Next Article in Journal
Drug Delivery Systems with a “Tumor-Triggered” Targeting or Intracellular Drug Release Property Based on DePEGylation
Next Article in Special Issue
Investigation of Nitrogen and Vacancy Defects in Synthetic Diamond Plates by Positron Annihilation Spectroscopy
Previous Article in Journal
A Multimethod Assessment of a New Customized Heat-Treated Nickel–Titanium Rotary File System
Previous Article in Special Issue
Microstructural Gradational Properties of Sn-Doped Gallium Oxide Heteroepitaxial Layers Grown Using Mist Chemical Vapor Deposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Zn1−xPbxCr2Se4—Single Crystals Obtained by Chemical Vapour Transport—Structure and Magnetic, Electrical, and Thermal Properties

1
Institute of Chemistry, University of Silesia in Katowice, 40-007 Katowice, Poland
2
Institute of Physics, University of Silesia in Katowice, 40-007 Katowice, Poland
3
Institute of Materials Science, University of Silesia in Katowice, 40-007 Katowice, Poland
4
Faculty of Natural Sciences, Comenius University, 842 15 Bratislava, Slovakia
5
Katowice Branch, Research Network Łukasiewicz—Institute of Mechanised Construction and Rock Mining, 02-673 Warszawa, Poland
*
Author to whom correspondence should be addressed.
Materials 2022, 15(15), 5289; https://doi.org/10.3390/ma15155289
Submission received: 5 June 2022 / Revised: 20 July 2022 / Accepted: 27 July 2022 / Published: 31 July 2022
(This article belongs to the Special Issue Growth and Properties of Crystal Materials)

Abstract

:
Monocrystalline chalcogenide spinels ZnCr2Se4 are antiferromagnetic and semiconductor materials. They can be used to dope or alloy with related semiconducting spinels. Therefore, their Pb-doped display is expected to have unique properties and new potential applications. This paper presents the results of dc and ac magnetic measurements, including the critical fields visible on the magnetisation isotherms, electrical conductivity, and specific heat of the ZnCr2S4:Pb single crystals. These studies showed that substituting the diamagnetic Pb ion with a large ion radius for the Zn one leads to strong short-range ferromagnetic interactions in the entire temperature range and spin fluctuations in the paramagnetic region at Hdc = 50 kOe.

1. Introduction

Magnetic and semiconductor spinels are of great interest because of their exotic phenomena and the fascinating ground states observed at the turn of the 20th and 21st centuries, such as [1]: behaviour of heavy fermions [2,3], complex spin order and spin dimerization [4,5,6], spin-orbital fluid [7,8], orbital glass [9], the coexistence of ferromagnetism and ferroelectricity [10,11,12], the so-called Jahn–Teller effect [13,14,15], and phenomena attributed to the competing exchange of ferromagnetism (FM) and anti-ferromagnetism (AFM) of almost equal strength, causing a strong bond frustration [16].
Many interesting properties were observed in the spinel series that fulfil Vegard’s law in the full-concentration region. In the Katowice group at the end of the 20th century, the Zn1−xCuxCr2Se4 spinel series was particularly intensively studied, the edge spinels of which ZnCr2Se4 and CuCr2Se4 were respectively antiferromagnet and ferromagnet, and at the same time, a semiconductor and a conductor [17,18,19]. In the early 2000s, Parker et al. [20] discovered a transition from positive to negative magnetoresistance there.
Single crystals are used in modern, innovative technology and the electronic industry. The semiconducting [21] and antiferromagnetic [22] ZnCr2Se4 spinel is a matrix for various diluted systems where the effects of the site disorder, lattice frustration, and random distribution of spin interactions [23] create novel potential applications in spin-based electronic technology. In ZnCr2Se4, the spin order occurring at the Néel temperature TN = 21 K [1] is an incommensurate spiral due to the narrow distances between Cr ions. Still, the larger ones cause ferromagnetic (FM) interactions confirmed by the positive Curie–Weiss temperature of θ = 90 K [21,22]. For this reason, differences in magnetic interactions between the nearest and farther neighbours in ZnCr2Se4 are related to the frustration of chemical bonds that play an essential role. The magnetic order is accompanied by structural transformation from cubic F d 3 ¯ m to tetragonal I41/amd symmetry with a slight contraction along the c axis [24,25], manifested by sharp first-order anomalies at TN on the specific heat curve [1].
Our group’s work has been done based on the ZnCr2Se4 matrix. This matrix has been doped with both non-magnetic ions, such as, for example, Al [26,27,28], Cu [17,18,19,29,30], Ga [17,18,31,32,33], In [18,33,34], Sn [35,36], and Ta [37], and with magnetic ones, such as, for example, Dy [38,39], Gd [40], Ho [41], Mn [42,43], Nd [44], and Ni [45]. All the above-mentioned doped spinels have two features in common. They are semiconductors and antiferromagnets, just like a matrix. An interesting feature is the appearance of the re-entrant-type spin-glass when non-magnetic Al [27] and Sn [36] ions are located in the octahedral sites of the spinel structure. It was also observed that Sn ions occupying octahedral sites reduce the n-type electrical conductivity by magnitude compared to samples in which they occupy tetrahedral ones [35]. This interesting phenomenon has been explained based on the quantum band model, which predicts a lowering of the Fermi level in the lowest Cr3+ Mott–Hubbard sub-band with a narrowed 3d3 t2g band, which gives an increase in the energy gap when Sn3+ ions (spin defects) are replaced at octahedral sites instead of Cr3+ ones [35]. Interestingly, geometric spin-glass was observed when substituting Dy ions, a paramagnetic rare-earth ion in both tetrahedral and octahedral positions [39].
This work continues our research that involved introducing the p-block element into the crystal lattice ZnCr2Se4. Pb belongs to group 14 of the periodic table. Sn also belongs to this group. ZnCr2Se4 single crystals doped with tin showed the phenomenon of spin-glass [36]. It was assumed that the presence of Pb ions in the ZnCr2Se4 crystal lattice would affect the physicochemical properties of the parent compound. A single crystal growth model was developed using a computer programme. The structural, magnetic, electrical, and thermal properties were determined using various methods (XRD—X-ray Diffraction, SEM—Scanning Electron Microscopy, DSC—Differential Scanning Calorimetry, TG—Thermogravimetry, SQUID—Superconducting Quantum Interference Device, QD-PPMS—Quantum Design Physical Property Measurement System). This work presents the structural, magnetic, electrical, thermal, and specific heat properties of five monocrystalline samples doped with diamagnetic Pb2+ ions with a large ion radius instead of Zn2+ ones. The novelty of this work is the study of the ac magnetic susceptibility measured in an external dc magnetic field up to 50 kOe. According to our knowledge, such crystals have not been described in the literature.

2. Materials and Methods

High-purity elements, Zn, Pb, Cr, Se (5N, Sigma Aldrich, Sofia, Bulgaria), and anhydrous CrCl3 (Sigma Aldrich), were applied to synthesise ZnCr2Se4 single crystals doped with Pb.

2.1. Computation of Thermodynamic Parameters for the Chemical Transport Reactions

The basis for the growth of single crystals by gaseous chemical transport is reversible transport reactions. The criterion determining the reaction’s transport capacity is the order of magnitude of the equilibrium constant Kp. The Kp value should be close to 1 (logKp ≈ 0). The following dependence is used to assess the transport capacity of the reaction:
Δ G T 0 = Δ H 298 K 0 T Δ S 298 K 0
where Δ H 298 K   0 and Δ S 298 K 0 are the corresponding changes of the reaction’s standard and enthalpy and entropy. Free energy variations can be used to assess the transportability of the reaction Δ G 0 instead of the equilibrium constant Kp because these quantities are related to the formula:
Δ G 0 = 2.303 R T l o g K p = Δ H 0 T Δ S 0 ,
where: R—gas constant; T—absolute temperature; Kp—equilibrium constant; Δ G 0 , Δ H 0 , Δ S 0 —changes in the reaction’s free energy, enthalpy, and entropy. Accompanying reactions related to the solvent or reaction products’ dissociation may occur during the transport process. Therefore, all independent reactions occurring in the tested system should be considered. The selection of an appropriate transport substance involves finding such transport reactions for which the values of Kp or Δ G 0 , in the selected temperature range, do not deviate too much from the equilibrium values. We are looking for a substance that will produce sizeable partial pressure differences Δpi, with the slightest possible temperature difference.
In the process of selenium chromite single crystal growth, anhydrous CrCl3 is used, which dissociates at a temperature above 773 K according to the following reactions [46,47]:
2CrCl3(g) = 2CrCl2(g) + Cl2(g),
CrCl2(g) + Cl2(g) = CrCl4(g).
In a closed system, in a gas solution, apart from CrCl3 and Cl2, there is also CrCl4. Therefore, the dissociation products should be taken into account when creating a set of all independent transport reactions that occur simultaneously in the ZnSe–PbSe–CrCl3 system:
  • 2ZnSe + 4CrCl3(g) = 2ZnCl2(g) + 4CrCl2(g) + Se2(g)
  • 2ZnSe + 4CrCl3(g) = Zn2Cl4(g) + 4CrCl2(g) + Se2(g)
  • 2PbSe + 2CrCl3(g) = 2PbCl(g) + 2CrCl2(g) + Se2(g)
  • 2PbSe + 4CrCl3(g) = 2PbCl2(g) + 4CrCl2(g) + Se2(g)
  • 2ZnSe + 2CrCl4(g) = 2ZnCl2(g) + 2CrCl2(g) + Se2(g)
  • 2ZnSe + 2CrCl4(g) = Zn2Cl4(g) + 2CrCl2(g) + Se2(g)
  • 2PbSe + CrCl4(g) = 2PbCl(g) + CrCl2(g) + Se2(g)
  • 2PbSe + 2CrCl4(g) = 2PbCl2(g) + 2CrCl2(g) + Se2(g)
  • 2ZnSe + 2Cl2(g) = 2ZnCl2(g) + Se2(g)
  • 2ZnSe + 2Cl2(g) = Zn2Cl4(g) + Se2(g)
  • 2PbSe + Cl2(g) = 2PbCl(g) + Se2(g)
  • 2PbSe + 2Cl2(g) = 2PbCl2(g) + Se2(g)
These hypothetical reactions were used to calculate the temperature dependence of the thermodynamic parameters (ΔH0, ΔG0, ΔS0, and logKp). The calculations were performed using the H.S.C. Chemistry 6 computer programme. Based on these calculations, the reaction conditions (dissolution zone temperatures and crystallisation and their difference (Table 1)) make it possible to obtain good-quality crystals.

2.2. Synthesis of ZnSe and PbSe

The polycrystalline ZnSe and PbSe were obtained by the high-temperature sintering of powders (1073 K) using a solid-state reaction in a high vacuum (10−5 mbarr).

2.3. Crystal Structure and Chemical Composition

The X-ray diffraction measurement was carried out at 293(1) K. Five samples with different nominal Pb content were selected under a stereoscopic Zeiss optical microscope. A small crystal (diameter of about 0.1 mm) was mounted on a glass capillary. The data were collected using a SuperNova X-ray diffractometer with a microfocus X-ray tube, optimised multi-layer optics for Mo-Kα (λ = 0.71073 Å) radiation, and an Atlas CCD detector. Accurate cell parameters were determined and refined with CrysAlisPro software (Version 1.171.37.35, Agilent Technologies, Wrocław, Poland, 2014). Moreover, the CrysAlisPro program was used to integrate the collected data. The spinel structure (space group no. 227) was refined using the SHELXL-2013 program [48]. All atoms were refined with anisotropic displacement parameters. The chemical composition of the single crystals was studied using a Scanning Electron Microscope JEM 6480 equipped with an energy-dispersive X-ray spectrometer (SEM/EDS). The solid elements (Zn, Pb, Cr, Se) were used as standards to determine the mass % content of the elements in the sample. Measurements were carried out at 20 locations of the single crystal to determine the average chemical composition. Each of the measuring areas was approximately 50 × 30 μm. Then, the average chemical composition was calculated, which is shown in Table 1. The error bar represents the standard deviation. Relatively low values of the standard deviation indicate good homogeneity of the chemical composition.

2.4. Magnetic and Electrical Measurements

Dynamic ac magnetic susceptibility, χac, was measured in the temperature range 2–300 K and at an internal oscillating magnetic field Hac = 1 Oe, with an internal frequency f = 120 Hz taken at external static dc magnetic fields Hdc = 0, 10, 20, 30, 40, and 50 kOe. Magnetisation isotherms, M(H), were measured at 2, 10, 20, 40, 60, and 300 K and in the static magnetic field up to 70 kOe. A Quantum Design MPMS-XL-7AC SQUID magnetometer (Quantum Design, San Diego, CA, USA) was used. The Néel (TN) and Tm temperatures and the critical fields were defined as the extremes corresponding to the derivative of dχac/dT vs. T and dM/dH vs. H. The effective magnetic moment was calculated using the equation [37,42,49]: μ e f f = 2.828 C , where C is the Curie constant. The magnetic super-exchange integrals for the first two coordination spheres, J1 and J2, were calculated using the Holland and Brown equations [50]: J 1 = 9 T N + θ / 60 and J 2 = 3 T N + θ / 120 . The electrical conductivity, σ(T), was measured by the dc method using a KEITHLEY 6517B Electrometer/High Resistance Meter (Keithley Instruments, L.L.C., Solon, OH, USA) in the temperature range of 77–400 K. The activation energy, Ea, was determined below room temperature and in the temperature range of 300–400 K from the formula σ = σ 0 exp ( E a k T ) , where k is the Boltzmann constant and σ0 is the reference conductivity. The details of methods used to determine the electrical properties, thermal analysis, and specific heat are described in [37,39,41,44].

3. Results and Discussion

3.1. Growth of Single Crystals and Chemical Composition

Thermodynamic calculations indicate that, for the ZnSe–PbSe–CrCl3 system, the simultaneous transport of ZnSe and PbSe will be realised using gaseous CrCl3 and CrCl4 (logKp and ΔG0 values are close to zero in the selected temperature range: 1000–1400 K). It is shown in Figure 1 and Figure 2 that, for chemical reactions with CrCl3 and CrCl4, the enthalpy values are positive (ΔH0 > 0), which indicates that the transport takes place from a higher temperature to a lower temperature in the direction of the formation of products.
On the other hand, the enthalpy values for the transport of reactions with chlorine (Cl2) are negative (ΔH0 < 0), which indicates that the reaction equilibrium is shifted towards the reactants. Based on these calculations, it has been confirmed that in the ZnSe–PbSe–CrCl3 system, ZnSe and PbSe are mainly transported by gaseous CrCl3 and CrCl4. Values of logKa are close to zero in the selected temperature range: 1000–1400 K (Figure 3).
Synthesis of single crystals of Zn1–xPbxCr2Se4 was carried out according to the reaction:
4(1 − x)ZnSe + 4xPbSe + 2CrCl3 = Zn1−xPbxCrxSe4 + 3(1 − x)ZnCl2 + 3xPbCl2
for x = 0.1, 0.2, 0.3.
Thermodynamic calculations determined the reaction conditions, such as dissolution and crystallisation temperatures, and their difference. The crystallisation zone temperature was between 1083 and 1153 K. The dissolution zone was between 1133 and 1233 K. Their difference was 50–70 K. Based on our experience with the growth of doped ZnCr2Se4 single crystals, we can state that a minor temperature difference (ΔT) is more favourable for crystal growth. Stoichiometric mixtures of the ZnSe, PbSe, and CrCl3 were introduced to quartz ampoules (length—200 mm, inner diameter—20 mm) evacuated to 10−5 mbarr. The sealed ampoules were placed in a two-zone tubular furnace and heated for 450 h. After heating, the ampoules were cooled at about 70 degrees per hour. Such a preparation process for the growth of single crystals allowed for obtaining single crystals of good quality (Figure 4). Five single crystals with various Pb content were chosen for X-ray diffraction measurements. The reaction conditions and results of SEM analysis, together with determined chemical formulae, are shown in Table 1.

3.2. Structural Study

Based on the structure of ZnCr2Se4, in which the Zn2+ ions occupy the tetrahedral position 8a: 1/8, 1/8, 1/8 (A-site), and the Cr3+ ions occupy the position 16d: ½, ½, ½ (B-site), according to the spinel structure, we considered two models of cation distribution. In the first model, the presence of Pb2+ ions in A-sites was assumed with coupled site occupancy factors (SOF) and constrained atomic displacements. The second model assumed that the Pb2+ ions would substitute the Cr3+ ones. Because of the strong correlation, the SOFs for Cr and Pb were refined separately in alternative calculations. The assumed oxidation state of Pb and the correlations of the refined parameters caused the rejection of the second model.
Based on this refinement, the first model allowed us to obtain the acceptable atomic displacement parameters and SOFs. We could describe the general chemical formula for obtained single crystals as presented in Table 2 and Table 3. The observed slight increase in thermal shift at the A-site points out a static disturbance in the location of the Zn2+/Pb2+ pseudo-ion, which is usually caused by the difference in ionic charges and in ionic and covalent radii (ri (Zn2+) = 0.60 Å, ri (Cr3+) = 0.62 Å, ri (Pb2+) = 0.98 Å and Rc (Zn2+) = 0.74 Å, Rc (Cr3+) = 0.76 Å, Rc (Pb2+) = 1.12 Å, respectively) [51].
The structural study confirmed the chemical composition obtained from SEM. The general formula of obtained single crystals can be described as Zn1−xPbxCr2Se4. The Pb ions substituted the Zn ones. The single crystals crystallised in a cubic system with space group Fd-3m (No. 227, Z = 8). The lattice parameters of the obtained single crystals are slightly larger than pure ZnCr2Se4 (a = 10.489 Å) and enhanced with the increase in lead content (Table 2). The increase in lattice parameters is connected with the difference in ionic radii between Zn2+ and Pb2+ (Figure 5).
The positional parameter of Se (u) is a measure of the anion sublattice distortion from the cubic-close packing. The value of u is the same for all obtained single crystals. Its value increases slightly compared to the ideal value of u = 0.250 (Table 3). The admixture of lead does not significantly affect the u-value. The same is observed in the metal–metal and metal–selenium distances (Table 4), where the differences are insignificant [22]. The structure refinement parameters, the selected bond distances, and angles are shown in Table 3 and Table 4.

3.3. Electrical and Magnetic Properties

The results of the electrical measurements are shown in Figure 6 and Table 5, and the magnetic ones are in Figure 7, Figure 8 and Figure 9 and Table 5 and Table 6. Electrical conductivity measurements (Figure 6) showed semiconducting properties in the temperature range of 77–400 K for all single crystals under study. Two regions of Arrhenius type electrical conductivity are visible: the extrinsic one with the activation energy of Ea ~0.08 eV below room temperature, independent of the Pb content in the sample, and the intrinsic one with Ea increasing in the range of 0.138–0.315 eV above room temperature with increasing Pb content (Table 5).
The increase in activation energy with the increase in lead content may be caused by the appearance of cationic vacancies (acting as double acceptors) because the radius of Pb2+ ions is much larger than that of Zn2+ ones [51]. It may result in acceptor vacancy levels in the energy gap. The activation energies mentioned above correlate well with Ea for the matrix [31] and with its admixtures, such as tin [35], tantalum [37], and dysprosium [22].
The temperature dependencies of ac magnetic susceptibility χac, recorded in an internal oscillating magnetic field Hac = 1 Oe with an internal frequency f = 120 Hz and with a zero external static magnetic field (Figure 7a–e), show AFM order with the Néel temperature TN = 22 K and the positive Curie–Weiss temperature (θ) slightly changing from 74 K to 84 K as the Pb content increases (Table 5).
The θ values are much lower than the literature data published in [21,22] and slightly lower than in [1], where θ is equal to 115 K and 90 K, respectively. Thus, only the TN temperature values are close to this temperature’s values for the ZnCr2Se4 matrix [1,21,22]. It means that the long-range AFM interactions are slightly more substantial. In contrast, the short-range FM ones are somewhat weaker in the sample containing Pb, which is confirmed by both the long bond lengths (Table 4) and the values of the J1 and J2 super-exchange integrals for the first two coordination spheres (Table 5). A significant difference was observed between the effective magnetic moment (μeff) and the effective number of Bohr magnetons, peff = 5.477, i.e., the spin contribution to the magnetic moment for Cr3+ ions per molecule 3d3 configuration. It may suggest that the orbital magnetic contribution has not been quenched at low amplitude, Hac = 1 Oe, of the applied oscillating magnetic field.
The magnetisation measurements (Figure 8a–e) show that the magnetic saturation of the studied samples is close to the value of 6 μB/f.u. for the matrix; the first critical field is kept constant at Hc1 = 12 kOe, and the second critical field is around Hc2 = 57 kOe (Table 5). The hysteresis loops have zero-field coercivity and zero remanences (Figure 8f).
The temperature dependencies of ac magnetic susceptibility χac, recorded in an internal oscillating magnetic field Hac = 1 Oe with an internal frequency f = 120 Hz and taken at external static magnetic fields Hdc = 0, 10, 20, 30, 40, and 50 kOe, are shown in Figure 9a–e. With the increase in Hdc, we observed a shift of TN towards lower temperatures and of θ towards higher ones (Table 6).
A strong magnetic field weakens the AFM order and strengthens the FM one. Moreover, the Curie constant, C, and the effective moment μeff, are close to the values typical for a chromium ion per molecule. For Hdc = 50 kOe, the J1 super-exchange integral for the first coordination sphere changes the sign from negative to positive, while the J2 integral remains positive (Table 6). It means that the short-range FM interaction extends over the entire temperature range. The χac(T) curves above TN in the paramagnetic region show characteristic wide maxima at Tm = 32–36 and 44–48 K in the fields Hdc = 40 and 50 kOe, respectively. These broad maxima of ac magnetic susceptibility may be due to the spin fluctuations that occur due to the amplification of short-range FM interactions by a static magnetic field, which is opposed by the thermal energy kT. Similar maxima were found in the ZnCr2Se4 matrix doped with Al [28], Ce, Ga, and In [33].

3.4. Specific Heat Studies

In Figure 10, the points represent the specific heat measured for samples with the formulae: Zn0.92Pb0.07Cr2Se4, Zn0.92Pb0.09Cr2Se4, Zn0.89Pb0.11Cr2Se4, and Zn0.88Pb0.12Cr2Se4.
We see that the Debye model fits the experimental data well for all the samples, with the Debye temperatures obtained from the fits 163–177 K, with no apparent dependence of Debye temperature on the Pb concentration. The number of atoms obtained from the fit lies within 7.2–9.5 for the investigated samples, without clear dependence on the Pb concentration. It can be explained by the fact that the change in sample composition is minute, and the changes in Zn concentration accompany the changes in Pb concentration. The upper inset of Figure 10 shows the specific heat in a narrow region around the magnetic ordering temperature. The sample Zn0.92Pb0.09Cr2Se4 exhibits quite a sharp peak in the vicinity of magnetic transition, whereas the remaining compositions show a broader two-peak-like structure. This is most likely caused by the disorder induced by the Pb substitution and the lack of full site occupancy of the Zn site. The bottom inset of Figure 10 shows the shift of the magnetic transition with the increase in magnetic field for the Zn0.92Pb0.07Cr2Se4 sample, which is a clear manifestation of antiferromagnetic order.
We also measured the resistivity for the Zn0.88Pb0.12Cr2Se4 sample (the sample with the highest Pb concentration) with the resistivity option in the PPMS instrument (Figure 11). Below 120 K, the resistivity exceeded the range of the instrument. However, in the 120–300 K range, the resistivity can be well-fitted with an activation law ρ = ρ0exp(Δ/kT) (ρ0—resistivity at 273.15 K), where for an indirect bandgap, we obtained Δ ~0.88 eV. The inset shows the resistivity measured vs. magnetic field at room temperature.
The magnetoresistivity is positive and almost linear with the magnetic field. It is also relatively small, below 2% for 9 T.

3.5. Thermal Analysis

The thermal analysis examined two crystals of Zn1−xPbxCr2Se4 containing minimum (0.06) and maximum (0.12) lead amounts. The DSC and TG curves are depicted in Figure 12.
The shape of the DSC curves is similar to pure ZnCr2Se4. For the sample Zn0.94Pb0.06Cr2Se4, two small endothermic peaks are observed at 711 °C and 733 °C. The observed peaks are in a similar position close to pure ZnCr2Se4 (755 °C) [44]. These peaks indicate that the melting process of the sample starts at about 711 °C. The peak confirms it on the DTG curve. However, the sample Zn0.88Pb0.12Cr2Se4 lacks typical endothermic peaks. On the TG curves for both samples, a few percent increase in weight is visible in the first step. This phenomenon can indicate that the investigated sample absorbs an inert gas during the heating or solid-state–gas reaction. The mass loss is observed with increasing temperature on the DTG curve. It is an endothermic reaction, as indicated by the downward deflection of the DSC curve. The partial destruction of investigated samples takes place at about 1300 °C. It suggests that the samples doped with Pb are thermally more stable than pure ZnCr2Se4.

4. Conclusions

In conclusion, we have demonstrated the family of Zn1−xPbxCr2Se4 (x = 0.06, 0.07, 0.09, 0.11, 0.12) spinel single crystals showing semiconducting and AFM behaviour. These single crystals were successfully obtained using chemical vapour transport. The technology of crystal growth was based on thermodynamic calculations. XRD analysis showed that these crystals possess a cubic structure (SG: Fd3m). Thermal analysis showed that the single crystals under study are thermally stable up to 1300 °C.
With the increase in the external magnetic field, a shift of TN and the specific heat peak towards lower temperatures and θ towards higher ones was observed, as well as a strong weakening of long-range AFM interactions visible in the reduction of the super-exchange integral for the first coordination sphere and the appearance of spin fluctuations in the paramagnetic region, visible in magnetic fields of 40 and 50 kOe. Below TN, the magnetic field dependence of magnetisation, M(H), showed two peculiarities at critical fields Hc1 = 12 and Hc2 = 57 kOe and a change of the sign of the J1-integral from negative to positive at Hdc = 50 kOe, which causes the short-range FM interaction to extend over the entire temperature range. Finally, we can conclude that doped chalcogenide spinels are suitable materials for application in many technological areas, such as magnetic materials for capacitors, transformers, or components in electronic products.

Author Contributions

Conceptualisation, I.J. and T.G. (Tadeusz Groń); methodology, I.J., T.G. (Tadeusz Groń), B.S., E.P., J.J., B.W.-K. and H.D.; validation, I.J. and T.G. (Tadeusz Groń); formal analysis, I.J., B.W.-K. and T.G. (Tadeusz Groń); investigation, I.J., Z.S., J.K., E.P., T.G. (Tomasz Goryczka), B.S., J.J., J.G. and H.D.; resources, I.J.; writing—original draft preparation, I.J., T.G. (Tadeusz Groń), and J.G.; writing—review and editing, I.J., T.G. (Tadeusz Groń) and J.G.; visualisation, I.J. and T.G. (Tadeusz Groń); supervision, I.J.; project administration, I.J. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by project VEGA 1/0116/22.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hemberger, J.; von Nidda, H.-A.K.; Tsurkan, V.; Loidl, A. Large Magnetostriction and Negative Thermal Expansion in the Frustrated Antiferromagnet ZnCr2Se4. Phys. Rev. Lett. 2007, 98, 147203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kondo, S.; Johnston, D.C.; Swenson, C.A.; Borsa, F.; Mahajan, A.V.; Miller, L.L.; Gu, T.; Goldman, A.I.; Maple, M.B.; Gajewski, D.A.; et al. LiV2O4: A Heavy Fermion Transition Metal Oxide. Phys. Rev. Lett. 1997, 78, 3729–3732. [Google Scholar] [CrossRef] [Green Version]
  3. Krimmel, A.; Loidl, A.; Klemm, M.; Horn, S.; Schober, H. Dramatic Change of the Magnetic Response in LiV2O4: Possible Heavy Fermion to Itinerant d-Metal Transition. Phys. Rev. Lett. 1999, 82, 2919–2922. [Google Scholar] [CrossRef]
  4. Lee, S.-H.; Broholm, C.; Ratcliff, W.; Gasparovic, G.; Huang, Q.; Kim, T.H.; Cheong, S.-W. Emergent excitations in a geometrically frustrated magnet. Nature 2002, 418, 856–858. [Google Scholar] [CrossRef] [PubMed]
  5. Radaelli, P.G.; Horibe, Y.; Gutmann, M.J.; Ishibashi, H.; Chen, C.H.; Ibberson, R.M.; Koyama, Y.; Hor, Y.-S.; Kiryukhin, V.; Cheong, S.-W. Formation of isomorphic Ir3+ and Ir4+ octamers and spin dimerisation in the spinel CuIr2S4. Nature 2002, 416, 155–158. [Google Scholar] [CrossRef] [PubMed]
  6. Schmidt, M.; Ratcliff, W.; Radaelli, P.G., II; Refson, K.; Harrison, N.M.; Cheong, S.W. Spin Singlet Formation in MgTi2O4: Evidence of a Helical Dimerization Pattern. Phys. Rev. Lett. 2004, 92, 056402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Fritsch, V.; Hemberger, J.; Büttgen, N.; Scheidt, E.-W.; Krug von Nidda, H.-A.; Loidl, A.; Tsurkan, V. Spin and Orbital Frustration in MnSc2S4 and FeSc2S4. Phys. Rev. Lett. 2004, 92, 116401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Krimmel, A.; Mücksch, M.; Tsurkan, V.; Koza, M.M.; Mutka, H.; Loidl, A. Vibronic and Magnetic Excitations in the Spin-Orbital Liquid State of FeSc2S4. Phys. Rev. Lett. 2005, 94, 237402. [Google Scholar] [CrossRef]
  9. Fichtl, R.; Tsurkan, V.; Lunkenheimer, P.; Hemberger, J.; Fritsch, V.; Krug von Nidda, H.-A.; Scheidt, E.-W.; Loidl, A. Orbital Freezing and Orbital Glass State in FeCr2S4. Phys. Rev. Lett. 2005, 94, 027601. [Google Scholar] [CrossRef] [Green Version]
  10. Hemberger, J.; Lunkenheimer, P.; Fichtl, R.; Krug von Nidda, H.-A.; Tsurkan, V.; Loidl, A. Relaxor ferroelectricity and colossal magnetocapacitive coupling in ferromagnetic CdCr2S4. Nature 2005, 434, 364–367. [Google Scholar] [CrossRef] [Green Version]
  11. Weber, S.; Lunkenheimer, P.; Fichtl, R.; Hemberger, J.; Tsurkan, V.; Loidl, A. Colossal Magnetocapacitance and Colossal Magnetoresistance in HgCr2S4. Phys. Rev. Lett. 2006, 96, 157202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Yamasaki, Y.; Miyasaka, S.; Kaneko, Y.; He, J.-P.; Arima, T.; Tokura, Y. Magnetic Reversal of the Ferroelectric Polarisation in a Multiferroic Spinel Oxide. Phys. Rev. Lett. 2006, 96, 207204. [Google Scholar] [CrossRef] [PubMed]
  13. Lee, S.-H.; Broholm, C.; Kim, T.H.; Ratcliff, W.; Cheong, S.-W., II. Local Spin Resonance and Spin-Peierls-like Phase Transition in a Geometrically Frustrated Antiferromagnet. Phys. Rev. Lett. 2000, 84, 3718. [Google Scholar] [CrossRef] [Green Version]
  14. Tchernyshyov, O.; Moessner, R.; Sondhi, S.L. Order by Distortion and String Modes in Pyrochlore Antiferromagnets. Phys. Rev. Lett. 2002, 88, 067203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Yamashita, Y.; Ueda, K. Spin-Driven Jahn-Teller Distortion in a Pyrochlore System. Phys. Rev. Lett. 2000, 85, 4960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Hemberger, J.; Rudolf, T.; Krug von Nidda, H.-A.; Mayr, F.; Pimenov, A.; Tsurkan, V.; Loidl, A. Spin-Driven Phonon Splitting in Bond-Frustrated ZnCr2S4. Phys. Rev. Lett. 2006, 97, 087204. [Google Scholar] [CrossRef] [Green Version]
  17. Groń, T.; Duda, H.; Warczewski, J. Influence of the valence of the substituted cations on the electrical properties of the magnetically modulated spinels Zn1-xAxCr2Se4 (A = Cu, Ga2/3). Phys. Rev. 1990, 41, 12424–12431. [Google Scholar] [CrossRef]
  18. Groń, T. Influence of vacancies and mixed valence on the transport processes in solid solutions with the spinel structure. Philos. Mag. B 1994, 70, 121–132. [Google Scholar] [CrossRef]
  19. Groń, T.; Wolff, J.; Hehenkamp, T.; Bärner, K.; Okońska-Kozłowska, I.; Jendrzejewska, I.; Malicka, E. Positron annihilation studies in single and polycrystals of Zn1-xCuxCr2Se4 spinel series. Radiat. Eff. Defects Solids 1996, 139, 97–107. [Google Scholar] [CrossRef]
  20. Parker, D.R.; Green, M.A.; Bramwell, S.T.; Wills, A.S.; Gardner, J.S.; Neumann, D.A. Crossover from Positive to Negative Magnetoresistance in a Spinel. J. Am. Chem. Soc. 2004, 126, 2710–2711. [Google Scholar] [CrossRef]
  21. Lotgering, F.K. Ferromagnetic interactions in sulphides, selenides and tellurides with spinel structure. In Proceedings of the International Conference on Magnetism, Nottingham, UK, 7–11 September 1964; Institute of Physics and the Physical Society London: London, UK, 1965; p. 533. [Google Scholar]
  22. Plumier, R. Étude par diffraction de neutrons de l’antiferromagnétisme hélicoïdal du spinelle ZnCr2Se4 en présence d’un champ magnétique. J. Phys. 1966, 27, 213–219. [Google Scholar] [CrossRef]
  23. Park, S.; Kwon, S.; Lee, S.; Khim, S.; Bhoi, D.; Park, C.B.; Kim, K.H. Interactions in the bond-frustrated helimagnet ZnCr2Se4 investigated by NMR. Sci. Rep. 2019, 9, 16627. [Google Scholar] [CrossRef] [PubMed]
  24. Kleinberger, R.; de Kouchkovsky, C.R. Étude radiocrystallographique à basse temperature du spinelle ZnCr2Se4. Acad. Sci. Paris Ser. B 1966, 262, 628–630. [Google Scholar]
  25. Yokaichiya, F.; Krimmel, A.; Tsurkan, V.; Margiolaki, I.; Thompson, P.; Bordallo, H.N.; Buchsteiner, A.; Stußer, N.; Argyriou, D.N.; Loidl, A. Spin-driven phase transitions in ZnCr2Se4 and ZnCr2S4 probed by high-resolution synchrotron X-ray and neutron powder diffraction. Phys. Rev. B 2009, 79, 064423. [Google Scholar] [CrossRef] [Green Version]
  26. Malicka, E.; Groń, T.; Pacyna, A.W.; Maciążek, E.; Duda, H.; Pawełczyk, M.; Zawisza, B.; Sitko, R. Influence of temperature on the critical fields in ZnCr2-xAlxSe4 antiferromagnets. J. Alloys Compd. 2009, 480, 67–69. [Google Scholar] [CrossRef]
  27. Groń, T.; Malicka, E.; Duda, H.; Pacyna, A.W.; Mydlarz, T.; Sitko, R.; Pawełczyk, M. Spin-glass-like behavior in ZnxCryAlzSe4. J. Phys. Chem. Sol. 2009, 70, 1175–1180. [Google Scholar] [CrossRef]
  28. Malicka, E.; Groń, T.; Ślebarski, A.; Pacyna, A.W.; Goraus, J.; Fijałkowski, M.; Heimann, J. Specific heat and magnetic susceptibility of single-crystalline ZnCr2-xAlxSe4 (x = 0.15, 0.23). J. Phys. Chem. Sol. 2011, 72, 974–979. [Google Scholar] [CrossRef] [Green Version]
  29. Groń, T.; Jendrzejewska, I.; Gołąbek, S.; Duda, H.; Krajewski, A.; Bärner, K. The thermoelectric power of ferromagnetically ordered ZnxCuyCrzSe4 single crystals. Phys. B 2003, 327, 88–95. [Google Scholar] [CrossRef]
  30. Mazur, S.; Groń, T.; Jendrzejewska, I. Influence of spin arrangement on thermopower in ZnxCuyCrzSe4 spinels. J. Alloys Compd. 2009, 480, 19–22. [Google Scholar] [CrossRef]
  31. Okońska-Kozłowska, I.; Lutz, H.D.; Groń, T.; Krok, J.; Mydlarz, T. Darstellung, elektrische und magnetische Eingenschaften von-Zn1-xGa0.667xCr2Se4-Spinell-Einkristallen. Mater. Res. Bull. 1984, 19, 1–5. [Google Scholar] [CrossRef]
  32. Groń, T.; Duda, H.; Warczewski, J. Transport phenomena in the antiferromagnetic spinels Zn1-xGa2x/3Cr2Se4 (where 0.0 < x < 0.5). J. Mag. Mag. Mater. 1990, 83, 487–489. [Google Scholar]
  33. Malicka, E.; Groń, T.; Ślebarski, A.; Gągor, A.; Pacyna, A.W.; Sitko, R.; Goraus, J.; Mydlarz, T.; Heimann, J. Specific heat and magnetic susceptibility of single-crystalline ZnCr2Se4 spinels doped with Ga, In and Ce. Mater. Chem. Phys. 2011, 131, 142–150. [Google Scholar] [CrossRef]
  34. Groń, T.; Kopyczok, J.; Okońska-Kozłowska, I.; Warczewski, J. Seebeck effect in the antiferromagnetic single crystals of ZnCr2-xInxSe4 (0.0 < x < 0.15). J. Mag. Mag. Mater. 1992, 111, 53–55. [Google Scholar]
  35. Jendrzejewska, I.; Groń, T.; Kusz, J.; Żelechower, M.; Maciążek, E.; Ślebarski, A.; Fijałkowski, M. Spin-glass-like behaviour in tin doped ZnCr2Se4 single crystals. J. Alloys Compd. 2015, 635, 238–244. [Google Scholar] [CrossRef]
  36. Jendrzejewska, I.; Mroziński, J.; Groń, T.; Duda, H.; Zajdel, P.; Pacyna, A.W.; Maciążek, E.; Hanc, A. Effect of cation substitution on Fermi level of n-type ZnxSnyCrzSe4 spinels. J. Alloys Compd. 2009, 480, 63–66. [Google Scholar] [CrossRef]
  37. Jendrzejewska, I.; Groń, T.; Kwapuliński, P.; Kusz, J.; Pietrasik, E.; Goryczka, T.; Sawicki, B.; Ślebarski, A.; Fijałkowski, M.; Jampilek, J.; et al. Study of the structure, magnetic, thermal and electrical characterisation of ZnCr2Se4:Ta single crystals obtained by chemical vapour transport. Materials 2021, 14, 2749. [Google Scholar] [CrossRef]
  38. Jendrzejewska, I.; Groń, T.; Knizek, K.; Pilch, M.; Ślebarski, A.; Goraus, J.; Zajdel, P.; Stokłosa, Z.; Pietrasik, E.; Goryczka, T.; et al. Preparation, structure and magnetic, electronic and thermal properties of Dy3+-doped ZnCr2Se4 with unique geometric type spin-glass. J. Sol. State Chem. 2021, 298, 122114. [Google Scholar] [CrossRef]
  39. Jendrzejewska, I.; Groń, T.; Maciążek, E.; Duda, H.; Kubisztal, M.; Ślebarski, A.; Pietrasik, E.; Fijałkowski, M. Specific heat and magnetic properties of single-crystalline ZnxDyyCrzSe4 spinels. J. Mag. Mag. Mater. 2016, 407, 122–128. [Google Scholar] [CrossRef]
  40. Maciążek, E.; Karolus, M.; Kubisztal, M.; Jendrzejewska, I.; Sitko, R.; Groń, T.; Ślebarski, A.; Fijałkowski, M. Magnetic and specific heat properties of a new Gd-doped ZnCr2Se4. Mater. Chem. Phys. 2015, 168, 187–192. [Google Scholar] [CrossRef]
  41. Jendrzejewska, I.; Groń, T.; Kusz, J.; Goraus, J.; Barsova, Z.; Pietrasik, E.; Czerniewski, J.; Goryczka, T.; Kubisztal, M. Growth, structure and physico-chemical properties of monocrystalline ZnCr2Se4:Ho prepared by chemical vapour transport. J. Sol. State Chem. 2020, 281, 121024. [Google Scholar] [CrossRef]
  42. Jendrzejewska, I.; Groń, T.; Goraus, J.; Pilch, M.; Pietrasik, E.; Barsova, Z.; Czerniewski, J.; Goryczka, T.; Witkowska-Kita, B.; Bienko, A.; et al. Synthesis and structural, magnetic, thermal and electronic properties of Mn-doped ZnCr2Se4. Mater. Chem. Phys. 2019, 238, 121901. [Google Scholar] [CrossRef]
  43. Jendrzejewska, I.; Zajdel, P.; Heimann, J.; Krok-Kowalski, J.; Mydlarz, T.; Mrzigod, J. Influence of manganese on magnetic and electronic properties of ZnCr2Se4 single Crystals. Mater. Res. Bull. 2012, 47, 1881–1886. [Google Scholar] [CrossRef]
  44. Jendrzejewska, I.; Groń, T.; Kusz, J.; Barsova, Z.; Pietrasik, E.; Goryczka, T.; Sawicki, B.; Ślebarski, A.; Fijałkowski, M.; Jampilek, J.; et al. Synthesis, crystal structure and characterization of monocrystalline ZnCr2Se4 doped with neodymium. J. Sol. State Chem. 2020, 292, 121661. [Google Scholar] [CrossRef]
  45. Jendrzejewska, I.; Waśkowska, A.; Mydlarz, T. Influence of nickel substitution on the cation distribution and magnetic properties of ZnCr2Se4. J. Alloys Compd. 2001, 327, 73–77. [Google Scholar] [CrossRef]
  46. Lenz, M.; Gruehn, R. Developments in Measuring and Calculating Chemical Vapor Transport Phenomena Demonstrated on Cr, Mo, W and Their Compounds. Chem. Rev. 1997, 97, 2967–2994. [Google Scholar] [CrossRef]
  47. Plies, V. Massenspektrometrische Untersuchungen der Gasphase über CrCl3 und CrCl3/Cl2. Z. Anorg. Allg. Chem. 1988, 556, 120–128. [Google Scholar] [CrossRef]
  48. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
  49. Morrish, A.H. Physical Principles of Magnetism; John Wiley & Sons, Inc.: New York, NY, USA, 1965. [Google Scholar]
  50. Holland, W.E.; Brown, H.A. Application of the Weiss molecular field theory to the B-site spinel. Phys. Status Solidi 1972, 10, 249–253. [Google Scholar] [CrossRef]
  51. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
Figure 1. The dependence of the thermodynamic parameters vs. temperature T for the ZnSe and PbSe with CrCl3 as a transport agent.
Figure 1. The dependence of the thermodynamic parameters vs. temperature T for the ZnSe and PbSe with CrCl3 as a transport agent.
Materials 15 05289 g001
Figure 2. The dependence of the thermodynamic parameters vs. temperature T for the ZnSe and PbSe with CrCl4 as a transport agent.
Figure 2. The dependence of the thermodynamic parameters vs. temperature T for the ZnSe and PbSe with CrCl4 as a transport agent.
Materials 15 05289 g002
Figure 3. The dependence of the thermodynamic parameters vs. temperature T for the ZnSe and PbSe with Cl2 as a transport agent.
Figure 3. The dependence of the thermodynamic parameters vs. temperature T for the ZnSe and PbSe with Cl2 as a transport agent.
Materials 15 05289 g003
Figure 4. (a) A part of quartz ampoule with obtained ZnCr2Se4:Pb single crystals. Examples of single crystals obtained in the ZnCr2Se4:Pb system: (b) Zn0.94Pb0.06Cr2Se4, (c) Zn0.92Pb0.09Cr2Se4, (d) Zn0.88Pb0.12Cr2Se4.
Figure 4. (a) A part of quartz ampoule with obtained ZnCr2Se4:Pb single crystals. Examples of single crystals obtained in the ZnCr2Se4:Pb system: (b) Zn0.94Pb0.06Cr2Se4, (c) Zn0.92Pb0.09Cr2Se4, (d) Zn0.88Pb0.12Cr2Se4.
Materials 15 05289 g004
Figure 5. Dependence of the lattice parameters of Zn1-xPbxCr2Se4 single crystals on the Pb amount.
Figure 5. Dependence of the lattice parameters of Zn1-xPbxCr2Se4 single crystals on the Pb amount.
Materials 15 05289 g005
Figure 6. Electrical conductivity (lnσ) vs. reciprocal temperature (103/T) for ZnCr2Se4 single crystals doped with Pb2+ ions having contents of 0.06, 0.07, 0.09, 0.11, and 0.12.
Figure 6. Electrical conductivity (lnσ) vs. reciprocal temperature (103/T) for ZnCr2Se4 single crystals doped with Pb2+ ions having contents of 0.06, 0.07, 0.09, 0.11, and 0.12.
Materials 15 05289 g006
Figure 7. Ac magnetic susceptibility χac and its reciprocal 1/χac vs. temperature T recorded at internal oscillating magnetic field Hac = 1 Oe with internal frequency f = 120 Hz for the ZnCr2Se4 single crystals doped with Pb2+ ions having contents of 0.06 (a), 0.07 (b), 0.09 (c), 0.11 (d), and 0.12 (e). The solid (black) line, (T–θ)/C, indicates a Curie–Weiss behaviour.
Figure 7. Ac magnetic susceptibility χac and its reciprocal 1/χac vs. temperature T recorded at internal oscillating magnetic field Hac = 1 Oe with internal frequency f = 120 Hz for the ZnCr2Se4 single crystals doped with Pb2+ ions having contents of 0.06 (a), 0.07 (b), 0.09 (c), 0.11 (d), and 0.12 (e). The solid (black) line, (T–θ)/C, indicates a Curie–Weiss behaviour.
Materials 15 05289 g007aMaterials 15 05289 g007b
Figure 8. Magnetization M vs. magnetic field H at 2, 4, 10, 20, 40, 60, and 300 K for the ZnCr2Se4 single crystals doped with Pb2+ ions having contents of 0.06 (a), 0.07 (b), 0.09 (c), 0.11 (d), and 0.12 (e). The M(H) isotherms recorded at 2 K with arrows denoting the first (Hc1) and second (Hc2) critical magnetic fields are shown in figure (f).
Figure 8. Magnetization M vs. magnetic field H at 2, 4, 10, 20, 40, 60, and 300 K for the ZnCr2Se4 single crystals doped with Pb2+ ions having contents of 0.06 (a), 0.07 (b), 0.09 (c), 0.11 (d), and 0.12 (e). The M(H) isotherms recorded at 2 K with arrows denoting the first (Hc1) and second (Hc2) critical magnetic fields are shown in figure (f).
Materials 15 05289 g008
Figure 9. Ac magnetic susceptibility χac vs. temperature T recorded at internal oscillating magnetic field Hac = 1 Oe with internal frequency f = 120 Hz for the ZnCr2Se4 single crystals doped with Pb2+ ions having contents of 0.06 (a), 0.07 (b), 0.09 (c), 0.11 (d), and 0.12 (e) taken at external magnetic fields Hdc = 0, 10, 20, 30, 40, and 50 kOe. Vertical arrows indicate the Néel TN and Tm temperatures, and horizontal ones indicate the shift of TN and Tm with increasing magnetic field Hdc.
Figure 9. Ac magnetic susceptibility χac vs. temperature T recorded at internal oscillating magnetic field Hac = 1 Oe with internal frequency f = 120 Hz for the ZnCr2Se4 single crystals doped with Pb2+ ions having contents of 0.06 (a), 0.07 (b), 0.09 (c), 0.11 (d), and 0.12 (e) taken at external magnetic fields Hdc = 0, 10, 20, 30, 40, and 50 kOe. Vertical arrows indicate the Néel TN and Tm temperatures, and horizontal ones indicate the shift of TN and Tm with increasing magnetic field Hdc.
Materials 15 05289 g009
Figure 10. Specific heat, C, as a function of temperature T, measured for the Zn1−xPbxCr2Se4 single crystals (x = 0.07, 0.09, 0.11, and 0.12) at zero magnetic field (main figure). The dashed lines show the Debye model fit experimental data for T > 40 K. The upper inset shows how the magnetic peak is affected by Pb substitution. The lower inset shows how the magnetic transition temperature is affected by the magnetic field for the Zn0.92Pb0.07Cr2Se4 sample. Note that the temperature scales on the lower and upper inset are not the same. Hence, the peaks in the upper inset appear much broader than in the lower inset or the main figure.
Figure 10. Specific heat, C, as a function of temperature T, measured for the Zn1−xPbxCr2Se4 single crystals (x = 0.07, 0.09, 0.11, and 0.12) at zero magnetic field (main figure). The dashed lines show the Debye model fit experimental data for T > 40 K. The upper inset shows how the magnetic peak is affected by Pb substitution. The lower inset shows how the magnetic transition temperature is affected by the magnetic field for the Zn0.92Pb0.07Cr2Se4 sample. Note that the temperature scales on the lower and upper inset are not the same. Hence, the peaks in the upper inset appear much broader than in the lower inset or the main figure.
Materials 15 05289 g010
Figure 11. The resistivity of the Zn0.88Pb0.12Cr2Se4 single crystal measured vs. temperature (central Figure) and vs. field at T = 300 K (inset). The Zn0.88Pb0.12Cr2Se4 sample was tiny. Hence, we do not provide the absolute value of the resistivity.
Figure 11. The resistivity of the Zn0.88Pb0.12Cr2Se4 single crystal measured vs. temperature (central Figure) and vs. field at T = 300 K (inset). The Zn0.88Pb0.12Cr2Se4 sample was tiny. Hence, we do not provide the absolute value of the resistivity.
Materials 15 05289 g011
Figure 12. DSC/TG curves for Zn0.94Pb0.06Cr2Se4 (a) and Zn0.88Pb0.12Cr2Se4 (b) single crystals.
Figure 12. DSC/TG curves for Zn0.94Pb0.06Cr2Se4 (a) and Zn0.88Pb0.12Cr2Se4 (b) single crystals.
Materials 15 05289 g012
Table 1. Conditions of growth and chemical composition of ZnCr2Se4:Pb single crystals.
Table 1. Conditions of growth and chemical composition of ZnCr2Se4:Pb single crystals.
NoAmounts of
Substrates
(mmol)
Td (K)Tc
(K)
ΔT
(K)
% WeightChemical Formula
nZnSenPbSeZnPbCrSe
(1)14.41.6122311636012.31 ± 0.093.23 ± 0.0620.66 ± 0.0163.71 ± 0.04Zn0.94Pb0.06Cr2Se4
(2)14.41.6113310835012.25 ± 0.082.95 ± 0.0720.45 ± 0.0264.54 ± 0.05Zn0.92Pb0.07Cr2Se4
(3)12.83.2120311535012.05 ± 0.063.43 ± 0.0620.49 ± 0.0564.10 ± 0.03Zn0.92Pb0.09Cr2Se4
(4)12.83.2120311436011.63 ± 0.024.51 ± 0.0420.30 ± 0.0663.56 ± 0.03Zn0.89Pb0.11Cr2Se4
(5)11.24.8120311535011.42 ± 0.034.98 ± 0.0520.15 ± 0.0363.42 ± 0.03Zn0.88Pb0.12Cr2Se4
Chemical compositions were determined using Scanning Electron Microscopy. Td is the temperature of the dissolution zone, Tc is the temperature of the crystallisation zone, and ΔT is the difference in temperatures between the dissolution and the crystallisation zones.
Table 2. Structural parameters of the Zn1-xPbxCr2Se4 single crystals.
Table 2. Structural parameters of the Zn1-xPbxCr2Se4 single crystals.
Chemical FormulaLattice Parameter (Å)Volume
3)
Density Calc.
(Mg/m3)
Absorption Coeff.
(mm−1)
Goodness of Fit on F2R Parameters
R1wR2
Zn0.94Pb0.06Cr2Se410.4910(3)1154.65(2)5.68232.8371.1500.02230.0662
Zn0.92Pb0.07Cr2Se410.4933(2)1155.41(3)5.68932.8831.2050.03330.1114
Zn0.92Pb0.09Cr2Se410.4946(4)1155.84(1)5.70932.8561.1750.02610.0839
Zn0.89Pb0.11Cr2Se410.4954(2)1156.10(2)5.75733.8411.1960.03330.1114
Zn0.88Pb0.12Cr2Se410.4959(1)1156.27(3)5.77233.8521.1910.03110.1012
Table 3. Atomic coordinates and equivalent isotropic displacement parameters of the Zn1−xPbxCr2Se4 single crystals. The Wyckoff positions of the atoms in the spinel structure are: Zn in 8 b (3/8, 3/8, 3/8); Cr/Ta in 16 c (½, ½, ½); and Se in 32 e (x, x, x). U (eq) is defined as one-third of the orthogonalized Uij tensor trace.
Table 3. Atomic coordinates and equivalent isotropic displacement parameters of the Zn1−xPbxCr2Se4 single crystals. The Wyckoff positions of the atoms in the spinel structure are: Zn in 8 b (3/8, 3/8, 3/8); Cr/Ta in 16 c (½, ½, ½); and Se in 32 e (x, x, x). U (eq) is defined as one-third of the orthogonalized Uij tensor trace.
SpinelAnion
Parameter u
Site OccupationUiso (Å × 103)
(A)(B)Zn/PbCrSe
(Zn0.94Pb0.06)[Cr2]Se40.2595(1)0.94(1):0.06(1)1.013(1)7(1)7(1)
(Zn0.92Pb0.07)[Cr2]Se40.2595(1)0.92(2):0.07(1)1.014(1)6(1)6(1)
(Zn0.92Pb0.09)[Cr2]Se40.2595(1)0.92(1):0.09(1)1.013(1)7(1)6(1)
(Zn0.89Pb0.11)[Cr2]Se40.2595(1)0.89(2):0.11(2)1.014(1)5(1)5(1)
(Zn0.88Pb0.12)[Cr2]Se40.2595(1)0.87(1):0.13(2)1.014(1)5(1)6(1)
Table 4. Selected interatomic distances (Å) and bond angles (deg) of the Zn1-xPbxCr2Se4 spinel single crystals.
Table 4. Selected interatomic distances (Å) and bond angles (deg) of the Zn1-xPbxCr2Se4 spinel single crystals.
SpinelBond DistancesBond Angles
Zn/Pb–SeCr–SeSe–Zn/Pb–SeSe–Cr–Se
(Zn0.94Pb0.06)[Cr2]Se42.4451(5)2.5282(3)109.5(0) × 6180.0(0) × 3
94.61(13) × 6
85.39(13) × 6
(Zn0.92Pb0.07)[Cr2]Se42.4442(7)2.5269(4)109.5(0) × 6180.0(0) × 3
94.61(2) × 6
85.39(2) × 6
(Zn0.92Pb0.09)[Cr2]Se42.4453(5)2.5274(3)109.5(0) × 6180.0(0) × 3
94.62(14) × 6
85.38(14) × 6
(Zn0.89Pb0.11)[Cr2]Se42.4453(8)2.5279(4)109.5(0) × 6180.0(0) × 3
94.62(2) × 6
82.38(2) × 6
(Zn0.88Pb0.12)[Cr2]Se42.4453(7)2.5280(4)109.5(0) × 6180.0(0) × 3
94.61(2) × 6
85.39(2) × 6
Table 5. Magnetic parameters of Zn1−xPbxCr2Se4 single crystals recorded in an internal oscillating magnetic field Hac = 1 Oe with an internal frequency f = 120 Hz and with a zero external static magnetic field: C is the Curie constant; TN and θ are the Néel and Curie–Weiss temperatures, respectively; µeff is the effective magnetic moment; J1 and J2 are the super-exchange integrals for the first two coordination spheres. Hc1 and Hc2 are the critical fields measured at the static magnetic field up to 70 kOe. Ea is the energy activation measured in the temperature range of 300–400 K. Experimental data for the ZnCr2Se4 matrix were taken from Refs. [1,21,22,31] for comparison.
Table 5. Magnetic parameters of Zn1−xPbxCr2Se4 single crystals recorded in an internal oscillating magnetic field Hac = 1 Oe with an internal frequency f = 120 Hz and with a zero external static magnetic field: C is the Curie constant; TN and θ are the Néel and Curie–Weiss temperatures, respectively; µeff is the effective magnetic moment; J1 and J2 are the super-exchange integrals for the first two coordination spheres. Hc1 and Hc2 are the critical fields measured at the static magnetic field up to 70 kOe. Ea is the energy activation measured in the temperature range of 300–400 K. Experimental data for the ZnCr2Se4 matrix were taken from Refs. [1,21,22,31] for comparison.
SpinelCTNθµeffMs(2K)J1J2Hc1Hc2Ea
(emu·K/mol)(K)(K)B/f.u.)B/f.u.)(K)(K)(kOe)(kOe)(eV)
ZnCr2Se44.08221905.7146.0−1.651.2810.065.00.135
Zn0.94Pb0.06Cr2Se45.47722746.6185.325−2.071.1712.055.00.138
Zn0.92Pb0.07Cr2Se45.56822776.6735.964−1.951.2312.056.00.161
Zn0.92Pb0.09Cr2Se44.72722816.1495.960−2.021.1912.061.00.146
Zn0.89Pb0.11Cr2Se45.76822836.7925.486−1.921.2412.055.00.177
Zn0.88Pb0.12Cr2Se45.89822846.8685.251−1.901.2512.057.00.315
Table 6. Magnetic parameters of the Zn1−xPbxCr2Se4 single crystals recorded in an internal oscillating magnetic field Hac = 1 Oe with an internal frequency f = 120 Hz and taken at external static magnetic fields Hdc = 0, 10, 20, 30, 40, and 50 kOe: C is the Curie constant; TN, θ, and Tm are the Néel, Curie–Weiss, and the spin fluctuation temperatures, respectively; µeff is the effective magnetic moment; and J1 and J2 are the super-exchange integrals for the first two coordination spheres.
Table 6. Magnetic parameters of the Zn1−xPbxCr2Se4 single crystals recorded in an internal oscillating magnetic field Hac = 1 Oe with an internal frequency f = 120 Hz and taken at external static magnetic fields Hdc = 0, 10, 20, 30, 40, and 50 kOe: C is the Curie constant; TN, θ, and Tm are the Néel, Curie–Weiss, and the spin fluctuation temperatures, respectively; µeff is the effective magnetic moment; and J1 and J2 are the super-exchange integrals for the first two coordination spheres.
SpinelHdcCTNθTmµeffJ1J2
(kOe)(emu·K/mol)(K)(K)(K)B/f.u.)(K)(K)
Zn0.94Pb0.06Cr2Se405.4772274-6.618−2.071.17
104.2112276-5.803−1.731.13
204.0861879-5.716−1.381.11
303.9861682-5.646−1.031.08
403.948129531.55.6190.221.09
503.8636108465.5580.901.05
Zn0.92Pb0.07Cr2Se405.5682277-6.673−1.951.23
104.1822078-5.783−1.61.18
204.1021981-5.728−1.451.18
304.0031684-5.658−0.91.11
403.9881296335.648−0.181.11
503.976799445.6390.601.00
Zn0.92Pb0.09Cr2Se404.7272281-6.149−2.021.19
103.9222182-5.601−1.851.18
203.7112084-5.448−1.651.17
303.6811685-5.426−1.001.10
403.602129733.55.367−0.201.10
503.6396108465.3690.901.05
Zn0.89Pb0.11Cr2Se405.7682283-6.792−1.921.24
104.2932084-5.859−1.601.20
204.1821985-5.783−1.41.18
304.1011686-5.727−0.971.12
404.0561291345.695−0.281.06
504.002697485.6570.720.96
Zn0.88Pb0.12Cr2Se405.8982284-6.868−1.901.25
103.7942086-5.508−1.571.22
203.7811988-5.499−1.381.21
303.7621690-5.485−0.901.15
403.7031294365.442−0.231.08
503.689698485.4320.730.97
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jendrzejewska, I.; Groń, T.; Kusz, J.; Stokłosa, Z.; Pietrasik, E.; Goryczka, T.; Sawicki, B.; Goraus, J.; Jampilek, J.; Duda, H.; et al. The Zn1−xPbxCr2Se4—Single Crystals Obtained by Chemical Vapour Transport—Structure and Magnetic, Electrical, and Thermal Properties. Materials 2022, 15, 5289. https://doi.org/10.3390/ma15155289

AMA Style

Jendrzejewska I, Groń T, Kusz J, Stokłosa Z, Pietrasik E, Goryczka T, Sawicki B, Goraus J, Jampilek J, Duda H, et al. The Zn1−xPbxCr2Se4—Single Crystals Obtained by Chemical Vapour Transport—Structure and Magnetic, Electrical, and Thermal Properties. Materials. 2022; 15(15):5289. https://doi.org/10.3390/ma15155289

Chicago/Turabian Style

Jendrzejewska, Izabela, Tadeusz Groń, Joachim Kusz, Zbigniew Stokłosa, Ewa Pietrasik, Tomasz Goryczka, Bogdan Sawicki, Jerzy Goraus, Josef Jampilek, Henryk Duda, and et al. 2022. "The Zn1−xPbxCr2Se4—Single Crystals Obtained by Chemical Vapour Transport—Structure and Magnetic, Electrical, and Thermal Properties" Materials 15, no. 15: 5289. https://doi.org/10.3390/ma15155289

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