLaS2 impurity in the composition of CuCr1-xLaxS2 samples did not significantly distort the VB structure.
The Seebeck coefficient (S) temperature dependencies for CuCr1-xLaxS2 solid solutions studied are plotted in Figure 8a. The S values measured in the current study are higher compared to the previously reported data for the samples obtained using metal oxide initial reagents [42]. Thus, the Seebeck coefficient of CuCrS2-based solid solutions could be increased by changing the synthesis route. In the case of initial metal oxide reagents, during the sulfidation procedure, the oxygen atoms in the reaction mixture are replaced by the sulfur atoms from the sulfidation atmosphere. When one uses sulfide metal initial reagents, the sulfidation atmosphere prevents sulfur depletion from the reaction mixture. Hence, in the first route, some residual oxygen could be preserved after the sulfidation procedure in the composition of samples. The residual oxygen atoms could introduce additional dopant or trap centers, affecting the total charge carrier concentration and S values. Note that the sample defectiveness increase discussed above could also impact the total carrier concentration. The Seebeck coefficient of CuCr1-xLaxS2 tended to saturate at high temperatures (T > 400 K). It should be noted that using the metal oxides as the initial reagents resulted in a shift of the saturation to a higher temperature region (T > 500 K) [42] due to the presence of the residual oxygen affecting the total carrier concentration [45]. The charge carrier concentration measured at room temperature is plotted in Figure 7b. The carrier concentration (holes according to the Hall voltage polarity) measured for the initial CuCrS2-matrix (1.5·1017) is lower compared to that previously reported for the samples obtained using metal oxygen reagents (6.8·1017) [42].
As was reported previously [32,33,35,36,42], the cationic substitution of chromium atoms in CuCrS2-matrix with lanthanum atoms leads to an increase in the S value, as observed in Figure 8. The maximum S value of 412 μV/K was measured for CuCr0.985La0.015S2 at 420 K. The obtained value is higher compared to the maximum value of 373 μV/K reported in [42]. Further increasing the lanthanum concentration to x = 0.03 caused a suppression of the Seebeck coefficient. The observed concentration behavior of the Seebeck coefficient correlated with electronic structure reconfiguration reported for lanthanum-doped and CuCrS2-based solid solutions [3,32,33,35,36,42]. For instance, at low concentrations x ≤ 0.015, the S value of CuCr1-xLaxS2 increases due to the shift of the valence band top to a higher binding energy region. The further concentration increase gradually leads to the band gap narrowing, ultimately resulting in it completely vanishing at high dopant concentrations [3,34]. The corresponding model is in good agreement with the measured carrier concentrations presented in Figure 8b. The carrier concentration decreasing at x = 0.005 could be related to the shift of the valence band top. The carrier concentration increased at x ≥ 0.01 due to the band gap narrowing. The deviation of the carrier concentration from the increasing trend at x = 0.015 could be attributed to the presence of the CuLaS2 impurity. Thus, the synthesis route could affect both the Seebeck coefficient value and carrier concentration. The variation of the initial reagents from the metal oxides to sulfides initial reagents could be utilized to increase the Seebeck coefficient.

4. Conclusions

A comprehensive study involving XPS spectroscopy, static magnetochemistry, the Seebeck coefficient, and the Hall voltage measurements for the lanthanum-substituted solid solutions CuCr1-xLaxS2 (x = 0, 0.005, 0.01, 0.015, and 0.03) synthesized using metal sulfide initial regents was carried out. It was shown that the cationic substitution occurs via the isovalent principle Cr3+→La3+ for all lanthanum concentrations. The cationic substitution did not affect the charge distribution on the matrix elements Cu+, Cr3+, and S2−. The most significant effect of the impurity CuLaS2 phase was observed in the Ln3d-region. The measured BE of Cu2p-, Cr2p-, S2p- and La3d-lines were shifted to the low-energy region compared to the previously reported data for the CuCrS2-based solid solutions obtained using the metal oxide initial reagents. This could be due to the significant covalence contribution of the M-S bonding compared to the M-O. The magnetic properties of CuCr1-xLaxS2 solid solutions synthesized using metal sulfide reagents cannot be interpreted using the simple isovalent Cr3+ to La3+ cationic substitution model. The defectiveness of the samples and the formation of the impurity CuLaS2 phase could be additional factors that affect the magnetic properties of CuCr1-xLaxS2 solid solutions. The concentration dependencies of the μeff, Θ, and |∑Jizi| were significantly influenced by the synthesis route. The cationic substitution did not significantly affect the partial density of states distribution in the experimental valence band structure of CuCr1-xLaxS2 solid solutions. The maximum Seebeck coefficient value of 412 μV/K was measured for CuCr0.985La0.015S2 at 420 K. An increase in the lanthanum concentration to x = 0.03 caused a suppression of the Seebeck coefficient. The variation of the initial reagents from the metal oxides to sulfides initial reagents can be utilized to increase the Seebeck coefficient.

Author Contributions

Conceptualization, E.V.K. and M.M.S.; Methodology, E.V.K., M.M.S. and I.Y.F.; Validation, E.V.K.; Formal analysis, M.M.S.; Investigation, E.V.K. and M.M.S.; Resources, I.Y.F. and E.V.K.; Data curation, E.V.K. and M.M.S.; Writing—original draft, E.V.K. and M.M.S.; Writing—review and editing, E.V.K. and M.M.S.; Visualization, E.V.K. and M.M.S.; Supervision, E.V.K.; Funding acquisition, E.V.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Russian Science Foundation (project No. 19-73-10073, https://rscf.ru/project/19-73-10073/ accessed 14 November 2022).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to the Ministry of Science and Higher Education of the Russian Federation. The authors thank Sotnikov A.V. for the assistance with obtaining the ceramic samples.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tewari, G.C.; Tripathi, T.S.; Rastogi, A.K. Thermoelectric properties of layer-antiferromagnet CuCrS2. J. Electron. Mater. 2010, 39, 1133–1139. [Google Scholar] [CrossRef]
  2. Tewari, G.C.; Tripathi, T.S.; Rastogi, A.K. Effect of chromium disorder on the thermoelectric properties of layered-antiferromagnet CuCrS2. Z. Für Krist. 2010, 225, 471–474. [Google Scholar] [CrossRef]
  3. Korotaev, E.V.; Syrokvashin, M.M.; Filatova, I.Y.; Pelmenev, K.G.; Zvereva, V.V.; Peregudova, N.N. Seebeck Coefficient of Cation-Substituted Disulfides CuCr1−xFexS2 and Cu1−xFexCrS2. J. Electron. Mater. 2018, 47, 3392–3397. [Google Scholar] [CrossRef]
  4. Kaltzoglou, A.; Vaqueiro, P.; Barbier, T.; Guilmeau, E.; Powell, A.V. Ordered-Defect Sulfides as Thermoelectric Materials. J. Electron. Mater. 2014, 43, 2029–2034. [Google Scholar] [CrossRef]
  5. Romanenko, A.I.; Chebanova, G.E.; Katamanin, I.N.; Drozhzhin, M.V.; Artemkina, S.B.; Han, M.-K.; Kim, S.-J.; Wang, H. Enhanced thermoelectric properties of polycrystalline CuCrS2−xSex (x = 0, 0.5, 1.0, 1.5, 2) samples by replacing chalcogens and sintering. J. Phys. D Appl. Phys. 2021, 55, 135302. [Google Scholar] [CrossRef]
  6. Hansen, A.-L.; Dankwort, T.; Groß, H.; Etter, M.; König, J.; Duppel, V.; Kienle, L.; Bensch, W. Structural properties of the thermoelectric material CuCrS2 and of deintercalated CuxCrS2 on different length scales: X-ray diffraction, pair distribution function and transmission electron microscopy studies. J. Mater. Chem. C 2017, 5, 9331–9338. [Google Scholar] [CrossRef]
  7. Chen, Y.-X.; Zhang, B.-P.; Ge, Z.-H.; Shang, P.-P. Preparation and thermoelectric properties of ternary superionic conductor CuCrS2. J. Solid. State Chem. 2012, 186, 109–115. [Google Scholar] [CrossRef]
  8. Krengel, M.; Hansen, A.L.; Hartmann, F.; van Dinter, J.; Bensch, W. Elucidation of the Sodium—Copper Extrusion Mechanism in CuCrS2: A High Capacity, Long-Life Anode Material for Sodium-Ion Batteries. Batter. Supercaps 2018, 1, 176–183. [Google Scholar] [CrossRef]
  9. Almukhametov, R.F.; Yakshibayev, R.A.; Gabitov, E.V.; Abdullin, A.R.; Kutusheva, R.M. Structural properties and ionic conductivities of CuCr1–xVxS2 solid solutions. Phys. Status Solidi 2003, 236, 29–33. [Google Scholar] [CrossRef]
  10. Almukhametov, R. Structural properties and ionic conductivity of new CuCr1−xVxSe2 solid solutions. Solid. State Ion. 2003, 158, 409–414. [Google Scholar] [CrossRef]
  11. Vasilyeva, I.G. Chemical aspect of the structural disorder in CuCrS2 and CuCr1–xVxS2 solid solutions. J. Struct. Chem. 2017, 58, 1009–1017. [Google Scholar] [CrossRef]
  12. Vassilieva, I.G.; Kardash, T.Y.; Malakhov, V.V. Phase transformations of CuCrS2: Structural and chemical study. J. Struct. Chem. 2009, 50, 288–295. [Google Scholar] [CrossRef]
  13. Peng, J.; Chen, Z.-J.; Ding, B.; Cheng, H.-M. Recent Advances for the Synthesis and Applications of 2-Dimensional Ternary Layered Materials. Research 2023, 6, 40. [Google Scholar] [CrossRef] [PubMed]
  14. Abramova, G.M.; Petrakovskii, G.A. Metal-insulator transition, magnetoresistance, and magnetic properties of 3d-sulfides (Review). Low. Temp. Phys. 2006, 32, 725–734. [Google Scholar] [CrossRef]
  15. Rasch, J.C.E.; Boehm, M.; Ritter, C.; Mutka, H.; Schefer, J.; Keller, L.; Abramova, G.M.; Cervellino, A.; Löffler, J.F. Magnetoelastic coupling in the triangular lattice antiferromagnet CuCrS2. Phys. Rev. B 2009, 80, 104431. [Google Scholar] [CrossRef]
  16. Tsujii, N.; Kitazawa, H.; Kido, G. Insulator to metal transition induced by substitution in the nearly two-dimensional compound CuCr1–xVxS2. Phys. Status Solidi 2006, 3, 2775–2778. [Google Scholar] [CrossRef]
  17. Engelsman, F.M.R.; Wiegers, G.A.; Jellinek, F.; Van Laar, B. Crystal structures and magnetic structures of some metal (I) chromium (III) sulfides and selenides. J. Solid. State Chem. 1973, 6, 574–582. [Google Scholar] [CrossRef]
  18. Karmakar, A.; Dey, K.; Chatterjee, S.; Majumdar, S.; Giri, S. Spin correlated dielectric memory and rejuvenation in multiferroic CuCrS2. Appl. Phys. Lett. 2014, 104, 052906. [Google Scholar] [CrossRef]
  19. Winterberger, M.; Allain, Y.; Winterberger, M.A.Y. Structure magnetique helicoïdale de CuCrS2. Solid. State Commun. 1987, 64, 1343–1346. [Google Scholar] [CrossRef]
  20. Sanchez Rodriguez, J.J.; Nunez Leon, A.N.; Abbasi, J.; Shinde, P.S.; Fedin, I.; Gupta, A. Colloidal Synthesis, Characterization, and Photoconductivity of Quasi-Layered CuCrS2 Nanosheets. Nanomaterials 2022, 12, 4164. [Google Scholar] [CrossRef]
  21. Lee, H.K.; Ban, Y.J.; Lee, H.J.; Kim, J.H.; Park, S.J. One-Pot Synthesis and Characterization of CuCrS2/ZnS Core/Shell Quantum Dots as New Blue-Emitting Sources. Materials 2023, 16, 762. [Google Scholar] [CrossRef]
  22. Xu, X.; Zhong, T.; Zuo, N.; Li, Z.; Li, D.; Pi, L.; Chen, P.; Wu, M.; Zhai, T.; Zhou, X. High- TCTwo-Dimensional Ferroelectric CuCrS2 Grown via Chemical Vapor Deposition. ACS Nano 2022, 16, 8141–8149. [Google Scholar] [CrossRef] [PubMed]
  23. Zhong, T.; Li, X.; Wu, M.; Liu, J.M. Room-temperature multiferroicity and diversified magnetoelectric couplings in 2D materials. Natl. Sci. Rev. 2020, 7, 373–380. [Google Scholar] [CrossRef] [PubMed]
  24. Kousar, H.S.; Srivastava, D.; Karppinen, M.; Tewari, G.C. Tunable Low-Temperature Thermoelectric Transport Properties in Layered CuCr(S1-xSex)2 System. Z. Fur Anorg. Und Allg. Chem. 2023, 649, e202300079. [Google Scholar] [CrossRef]
  25. Wu, D.; Huang, S.; Feng, D.; Li, B.; Chen, Y.; Zhang, J.; He, J. Revisiting AgCrSe2 as a promising thermoelectric material. Phys. Chem. Chem. Phys. 2016, 18, 23872–23878. [Google Scholar] [CrossRef] [PubMed]
  26. Yin, Y.; Baskaran, K.; Tiwari, A. A Review of Strategies for Developing Promising Thermoelectric Materials by Controlling Thermal Conduction. Phys. Status Solidi Appl. Mater. Sci. 2019, 216, 6–8. [Google Scholar] [CrossRef]
  27. Zebarjadi, M.; Esfarjani, K.; Dresselhaus, M.S.; Ren, Z.F.; Chen, G. Perspectives on thermoelectrics: From fundamentals to device applications. Energy Environ. Sci. 2012, 5, 5147–5162. [Google Scholar] [CrossRef]
  28. Tan, X.; Ding, J.; Luo, H.; Delaire, O.; Yang, J.; Zhou, Z.; Lan, J.L.; Lin, Y.H.; Nan, C.W. High Thermoelectric Performance of AgSb1-xPbxSe2Prepared by Fast Nonequilibrium Synthesis. ACS Appl. Mater. Interfaces 2020, 12, 41333–41341. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, J.; Yan, X.; Zhang, Z.; Guo, R.; Ying, H.; Han, G.; Han, W.-Q. Rational Design of an Electron/Ion Dual-Conductive Cathode Framework for High-Performance All-Solid-State Lithium Batteries. ACS Appl. Mater. Interfaces 2020, 12, 41323–41332. [Google Scholar] [CrossRef]
  30. Li, S.; Hou, S.; Xue, W.; Yin, L.; Liu, Y.; Wang, X.; Chen, C.; Mao, J.; Zhang, Q. Manipulation of phase structure and Se vacancy to enhance the average thermoelectric performance of AgBiSe2. Mater. Today Phys. 2022, 27, 100837. [Google Scholar] [CrossRef]
  31. Wu, H.; Shi, X.; Duan, J.; Liu, Q.; Chen, Z.-G. Advances in Ag 2 Se-based thermoelectrics from materials to applications. Energy Environ. Sci. 2023, 16, 1870–1906. [Google Scholar] [CrossRef]
  32. Korotaev, E.V.; Syrokvashin, M.M.; Filatova, I.Y.; Trubina, S.V.; Nikolenko, A.D.; Ivlyushkin, D.V.; Zavertkin, P.S.; Sotnikov, A.V.; Kriventsov, V.V. XANES investigation of novel lanthanide-doped CuCr0.99Ln0.01S2 (Ln = La, Ce) solid solutions. Appl. Phys. A 2020, 126, 537. [Google Scholar] [CrossRef]
  33. Korotaev, E.V.; Syrokvashin, M.M.; Filatova, I.Y.; Sotnikov, A.V.; Kalinkin, A. V The Charge Distribution, Seebeck Coefficient, and Carrier Concentration of CuCr0.99Ln0.01S2 (Ln = Dy–Lu). Materials 2023, 16, 2431. [Google Scholar] [CrossRef] [PubMed]
  34. Korotaev, E.V.; Syrokvashin, M.M.; Filatova, I.Y.; Zvereva, V.V. Magnetic Properties of Novel Layered Disulfides CuCr0.99Ln0.01S2 (Ln = La…Lu). Materials 2021, 14, 5101. [Google Scholar] [CrossRef] [PubMed]
  35. Korotaev, E.V.; Syrokvashin, M.M.; Filatova, I.Y.; Kalinkin, A.V.; Sotnikov, A.V. Valence band structure and charge distribution in the layered lanthanide-doped CuCr0.99Ln0.01S2 (Ln = La, Ce) solid solutions. Sci. Rep. 2021, 11, 18934. [Google Scholar] [CrossRef]
  36. Korotaev, E.V.; Syrokvashin, M.M.; Filatova, I.Y.; Sotnikov, A.V.; Kalinkin, A.V. Charge Distribution in Layered Lanthanide-Doped CuCr0.99Ln0.01S2 (Ln = Pr–Tb) Thermoelectric Materials. Materials 2022, 15, 8747. [Google Scholar] [CrossRef]
  37. Mori, T.; Maignan, A. Thermoelectric materials developments: Past, present, and future. Sci. Technol. Adv. Mater. 2021, 22, 998–999. [Google Scholar] [CrossRef]
  38. Tsujii, N.; Kitazawa, H. Substitution effect on the two-dimensional triangular-lattice system CuCrS2. J. Phys. Condens. Matter 2007, 19, 145245. [Google Scholar] [CrossRef]
  39. Korotaev, E.V.V.; Syrokvashin, M.M.M.; Filatova, I.Y.; Zvereva, V.V. Vanadium doped layered copper-chromium sulfides: The correlation between the magnetic properties and XES data. Vacuum 2020, 179, 109390. [Google Scholar] [CrossRef]
  40. Hollander, J.C.T.; Sawatzky, G.; Haas, C. Monovalent copper in the chalcogenide spinel CuCr2Se4. Solid. State Commun. 1974, 15, 747–751. [Google Scholar] [CrossRef]
  41. Mazalov, L.N.; Sokolov, V.V.; Kryuchkova, N.A.; Vovk, E.I.; Filatova, I.Y.; Abramova, G.M. X-ray photoelectron spectroscopic studies of the charged state of 3d metal ions in CuCr1−XvXs2 (X = 0–0.4). J. Struct. Chem. 2009, 50, 439–445. [Google Scholar] [CrossRef]
  42. Korotaev, E.V.; Syrokvashin, M.M.; Sulyaeva, V.S.; Filatova, I.Y. Magnetic Properties of CuCr1−xLaxS2 Thermoelectric Materials. Magnetochemistry 2023, 9, 168. [Google Scholar] [CrossRef]
  43. CasaXPS Software. Available online: http://www.casaxps.com/ (accessed on 14 November 2022).
  44. Inorganic Crystal Structure Database, Version 2.1.0; Leibniz Institute for Information Infrastructure, FIZ Karlsruhe: Eggenstein-Leopoldshafen, Germany. Available online: https://icsd.products.fiz-karlsruhe.de/ (accessed on 17 August 2021).
  45. Shalimova, K.V. Semiconductors Physics; Lan: St. Petersburg, Russia, 2021; ISBN 978-5-8114-0922-8. [Google Scholar]
  46. Kabongo, G.L.L.; Mhlongo, G.H.H.; Mothudi, B.M.M.; Hillie, K.T.T.; Mbule, P.S.S.; Dhlamini, M.S.S. Structural, photoluminescence and XPS properties of Tm 3+ ions in ZnO nanostructures. J. Lumin. 2017, 187, 141–153. [Google Scholar] [CrossRef]
  47. de Groot, F.; Kotani, A. Core Level Spectroscopy of Solids; CRC Press: Boca Raton, FL, USA, 2008; ISBN 9780429195792. [Google Scholar]
  48. Larsson, S. Satellites in ESCA inner-shell spectra of 3d0 transition metal complexes. J. Electron. Spectros. Relat. Phenom. 1976, 8, 171–178. [Google Scholar] [CrossRef]
  49. Vasil’eva, I.G.; Kriventsov, V.V. Structural Study of CuCr1-XVxS2 Substitutional Solid Solutions. J. Surf. Investig. 2010, 4, 640–644. [Google Scholar] [CrossRef]
  50. NIST X-ray Photoelectron Spectroscopy (XPS) Database, Version 3.5. Available online: https://srdata.nist.gov/xps/ (accessed on 14 November 2022).
  51. XPS, AES, UPS and ESCA Database. Available online: http://www.lasurface.com/database/elementxps.php (accessed on 14 November 2022).
  52. Starodub, V.A. Ternary and Quaternary Chalcogenides of Group IB Elements. Usp. Khim. 1999, 68, 902–903. [Google Scholar] [CrossRef]
  53. Mikhlin, Y.L.L.; Tomashevich, Y.V.V.; Pashkov, G.L.L.; Okotrub, A.V.V.; Asanov, I.P.P.; Mazalov, L.N.N. Electronic structure of the non-equilibrium iron-deficient layer of hexagonal pyrrhotite. Appl. Surf. Sci. 1998, 125, 73–84. [Google Scholar] [CrossRef]
  54. Mikhlin, Y.; Karacharov, A.; Vorobyev, S.; Romanchenko, A.; Likhatski, M.; Antsiferova, S.; Markosyan, S. Towards Understanding the Role of Surface Gas Nanostructures: Effect of Temperature Difference Pretreatment on Wetting and Flotation of Sulfide Minerals and Pb-Zn Ore. Nanomaterials 2020, 10, 1362. [Google Scholar] [CrossRef] [PubMed]
  55. Selwood, P. Magnetochemistry, 2nd ed.; Interscience Publishers: New York, NY, USA, 1956. [Google Scholar]
  56. Blundell, S. Magnetism in Condensed Matter; OXFORD University Press: Oxford, UK, 2001. [Google Scholar] [CrossRef]
  57. Le Nagard, N.; Collin, G.; Gorochov, O. Etude structurale et proprietes physiques de CuCrS2. Mater. Res. Bull. 1979, 14, 1411–1417. [Google Scholar] [CrossRef]
  58. Abramova, G.M.; Petrakovskǐ, G.A.; Vorotynov, A.M.; Velikanov, D.A.; Kiselev, N.I.; Bovina, A.F.; Szymczak, R.; Al’mukhametov, R.F. Phase transitions and colossal magnetoresistance in CuVxCr 1-x S2 layered disulfides. JETP Lett. 2006, 83, 118–121. [Google Scholar] [CrossRef]
Figure 1. XRD patterns for CuCr1-xLaxS2 powder samples. *—most intense peak of CuLaS2 impurity phase.
Figure 1. XRD patterns for CuCr1-xLaxS2 powder samples. *—most intense peak of CuLaS2 impurity phase.
Jcs 07 00436 g001
Figure 2. XPS lines for CuCr1-xLnxS2 and CuLaS2: (a) Cu2p and (b) S2p region.
Figure 2. XPS lines for CuCr1-xLnxS2 and CuLaS2: (a) Cu2p and (b) S2p region.
Jcs 07 00436 g002
Figure 3. XPS lines for CuCr1-xLnxS2 and CuLaS2: (a) Cr2p- and (b) La3d-region.
Figure 3. XPS lines for CuCr1-xLnxS2 and CuLaS2: (a) Cr2p- and (b) La3d-region.
Jcs 07 00436 g003
Figure 4. Temperature dependencies of magnetic susceptibility (a,d), inverse magnetic susceptibility (b), and effective magnetic moment (c,d) for CuCr1-xLaxS2 (ac) solid solutions and CuLaS2 (d). *—data after correction procedure.
Figure 4. Temperature dependencies of magnetic susceptibility (a,d), inverse magnetic susceptibility (b), and effective magnetic moment (c,d) for CuCr1-xLaxS2 (ac) solid solutions and CuLaS2 (d). *—data after correction procedure.
Jcs 07 00436 g004
Figure 5. Minimum temperature of effective magnetic moment for CuCr1-xLaxS2 solid solutions.
Figure 5. Minimum temperature of effective magnetic moment for CuCr1-xLaxS2 solid solutions.
Jcs 07 00436 g005
Figure 6. Concentration dependencies of effective magnetic moment (a), Weiss constants (b), and total magnetic exchange interaction absolute value (c) for CuCr1−xLaxS2 solid solutions.
Figure 6. Concentration dependencies of effective magnetic moment (a), Weiss constants (b), and total magnetic exchange interaction absolute value (c) for CuCr1−xLaxS2 solid solutions.
Jcs 07 00436 g006
Figure 7. Density of states distribution (a), theoretical valence band spectra (b), and experimental XPS valence band spectra (c) of CuCr1-xLaxS2 solid solutions.
Figure 7. Density of states distribution (a), theoretical valence band spectra (b), and experimental XPS valence band spectra (c) of CuCr1-xLaxS2 solid solutions.
Jcs 07 00436 g007
Figure 8. Seebeck coefficient temperature dependence (a) and carrier concentration at room temperature (b) for CuCr1−xLaxS2 solid solutions.
Figure 8. Seebeck coefficient temperature dependence (a) and carrier concentration at room temperature (b) for CuCr1−xLaxS2 solid solutions.
Jcs 07 00436 g008
Table 1. The lattice parameters calculated for CuCr1-xLaxS2.
Table 1. The lattice parameters calculated for CuCr1-xLaxS2.
a, Åc, Å
CuCrS23.4807(9)18.701(6)
CuCr0.995La0.005S23.4816(8)18.707(6)
CuCr0.99La0.01S23.4820(8)18.708(6)
CuCr0.985La0.015S23.4807(9)18.702(6)
CuCr0.97La0.03S23.478(1)18.689(6)
Table 2. BE values for Cu2p-, Cr2p-, S2p-, and La3d-lines.
Table 2. BE values for Cu2p-, Cr2p-, S2p-, and La3d-lines.
Cu2p3/2Cr2p3/2S2p3/2La3d5/2
CuCrS2931.6574.2
577.2
160.6
162.5
CuCr0.995La0.005S2931.6574.2
576.6
160.7
162.6
834.9
CuCr0.99La0.01S2931.8574.3
576.8
160.9
162.7
834.7
CuCr0.985La0.015S2931.6574.2
576.7
160.7
162.8
834.8
CuCr0.97La0.03S2932.0574.5
577.1
161.0
162.7
834.7
CuLaS2931.9160.4
162.1
835.2
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

Korotaev, E.V.; Syrokvashin, M.M.; Filatova, I.Y. Thermoelectric and Magnetic Properties and Electronic Structure of Solid Solutions CuCr1-xLaxS2. J. Compos. Sci. 2023, 7, 436. https://doi.org/10.3390/jcs7100436

AMA Style

Korotaev EV, Syrokvashin MM, Filatova IY. Thermoelectric and Magnetic Properties and Electronic Structure of Solid Solutions CuCr1-xLaxS2. Journal of Composites Science. 2023; 7(10):436. https://doi.org/10.3390/jcs7100436

Chicago/Turabian Style

Korotaev, Evgeniy V., Mikhail M. Syrokvashin, and Irina Yu. Filatova. 2023. "Thermoelectric and Magnetic Properties and Electronic Structure of Solid Solutions CuCr1-xLaxS2" Journal of Composites Science 7, no. 10: 436. https://doi.org/10.3390/jcs7100436

Article Metrics

Back to TopTop