Next Article in Journal / Special Issue
Batteryless Electronic System Printed on Glass Substrate
Previous Article in Journal / Special Issue
Structural, Electronic, and Optical Properties of p-Type Semiconductors Cu2O and ZnRh2O4: A Self-Consistent Hybrid Functional Investigation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Method to Significantly Improve the Mechanical Properties of n-Type Bi(1−x)Sbx Thermoelectrics Due to Plastic Deformation

1
RusTec LTC, Peschany Carier 3, 109383 Moscow, Russia
2
Department of Materials Engineering, Ben-Gurion University of the Negev, Beer-Sheva 84105, Israel
*
Author to whom correspondence should be addressed.
Electron. Mater. 2021, 2(4), 511-526; https://doi.org/10.3390/electronicmat2040036
Submission received: 16 September 2021 / Revised: 2 October 2021 / Accepted: 18 October 2021 / Published: 2 November 2021
(This article belongs to the Special Issue Feature Papers of Electronic Materials)

Abstract

:
A unique method was developed to significantly improve the strength of Bi(1−x)Sbx single crystals, the most effective thermoelectric (TE) materials in the temperature range from 100 to 200 K due to their plastic deformation by extrusion. After plastic deformation at room temperature under all-round hydrostatic compression in a liquid medium, n-type Bi–Sb polycrystalline solid solutions show a significant increase in mechanical strength compared to Bi–Sb single crystals in the temperature range from 300 to 80 K. The significantly higher strength of extruded alloys in comparison with Bi–Sb single crystals is associated with the development of numerous grains with a high boundary surface as well as structural defects, such as dislocations, that accumulate at grain boundaries. Significant stability of the structure of extruded samples is achieved due to the uniformity of crystal plastic deformation under all-round hydrostatic compression and the formation of the polycrystalline structure consisting of grains with the orientation of the main crystallographic directions close to the original single crystal. The strengthening of Bi–Sb single crystals after plastic deformation allows for the first time to create workable TE devices that cannot be created on the basis of single crystals that have excellent TE properties, but low strength.

1. Introduction

A significant increase in the efficiency of thermoelectric (TE) cooling at temperatures below 200 K is currently one of the urgent problems of thermoelectricity. (Bi,Sb)2(Te,Se)3 solid solutions have remained the preferred thermoelectric materials for cooling purposes for many decades [1]. However, at cold temperatures of Tc ≤ 200 K, thermoelectric materials based on Bi2Te3 compounds are not effective for refrigeration due to low thermoelectric efficiency at these temperatures [1,2,3]. Their figure of merit Z for both n- and p-type legs does not exceed 2.5 × 10−3 K−1 at T = 200 K, restricting their practical application at cold temperatures of Tc ≤ 200 K [2,3,4,5].
An effective way to increase the efficiency of the TE device at cold temperatures is to maximize Z values. The most effective known thermoelectric (TE) materials at temperatures of T ≤ 180 K are single crystals of n-type Bi–Sb solid solutions with Sb content from 7 to 15 at.% [2,3,4,5]. Below this temperature, the Z value of these crystals is two to three times higher than that of the best n-type TE materials based on Bi2(Te,Se)3 solid solutions in a transverse magnetic field of B ≤ 1 T [3,5]. The thermoelectric properties of these TEs were first measured more than 100 years ago by G. Gehlhoff and F. Neumeier, as indicated by Smith and Wolfe [6], but in all of the early works on these materials, polycrystalline specimens of unknown purity were used. The ultimate bending strength σb of these Bi–Sb single crystals can reach 40 MPa at room temperature when the loading force is perpendicular to the trigonal axis [3,7,8,9]. To improve the bending strength of Czochralski-grown ingots, the effect of the growth parameters and the size and surface quality of the samples on σb has been investigated in detail by Belaya et al. [3,7]. At 77 K, the bending strength σb is relatively low (10–20 MPa); for example, when the bending force in a three-point test is orientated along the <10 1 ¯ > and <1 2 ¯ 1> axes, the σb values of Bi0.91Sb0.09 single crystals were about 18–28 MPa [8]. At T < 200 K, it is promising to use n-type branches from Bi–Sb single crystals together with superconducting branches from high-temperature superconductors instead of p-type branches [8,9].
However, single crystals of Bi–Sb solid solutions have not received practical application as a material for TE legs. This is mainly due to the ease of cleavage of single-crystal branches along the (111) plane perpendicular to the direction of the maximum values of Z [2,5,7,10]. In other words, Bi–Sb single crystals cannot withstand any bending along their trigonal axis, although it is in this direction that the figure of merit is the best. Therefore, the decisive condition for the successful use of TE-legs from Bi–Sb solid solutions is the development of methods for increasing the strength properties of these TE materials.
The processing (hot and cold pressing, extrusion, etc.) of polycrystalline TE materials may significantly affect the microstructure and improve their properties. A number of works indicate that extruded Bi–Sb solid solutions are promising as n-type TE material at temperatures of T ≤ 180 K [2,5,8,9,10]. Extrusion of Bi–Sb materials may introduce a preferential orientation of the grains. From a thermoelectric point of view, the presence of a strong texture in this material would be favorable. Actually, if all the grains could be orientated so that their trigonal axes are nearly parallel, then their thermoelectric properties should approach those of the single crystals [2,3,4]. Annealing the extruded Bi–Sb alloys restores high mobility of electrons but destroys the preferred grain orientation in the <111> direction along the extrusion axis. As a result of high deformation during extrusion and subsequent annealing of Bi–Sb alloys, a textured polycrystal structure is formed with a characteristic grain size of 0.02–0.05 mm and high-angle grain boundaries [5,8].
Mechanical alloying of the elemental materials as reported by Martin-Lopez et al. [10] is also a powerful method for the synthesis of homogeneous powders of the Bi(1−x)Sbx alloy at x = 0.07, 0.1, 0.12, 0.15, 0.22, with grain sizes about 10 μm, using relatively short milling times (4–15 h). The mechanical properties of the Bi0.85Sb0.15 alloy prepared by mechanical alloying and consolidated by hot extrusion were investigated at two fixed temperatures (77 and 293 K). The modulus of rupture of the polycrystalline material was roughly the same at these temperatures, with a value of about 100 MPa compared to 10 and 20 MPa for a Bi0.85Sb0.15 single crystal [10]. However, sintering methods offer the advantage of improving the mechanical strength of alloys but, unfortunately, degrade the thermoelectric performance as a result of random orientation. Moreover, impurities such as oxygen during power elaboration could be factors hindering improvement of the figure of merit [2].
However, the application of known extrusion methods [11] to Bi–Sb solid solutions does not make it possible to obtain a material possessing high strength and thermoelectric characteristics simultaneously. The effect of plastic deformation during extrusion on the structural, mechanical, and TE properties of Bi–Sb solid solutions has been little studied. This work is devoted to the development of a method for cold deformation of Bi–Sb single crystals at high hydrostatic pressure, which will improve the strength characteristics of the material without substantial decrease in the TE efficiency. A detailed study of the structure, texture, and mechanical properties of deformed Bi–Sb alloy was carried out in the temperature range from 300 to 80 K.

2. Materials and Methods

2.1. Growth of Single Crystals

The initial Bi0.91Sb0.09 single crystals were obtained using the Czochralski method [3,5,9,12,13,14], which is currently used in the production of more than half of the technically important crystals grown from the melt; at the same time, about 60% of all single crystals are used to obtain semiconductors [13. The single crystal was grown on a single crystal seed, and the melt was fed with solid antimony during the growth without additional alloying. The purity of the starting materials was 99.9999%. The single crystals 25–35 mm in diameter and 40–50 mm in length were grown in an atmosphere of high-purity helium at 140–150 kPa at the crystal pulling speed of 1.7 × 10−6 m s−1.

2.2. Extrusion of Single Crystals

To increase the strength of Bi0.91Sb0.09 single crystals, the method for their plastic deformation in a liquid medium at high hydrostatic pressure was used [5,8,9], for which the scheme and view of the extruder are shown in Figure 1. Here, castor oil used as a working fluid was first compressed to a pressure P, and the TE material was then extruded. Bi–Sb single-crystal samples for plastic deformation 12 mm in diameter and length of 25–30 mm were cut on an electric discharge machine and etched in concentrated HNO3 to remove the damaged layer of 40–50 μm depth. The cylindrical axis was oriented along the trigonal axis [111] of the single crystal.
For extrusion, blanks with a smooth surface without visible cracks were used. Quality control of the workpiece surfaces was carried out using an MBS-9 binocular microscope. The pressure of the working fluid P required for the extrusion was selected based on the microhardness HV of Bi0.91Sb0.09 single crystals, which was measured using a PMT-3 microhardness tester on the cleavage plane (111) at room temperature and ranged from 460 to 500 MPa. Considering the plastic deformation of a crystal with an internal crack placed in a medium with a high hydrostatic pressure P, a criterion P > HV was developed for preventing crystal cracking during plastic deformation, where HV is the microhardness of the crystalline material. To perform plastic deformation of Bi0.91Sb0.09 single crystals at room temperature (T = 300 K), the pressure P of the surrounding liquid was chosen as 600–1100 MPa.
The extrusion process in a hydrostat consists of two stages: preparatory and extrusion. In the preparatory phase, the working fluid is compressed as the plunger moves inside the pressure chamber, prior to extrusion of the TE material. The preparatory stage ends when the hydrostat plunger contacts the extruder plunger. Extrusion ratio, showing how many times the length increased or the cross-sectional area of the product decreased in one pass of extrusion, was 1.2, 3, 5, and 10.

2.3. Study of the Crystal Structure of Single Crystals and Extruded Specimens

The deformation of the material in the surface layer occurs under conditions that differ significantly from the conditions of deformation of the bulk of the crystal. Therefore, the crystal structure of the surface layer differs from the crystal structure of the bulk of the crystal. The 10–15 μm thick surface layer was removed by etching in 70% nitric acid; then, the sample was washed sequentially in concentrated hydrochloric acid, distilled water, and isopropyl alcohol and dried.
For structural study of the perfection of the grown crystals, a method with reflection recording was used. In this case, the cleavage plane (111) was investigated. For the topographic study, the Schultz method was used with a point source of white X-ray radiation [15,16]. The cleavage plane of the sample was located at an angle of 20–25° to the source. On radiographs, the surface protrusions are displayed as light stripes, and the depressions are dark. The width of these stripes was used to determine the angle of disorientation of the substructural components. Additionally, the divergent beam X-ray technique [17] has been used for precision measurement of the lattice period of perfect single crystals and for studying structural changes in deformed crystals. By changing the shape and width of the lines, it is possible to determine the angles of rotation of the substructural components of the crystal and the perfection of a small region with a length of 10–50 μm.
To determine the quantitative characteristics of the preferential structural orientation, we used a model in which the texture of a real crystal is considered as a combination of perfectly oriented substructures corresponding to the observable crystalline textures. In this model, a technique for magnetometric analysis was developed, which is almost independent of the state of grain boundaries, the presence of micropores at grain boundaries, as well as inevitable fluctuations in the concentration of impurities and composition [18]. This method is based on the magnetic properties of Bi–Sb crystals, which are anisotropic diamagnets at room temperature, with a weak dependence of the magnetic susceptibility on composition.
The rotating moment which acts on a textured diamagnetic sample of the mass m in a magnetic field can be determined through the free energy of the sample F as [18]
M ( ) = F / ( ) = M i w i f i /
where i is one of the textures, wi = mi/m is the specific content of the ith texture; fi is the specific (per unit mass) free energy of the ith texture in the magnetic field, and ∅ is the angle between the chosen direction in the sample, which is usually the one with the highest symmetry.
The magnetic properties of single crystals of diamagnetic Bi–Sb alloys are determined by two main values of the specific magnetic susceptibility tensor, namely k3 along the (111) trigonal axis and k1 perpendicular to the (111) direction. If the (111) axis in the ith texture is tilted at an angle ∅i to the direction of the uniform magnetic field H, then
f i =   1 2   H 2 [ k 3     ( k 1 k 3 ) s i n 2 i ] .
By varying the orientation of poly- and monocrystalline samples with respect to the magnetic field so that the rotating moments are maximal for the given field, we obtain the specific content of the prevailing texture
w 1 =   m 0 / m × ( H 0 / H ) 2 × M ¯ / M ¯ 0
where M ¯ and M ¯ 0 are the maximal rotating moments for a textured polycrystal in the field H and for a single crystal of the same composition in the field H0, respectively.
Since Bi–Sb alloys have a predominant <111> texture at a degree of deformation over 90%, the {111} axial texture prevails, and in this case, the {111} basal planes of the grains are oriented along the extrusion axis. Using this method, which does not require direct measurements of the magnetic susceptibility, the content of texture <111> (W<111>) was obtained by changing the orientation of the polycrystalline samples with respect to the magnetic field so that the torques were maximum for a given field.

2.4. Three-Point Bending Test of Bi0.91Sb0.09 Alloy

A schematic diagram of the rhombohedral and hexagonal lattice of Bi–Sb solid solutions is shown in Figure 2a. A hexagonal lattice system is conventionally used with a 1, a 2, a 3, and c lattice vectors. The angles between the a vectors equal 120°, and they lie in the plane normal to the c vector. A Cartesian coordinate system is also shown with axes denoted binary (OX), bisectrix (OY), and trigonal (OZ). The trigonal axis is aligned with the c direction in the lattices. The bisectrix axis is defined as being orthogonal to the trigonal and binary axes.
For n-type branches from Bi–Sb crystals, the most dangerous deformation is bending perpendicular to the direction of maximum TE efficiency, which coincides with the trigonal axis for single crystals or with the extrusion axis for crystals after plastic deformation. Therefore, from the point of view of their use as a TE material, the most important strength parameter for crystals of Bi–Sb solid solutions is the ultimate bending strength σb, when a bending force Fb is applied perpendicular to the trigonal or extrusion axis. Bending tests of samples by the three-point test method (Figure 2b) were carried out on an Instron 37-10-16 at temperatures of 77–300 K in a special chamber cooled with a mixture of gaseous and liquid nitrogen [9].
The surfaces of the samples undergoing bending were carefully mechanically ground and then chemically polished in a solution composed of 250 mL of H2O + 70 mL of HCl + 83 g of CrO3. The thickness of the removed layer was 5–10 μm. The average grain size in polycrystals of extruded samples was determined on the basis of the number N of grains observed in a given area S = 0.5 × 0.5 mm2, containing at least 100 grains. Then, the diameter d of a circle with an area equal to the area of one grain was determined. The value d was taken as the average grain size:
d = (4/π)0.5(S/N)0.5
In bending tests, Bi–Sb crystals of rectangular cross-section were used with dimensions h = (1 ± 0.1) mm, b = (4 ± 0.1) mm, l = 10 ± 0.1 mm, for which l / h ≈ 10, which justifies the application of the Equations (5) and (6) for calculation of the maximum tensile deformation εb and strength σb during bending [19]:
εb = 6·Δb·h/l2,
σb = l.5·Fb·l/(bh2),
where Δb is the maximum deflection of the specimen and Fb is the bending force.

3. Results

3.1. Crystal Structure of Bi0.91Sb0.09 Single and Polycrystalline Crystals

The dislocations within and at the boundaries of the blocks are responsible for the total rotation field of the crystal lattice. In Figure 3a, an X-ray diffraction pattern of Bi–Sb crystals is presented. The shape and width of the diffraction lines indicate the high perfection of the grown crystals. Figure 3b shows the images of diffraction spots from the crystal grown using the Schultz method. The images show images of substructural components with boundaries parallel and perpendicular to the growth axis. The size of the reflexes from individual substructural components and the width of the gaps between them were used to determine the sizes of substructural components and the angles of misorientation between them.
It was shown by the Schultz topographic method that, first, the blocks are predominantly elongated in the direction of crystal elongation; second, the transverse dimensions of the blocks are from 0.1 to 2.0 mm; third, the angles of rotation of the blocks relative to each other are in the range from several minutes to 1°. Note that a substructure in the form of weakly misoriented single-crystal blocks was also observed earlier in Bi–Sb crystals grown by the Czochralski method [5,9,12,13,14]. The formation of blocks is associated with the fact that the growth of Bi–Sb crystals is accompanied by the appearance of local concentration overcooling regions at the crystallization front. The results obtained using these methods indicate the presence of weakly misoriented blocks, relatively free of dislocations, with boundaries consisting of dislocation networks. The formation of blocks is due to local concentration of overcooling regions at the crystallization front. Blocks are regions that are relatively free of dislocations, and their boundaries consist of dislocation networks. The length and width of single-crystal blocks along the direction of crystal growth and perpendicular to it were 0.7–1.4 mm and 0.1–0.2 mm, respectively, at an angle of rotation of the blocks relative to each other of 3–5 arc min.
In Figure 4, the typical diffraction pattern of the sample after deformation is compared with the isotropic polycrystalline structure of the reference sample. This indicates the predominant orientation of the axes of trigonal crystallites in the direction of the extrusion axis, namely, the predominance of the <111> texture. As noted above, to quantify the degree of this texture, a magnetometric method was used in which the W<111> content in the direction of the extrusion axis was calculated using Equation (3). With an increase in the extrusion ratio from 1.2 to 10, the value of W<111> reduces from 88 to 50% (Table 1). The W<111> values were determined by averaging five identically prepared samples over the results of magnetometric measurements. Table 1 also shows the root-mean-square errors of the reduced values of W<111>. The values of pole density Phkl for extruded samples before and after additional annealing are presented in Table 2. Annealing was carried out at a temperature of 180 °C for 8 h in an inert gas atmosphere at a pressure of 150 kPa. In calculating Phkl, only the most intense diffraction lines were taken into account. According to the data presented in Table 2, the values of the pole density P003 with Miller indices in multiples of (001) significantly exceed Phkl of other observed diffraction lines. This indicates the predominant orientation of the trigonal axes of the crystallites along the extrusion axis or, in other words, the dominance of the <111> texture.
Annealing of Bi–Sb extruded crystals causes a further decrease in the W<111> values. As shown in Figure 5 and listed in Table 1, annealing of extruded samples causes intense recrystallization, leading to reorientation and coarsening of polycrystalline grains. In this case, the scattering of the <111> texture occurs more intensely for the samples obtained at higher values of the Ke (Figure 5).
Figure 6 shows examples of the structure of cleavages of Bi–Sb extruded samples at Ke = 1.2, 3, and 5. After annealing, on the cleavages perpendicular to the extrusion axis, there are triangular etching pits typical for the etched surface of cleavages along the (111) plane of single-crystal samples (Figure 6a). This indicates that during the extrusion process, the initial crystallographic orientation of the crystals is largely retained. Elongated grains coincide with the direction of extrusion (Figure 6b). At high magnification, the features of the cleavage of the sample extruded at Ke 5 show that the structure was practically isotropic in the plane perpendicular to the extrusion axis (Figure 6c). In Figure 6d, the scanning electron image of the fracture surface for the representative extruded Bi0.91Sb0.09 sample shows the excellent densification and uniform distribution of grains. No pores or cracks were identified on the surface of the sample, which is in agreement with the high density of the extruded samples, ~99% of the theoretical density. The estimated average grain size ranges from 20 to 50 μm, and the size of the largest grains does not exceed ~80 μm.
As a result of structural studies, it is shown that a strongly textured polycrystalline Bi–Sb at Ke ≤ 3 is formed with a predominant orientation of the crystallographic axes of grains with an average size varying from 20 to 50 μm, coinciding with the orientation of the corresponding crystallographic axes of the single crystal before extrusion. At the extrusion ratio Ke ≥ 5, the structure was practically isotropic in the plane perpendicular to the extrusion axis. For Ke = 10, as a result of extrusion, the axial texture < 11 1 ¯ > dominates in the direction of the extrusion axis, and the structure of the extruded polycrystal is practically independent of the orientation of a single crystal before extrusion.

3.2. Mechanical Properties of Bi0.91Sb0.09 Single Crystals and Polycrystalline Material

After homogenizing annealing at 453 K for 8 h, bending tests were performed on Bi0.91Sb0.09 single crystals at 77–300 K. The tested crystals were made so that the direction OZ (Figure 2b), coinciding with the trigonal axis of the crystal, was perpendicular to the direction of the bending force Fb. In a plane perpendicular to the trigonal axis, the bending force Fb was oriented along the bisectrix or binary axes (OY or OX), that is, Fb || <1 2 ¯ 1> or Fb || <10 1 ¯ >, respectively. Ultimate bending strength σb of Bi–Sb extruded samples was much higher than that of single crystals and increased significantly with an increase in the extrusion ratio (Figure 7) and insignificantly with a decrease in the test temperature (Table 3). After extrusion at Ke ≥ 3, the bending strength of the extruded crystals was practically independent of Fb orientation in the plane perpendicular to the direction of the extrusion axis.
Experimental stress–strain σbεb dependencies (Figure 8) illustrate the main stress state during bending of samples. Fracture of the single crystals upon bending in the direction of the bisectrix axis (Figure 2, Fb || <1 2 ¯ 1>) occurs almost simultaneously with the appearance of the first stress breakdown at practically zero plastic deformation. In the meantime, at Fb, directed along the binary axis (Fb || <10 1 ¯ >), single crystals can undergo plastic deformation about 0.08%. In the directions of the binary axes of single crystals (Figure 2a, Fb || <10 1 ¯ >), just after the yield point at 22.7 MPa, the stress-strain curves show progressive plastic deformation with sharp drops and rises in stress (Figure 8a).
For the samples after extrusion at Ke 3, by analogy with the case of bending of single crystals, the dependences σbεb were investigated for two variants of orientation of the bending force Fb, namely (a) Fb || <10 1 ¯ >, that is, parallel to the predominant direction of the binary axes of crystallites (Figure 2a), which coincides with the direction of these axes of the initial single crystal and (b) Fb || <1 2 ¯ 1>, that is, parallel to the predominant direction of the bisectrix axis OY. Both orientations of the bending force Fb were found to have little effect on bending strength of extruded samples at 80 K. For example, for Fb || <10 1 ¯ > or <1 2 ¯ 1> directions, the bending strength was 31 ± 3 and 28 ± 3 MPa at Ke 1.2 as well as 42 ± 3 and 41 ± 3 MPa at Ke 3, respectively.
For samples extruded at Ke ≥ 3, the dependences σb–εb were investigated at an arbitrary orientation of the bending force Fb in the plane perpendicular to the extrusion axis. For specimens extruded at Ke = 1.2, the bending force was oriented along the predominant direction of the binary axes. As a result of the extrusion of Bi–Sb crystals at Ke ≥ 5, a TE material with an axially symmetric structure is formed. The extruded samples were investigated in the temperature range 80–300 Κ before and after annealing at 453 K for 8 h. For samples extruded at Ke = 10, the stress–strain dependences at room temperature had a relatively pronounced nonlinear region without significant disruptions of the σb values with increasing εb up to the destruction of the sample. With decreasing temperature, a decrease in the nonlinear section on the σbεb curves was observed. At the lowest temperature of 80 K, the nonlinear section on these curves corresponding to plastic deformation was practically absent (Figure 8b). As a result of extrusion at Ke ≥ 5, a polycrystalline structure with a high ultimate bending strength was formed, reaching values over 70 MPa in the temperature range 80–300 K.
Thus, the data obtained on the effect of extrusion on ultimate strength to bending show that plastic deformation of Bi–Sb single crystals under conditions of high hydrostatic pressure, even at low extrusion ratios, makes it possible to obtain a thermoelectric material with a significantly higher ultimate strength to bending compared to the original single crystals in the entire temperature range 80–300 Κ. In addition, in the same temperature range at different extrusion ratios, elongation-to-fracture (plastic deformation) of the extruded crystals increases two to three times compared to single-crystal samples, remaining rather low (Figure 9).

4. Discussion

As follows from Figure 7, Figure 8 and Figure 9, plasticity (elongation-to-fracture) and the strength of extruded samples at high hydrostatic pressure can reach values several times higher than those for single crystals. This is consistent with the data of compression tests of Fe3Si single crystals [20] and AISI 4330 steel [21,22]. As reported by Lorrek and Pawelski [20], brittle materials are usually successfully deformed by additional applying high hydrostatic pressure, which leads to an increase in the plasticity of single crystals of iron silicide by several times, especially at a certain crystallographic orientation of the sample relative to the loading axis (Figure 10a). In the Tadano and Hagihara model [21], the calculated flow stresses under uniaxial compression are very close to the experimental results obtained by Spitzig et al. [22] at hydrostatic pressures p = 0.552, and 1104 MPa for AISI 4330 steel (Figure 10b). In this analysis, the hydrostatic pressure is applied as the initial stress, which means that σ11 = σ22 = σ33 = −p at the initial state. In general, the ultimate plastic deformability of a material is a function of stress configuration, strain rate, and temperature. The effect of the stress configuration is the most significant and characterized by the mean normal stress σm = 1/3 (σ1 + σ2 + σ3), which may not be low enough to avoid cracking or fracture during the deformation process of a brittle material. In this case, σm can be reduced by applying hydrostatic pressure.
The significant stability of the structure of the samples obtained by extrusion of single crystals in a high pressure liquid medium at Ke ≤ 3 is explained by the fact that the all-round hydrostatic compression used in the process of extrusion increases the uniformity of crystal plastic deformation, which leads to the formation of a more stable polycrystalline structure of extruded samples, consisting of blocks with an orientation main crystallographic directions close to the original single crystal.
As follows from Figure 5, annealing of the extruded samples causes a decrease in the specific content of the crystallographic texture W<111>, which reflects intense recrystallization leading to reorientation and coarsening of polycrystalline grains. However, such heat treatment improves the TE properties of the alloy [1,2]. Of course, the value of W<111> reaches the maximum if the trigonal axis of the single crystals is directed along the extrusion axis.
At low values of extrusion ratio (Ke ≤ 3), Bi–Sb polycrystalline alloy is strongly textured and has a preferred orientation of the crystallographic axes of grains coinciding with the orientation of the corresponding crystallographic axes of the single crystal before extrusion. However, a higher plastic deformation of single crystals (Ke ≥ 5) leads to practically isotropic grains in the plane perpendicular to the extrusion axis (Figure 6c). At the highest plastic deformation of the single crystal at Ke = 10, the axial texture <111> dominates in the direction of the extrusion axis, and the structure of the polycrystalline extruded material is practically independent of the orientation of a single crystal before extrusion.
At room temperature, elastic elongation of Bi0.91Sb0.09 alloy depends on applied stress, which can be calculated from Young’s modulus and is about 0.08% (Figure 8). Gopinathan and Padmini [23] observed anomalous variations in elastic properties in the Sb concentration range from 2.2 to 14.89 at.%, in which, for example, Young’s modulus (E) varied nonmonotonically from 25 to 37.5 GPa. In this defined region of critical concentrations, this is interpreted in terms of changes in lattice parameter and carrier concentration. In accordance to Woodcox et al. [24], the E values for the Bi0.91Sb0.09 alloy are in the range from 26 to 32 GPa. Taking the modulus value equal to 30 GPa, and stress at the transition point of the linear dependence σb–εb into the nonlinear one (yield point) equal to 23 MPa, we find elastic deformation is 23 × 100%/30,000 = 0.077%. Thus, the plastic deformation (elongation-to-fracture) of extruded samples at 300 K, obtained at extrusion ratios of 3 and 10, is 0.08 and 0.28%, respectively (Figure 8b and Figure 9). At 80 K, the plastic deformation of the extruded samples is zero and 0.09% for the Ke values of 3 and 10, respectively.
As shown above in Figure 8a, at bending force directed along the binary axis OZ (Fb || <10 1 ¯ >), the stress–strain curves of single crystals show sharp drops and rises after the yielding point at 22.7 MPa. The occurrence of sharp sawtooth stress drops or serrations during compression and tension deformations of bismuth and antimony single crystals due to twinning was indicated earlier [25,26,27,28]. Klassen-Neklyudova [25] summarized that elastic and residual twinning is formed under loading of bismuth and antimony crystals. Slonaker and Smutz [26] found that the transition from slip to twinning as the predominant deformation mechanism of 99.99% Bi was observed at tensile deformation when the angle Θ between the normal to the true cleavage plane (111), and the major axis of the single crystal equals 70°. The value of the critical resolved shear stress for slip on the (111) plane in bismuth at room temperature, obtained by Steegmuller and Daniel [27], is in agreement with the results of Slonaker and Smutz [26]. Deformation modes observed after compression tests at the strain rate of 1.7 × 10−4 on bismuth crystals of various orientations, mainly at the temperature −80 °C (193 K) and more, are slip on (111) plane and twinning along the (011), (101), and (110) planes [27]. In the tensile test, twinning easily occurs at a strain rate of 1 × 10−3 s−1 [28] and is suppressed at a strain rate of 1 × 10−4 s−1 or lower [29,30].
The destruction of single crystals upon bending when Fb || <1 2 ¯ 1> occurs almost simultaneously with the appearance of the first breakdown of stress at a deformation of about 0.05%, while with Fb directed along the binary axis (Fb || <10 1 ¯ >) single crystals can undergo plastic deformation three times greater. Thus, when the bending force Fb is directed along the binary or bisectrix axis of the single crystal, differences in σb values are apparently determined by the conditions for the formation of microcracks μ (Figure 11) in the vicinity of the B–C line (Figure 2b), where maximum tensile forces T and stresses act upon the bending of single crystals [19,31].
Under the action of sufficient compression or tension stresses directed along one of the equivalent <100> axes, twins can be formed in Bi–Sb single crystals with a rhombohedral lattice symmetry in the form of thin interlayers with twinning planes {110} and twinning directions <100> [25]. With this bending scheme, one of the {110} twinning planes makes an angle of 71° with the (111) plane, and one of the twinning directions <100> is perpendicular to the binary or bisector axis coinciding with the B–C line, depending on the orientation Fb. When a single crystal is bent, tensile forces T act in the vicinity of the B–C line. As a result, conditions are realized for the formation of twinning layers along the {110} plane. The symmetry of the tensile forces upon bending of single crystals with respect to the middle cross-section ABCD (Figure 11), which coincides with the (111) plane, leads to the formation of a twin structure symmetric with respect to the ABCD plane. As a result, twin intersection regions appear in the vicinity of the BC line.
In the region of intersection of twins, there is a high density of dislocations and the appearance of cavities—Rose channels [32]—that have been actually reported as early as 1868 by Rose, who described these channels, or voids, formed by intersecting twins in calcite [24,31]. These were later described as microcracks in body-centered cubic (BCC) metals by Priestner [33], and by Sleeswyk [34] to explain ductile–brittle fracture transition in BCC iron through a complex emissary dislocation proposed mechanism. These cavities can penetrate the crystal in the direction of the binary axis at macroscopic distances and are the nuclei of microcracks. Therefore, when a single crystal is bent, a microcrack μ will most likely appear along the direction of the binary axis <10 1 ¯ >. As a result, firstly, microcracks μ most likely arise along the B–C line, which coincides with the direction of one of the binary axes, when Fb || <1 2 ¯ 1>, and secondly, when Fb || <10 1 ¯ >, then the line B–C coincides with one of the bisectrix axes <1 2 ¯ 1> (Figure 11c) in this case, while microcracks μ most likely occur along a binary axis oriented at an angle of 30° to the B–C line.
Obviously, in the case of Fb || <1 2 ¯ 1>, the action of tensile stresses during bending of a single crystal, which form and open a microcrack μ, is more effective than in the case of Fb || <10 1 ¯ >. This explains the higher values of the ultimate bending strength observed during bending of single crystals, when Fb || <10 1 ¯ >, compared with bending of single crystals, when Fb || <1 2 ¯ 1>.
Plastic deformation at compression by an all-round load can develop both due to sliding and due to twinning depending on the orientation of single crystals, temperature, and other factors. For example, for Bi single crystals deformed at a rate of ~10−4 s−1, Steegmuller and Daniel [27] found a slip mode at room temperature for the crystal orientation { 1 ¯ 11}, (111), and {100} and twinning mode for orientations (011), (101), and (110).
In the process of bending of extruded samples, the increased density of dislocations and other structural defects and the development of grain boundaries (GBs) determine a significant resistance to twinning. The twins that appear in individual grains are small in size, and in the areas of their intersection, extended microcracks—which can become nuclei of the destruction of the entire polycrystal—cannot form. This apparently explains the absence of sharp drops in σb during bending of extruded specimens, regardless of the orientation of Fb. In addition, GBs, dislocation pile-ups and other defects that prevent the development of microcracks increase the ultimate strength σb.
Thus, as previously reported [9], the TE properties of the Bi0.91Sb0.09 alloy obtained by extrusion at high hydrostatic pressure are close to those of single crystals, while having several times higher strength (Figure 7 and Figure 8). For example, the dimensionless thermoelectric figure of merit ZT at 80 K decreases from 0.42 to 0.36 with an increase in the extrusion ratio Ke from 1.2 to 10, compared to 0.44 for single crystals measured without a transverse magnetic field. Recently, a unique thermoelectric solid-state multistage cooler for operating temperatures up to ~140 K was developed. In this case, the two-stage TE module for 140–200 K temperature range was developed on the n-type Bi0.91Sb0.09 crystal after plastic deformation and p-type Bi1.6Sb0.4Te3 extruded crystal [35].

5. Conclusions

(1)
A new method was developed to significantly improve the mechanical properties of such effective thermoelectric (TE) materials as Bi(1−x)Sbx single crystals due to their plastic deformation by extrusion under all-round hydrostatic compression at room temperature in a liquid medium.
(2)
A detailed study of n-type Bi–Sb solid solutions in a wide temperature range shows a significant increase in the mechanical strength of Bi–Sb crystals after extrusion in comparison with high-quality Bi–Sb single crystals. The increase in the strength of the Bi–Sb single crystals after plastic deformation is associated with the development of numerous grains with a high boundary surface as well as structural defects, such as dislocations that accumulate at grain boundaries.
(3)
The significant stability of the structure of the extruded samples can be explained by the uniformity of the plastic deformation of the crystal under all-round hydrostatic compression, which leads to the formation of a more stable polycrystalline structure of the thermoelectrics, consisting of blocks with the orientation of the main crystallographic directions close to the original single crystal.
(4)
Strengthening of Bi–Sb crystals after plastic deformation under all-round compression, as reported earlier, makes it possible to develop a unique thermoelectric solid-state multistage cooler for operating temperatures up to T~140 K for the first time.

Author Contributions

Formal analysis, R.S.; Investigation, N.S.; Methodology, N.S.; Supervision, Z.D.; Visualization, N.S. and Y.U.; Writing—original draft, N.S. and Y.U.; Writing—review & editing, Z.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Patient consent was waived.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also form part of an ongoing study.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Goldsmid, H.J. Introduction to Thermoelectricity, 2nd ed.; Springer-Verlag: Berlin/Heidelberg, Germany, 2016; Volume 121. [Google Scholar] [CrossRef]
  2. Lenoir, B.; Scherrer, H. An Overview of Recent Developments for Bi-Sb Alloys. Semicond. Semimet. 2001, 69, 101–137. [Google Scholar] [CrossRef]
  3. Yim, W.M.; Amith, A. Bi-Sb Alloys for Magneto-Thermoelectric and Thermomagnetic Cooling. Solid-State Electron. 1972, 15, 1141–1165. [Google Scholar] [CrossRef]
  4. Zemskov, V.S.; Belaya, A.D.; Beluy, U.S.; Kozhemyakin, G.N. Growth and Investigation of Thermoelectric Properties of Bi-Sb Alloy Single Crystals. J. Cryst. Growth 2000, 212, 161–166. [Google Scholar] [CrossRef]
  5. Sidorenko, N.A.; Ivanova, L.D. Bi-Sb Solid Solutions: Potential Materials for High-Efficiency Thermoelectric Cooling to below 180 K. Inorg. Mater. 2001, 37, 331–335. [Google Scholar] [CrossRef]
  6. Smith, G.E.; Wolfe, R. Thermoelectric Properties of Bismuth-Antimony Alloys. J. Appl. Phys. 1962, 33, 841–846. [Google Scholar] [CrossRef]
  7. Belaya, A.D.; Zayakin, S.A.; Zemskov, V.S.; Kopiev, I.M. Ultimate Bending Strength of Single Crystals of Bi-Sb Solid Solutions. Izv. Acad. Sci. USSR Met. 1990, 4, 166–169. [Google Scholar]
  8. Sidorenko, N.A. Thermoelectric Materials for Peltier Cryogenic Coolers. Cryogenics 1992, 32, 40–43. [Google Scholar] [CrossRef]
  9. Sidorenko, N.; Parashchuk, T.; Maksymuk, M.; Dashevsky, Z. Development of Cryogenic Cooler Based on n-Type Bi-Sb Thermoelectric and HTSC. Cryogenics 2020, 112, 103197. [Google Scholar] [CrossRef]
  10. Martin-Lopez, R.; Zayakin, Z.; Lenoir, B.; Brochin, F.; Dauscher, A.; Scherrer, H. Mechanical Properties of Extruded Bi85Sb15 Alloy Prepared by Mechanical Alloying. Philos. Mag. Lett. 1998, 78, 283–287. [Google Scholar] [CrossRef]
  11. Ivanova, L.D.; Petrova, L.I.; Granatkina, Y.V.; Zemskov, V.S.; Sokolov, O.B.; Skipidarov, S.Y.; Duvankov, N.I. Extruded Materials for Thermoelectric Coolers. Inorg. Mater. 2008, 44, 687–691. [Google Scholar] [CrossRef]
  12. Czochralski, J. Ein Neues Verfahren zur Messung der Kristallisation-Sgeschwindigkeit der Metalle. Z. Physikal. Chem. 1918, 92, 219–221. [Google Scholar] [CrossRef]
  13. Mayanov, E.P.; Gasanov, A.A.; Naumov, A.V. The Czochralski Method (cz): History and Development. Mater. Electron. Eng. 2018, 19, 59–70. [Google Scholar] [CrossRef] [Green Version]
  14. Zemskov, V.S.; Belaya, A.D.; Kozhemyakin, G.N.; Lyuttsaw, V.G.; Kostyukova, Y.P. Growth of Single Crystals of Bismuth-Antimony Alloys by Chochralski Method. J. Cryst. Growth 1985, 71, 243–245. [Google Scholar] [CrossRef]
  15. Schulz, L.G. A Direct Method of Determining Preferred Orientation of a Flat Reflection Sample Using a Geiger Counter X-ray Spectrometer. J. Appl. Phys. 1949, 20, 1030–1033. [Google Scholar] [CrossRef]
  16. Raghothamachar, B.; Dhanaraj, G.; Bai, J.; Dudley, M. Defect Analysis in Crystals Using X-ray Topography. Microsc. Res. Tech. 2006, 69, 343–358. [Google Scholar] [CrossRef]
  17. Weissmann, S.; Lee, L.H. Applications of the Divergent Beam X-ray Technique. Prog. Cryst. Growth Charact. 1989, 18, 205–226. [Google Scholar] [CrossRef]
  18. Sidorenko, N.A.; Tsvetkova, N.A.; Dashevskii, Z.M. A Method for Preferential Structural Orientation Measurement in Thermoelectric Materials. Supercond. Sci. Technol. 1993, 6, 62–66. [Google Scholar] [CrossRef]
  19. Callister, W.D., Jr. Materials Science and Engineering, 6th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2003. [Google Scholar]
  20. Lorrek, W.; Pawelski, O. The Influence of Hydrostatic Pressure on the Plastic Deformation of Metallic Materials. In Proceedings of the Fifteenth International Machine Tool Design and Research Conference; Palgrave: London, UK, 1975; pp. 703–712. [Google Scholar]
  21. Tadano, Y.; Hagihara, S. Hydrostatic Pressure Dependent Crystal Plasticity by Homogenization-based Finite Element Method. ISIJ Int. 2016, 56, 700–707. [Google Scholar] [CrossRef] [Green Version]
  22. Spitzig, W.A.; Sober, R.J.; Richmond, O. Pressure Dependence of Yielding and Associated Volume Expansion in Tempered Martensite. Acta Metall. 1975, 23, 885–893. [Google Scholar] [CrossRef]
  23. Gopinathan, K.K.; Padmini, A.R.K.L. Ultrasonic Studies on Bismuth-Antimony Alloys. J. Phys. D Appl. Phys. 1974, 7, 32–40. [Google Scholar] [CrossRef]
  24. Woodcox, M.; Young, J.; Smeu, M. Ab Initio Investigation of the Elastic Properties of Bismuth-Based Alloys. Phys. Rev. B 2019, 100, 104105. [Google Scholar] [CrossRef]
  25. Klassen-Neklyudova, M.V. Mechanical Twinning of Crystals; Springer: Berlin/Heidelberg, Germany, 1964; p. 213. [Google Scholar]
  26. Slonaker, R.E.; Smutz, M. Factors Affecting the Growth and the Mechanical and Physical Properties of Bismuth Single Crystals. J. Less Common Met. 1965, 8, 327–338. [Google Scholar] [CrossRef] [Green Version]
  27. Steegmuller, C.; Daniel, J.S. Slip in Bismuth. J. Less Common Met. 1972, 27, 81–85. [Google Scholar] [CrossRef]
  28. Yanaka, Y.; Kariya, Y.; Watanabe, H.; Hokazono, H. Plastic Deformation Behavior and Mechanism of Bismuth Single Crystals in Principal Axes. Mater. Trans. 2016, 57, 819–823. [Google Scholar] [CrossRef] [Green Version]
  29. Motohashi, I.; Otake, S. Plastic Deformation and Dislocation Etch-Pit Distribution Duel the (001) <110> Glide System for Bismuth Single Crystals. J. Jpn. Inst. Met. 1973, 37, 978–985. [Google Scholar] [CrossRef] [Green Version]
  30. Otake, S.; Namazuea, H.; Matsuno, N. Critical Resolved Shear Stresses of Two Slip Systems in Bismuth Single Crystals. Jpn. J. Appl. Phys. 1980, 19, 433–437. [Google Scholar] [CrossRef]
  31. ASTM International. Standard Test Methods for Flexural Properties of Unreinforced and Reinforced Plastics and Electrical Insulating Materials; ASTM International: West Conshohocken, PA, USA, 2003. [Google Scholar]
  32. Rose, G. Über die im Kalkspat Vorkommenden Hohlen Canäle; Akademie Der Wissenschaften: Berling, Germany, 1868; Volume 23, pp. 57–79. [Google Scholar]
  33. Priestner, R. The Relationship between Brittle Cleavage and Deformation Twinning in BCC Metals. In Deformation Twinning; Reed-Hill, R.E., Hirth, J.P., Rogers, H.C., Eds.; Grodon and Breach Science: New York, NY, USA, 1964; pp. 321–355. [Google Scholar]
  34. Sleeswyk, A.W. Emissary Dislocations: Theory and Experiments on the Propagation of Deformation Twins in a-Iron. Acta Metall. 1962, 10, 705–725. [Google Scholar] [CrossRef]
  35. Parashchuk, T.; Sidorenko, N.; Ivantsov, L.; Sorokin, A.; Maksymuk, M.B.; Dzundza, B.; Dashevsky, Z. Development of Solid-State Multi-Stage Thermoelectric Cooler. J. Power Sources 2021, 496, 229821. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of plastic deformation of Bi–Sb crystals at high hydrostatic pressure (a) and the view of the extruder (b). 1—hydrostat plunger, 2—high pressure chamber, 3—working fluid, 4—extruder plunger, 5—Bi–Sb single crystal, 6—die, 7—as extruded polycrystalline TE material, 8—lower support ring.
Figure 1. Schematic diagram of plastic deformation of Bi–Sb crystals at high hydrostatic pressure (a) and the view of the extruder (b). 1—hydrostat plunger, 2—high pressure chamber, 3—working fluid, 4—extruder plunger, 5—Bi–Sb single crystal, 6—die, 7—as extruded polycrystalline TE material, 8—lower support ring.
Electronicmat 02 00036 g001
Figure 2. The main axes of the Bi–Sb crystal lattice (a) and the bending test scheme (b). (a) a1, a2, and a3 are binary axes in a hexagonal lattice, where a1 coincides with the OX axis, OY is the bisectrix axis, and OZ coincides with the trigonal and extrusion axes. (b) 1—loading prism, 2—sample, 3—supports, B–C—line of maximum tensile forces Ts; h and b are the height and width of the specimen, respectively, l—distance between the supports. Bending force Fb is parallel to the extrusion axis (Fb || OZ || <10 1 ¯ > or <1 2 ¯ 1>).
Figure 2. The main axes of the Bi–Sb crystal lattice (a) and the bending test scheme (b). (a) a1, a2, and a3 are binary axes in a hexagonal lattice, where a1 coincides with the OX axis, OY is the bisectrix axis, and OZ coincides with the trigonal and extrusion axes. (b) 1—loading prism, 2—sample, 3—supports, B–C—line of maximum tensile forces Ts; h and b are the height and width of the specimen, respectively, l—distance between the supports. Bending force Fb is parallel to the extrusion axis (Fb || OZ || <10 1 ¯ > or <1 2 ¯ 1>).
Electronicmat 02 00036 g002
Figure 3. An X-ray diffraction pattern (a) and diffraction spots (b) of Bi0.91Sb0.09 grown crystals.
Figure 3. An X-ray diffraction pattern (a) and diffraction spots (b) of Bi0.91Sb0.09 grown crystals.
Electronicmat 02 00036 g003
Figure 4. Diffraction patterns for a crystal Bi0.91Sb0.09 after plastic deformation. (a) Textureless reference sample, (b) sample after extrusion at Ke = 5 and annealing at 453 K for 8 h. Miller indices are given in the hexagonal setting for the [111] planes; is the doubled Bragg angle.
Figure 4. Diffraction patterns for a crystal Bi0.91Sb0.09 after plastic deformation. (a) Textureless reference sample, (b) sample after extrusion at Ke = 5 and annealing at 453 K for 8 h. Miller indices are given in the hexagonal setting for the [111] planes; is the doubled Bragg angle.
Electronicmat 02 00036 g004
Figure 5. Influence of the extrusion ratio on the content of the crystallographic texture W<111> in as extruded Bi0.91Sb0.09 crystals (1) and after annealing at 453 K for 8 h (2).
Figure 5. Influence of the extrusion ratio on the content of the crystallographic texture W<111> in as extruded Bi0.91Sb0.09 crystals (1) and after annealing at 453 K for 8 h (2).
Electronicmat 02 00036 g005
Figure 6. SEM microphotographs of cleavage features (ac) and fracture surface of a typical sample (d) in Bi0.91Sb0.09 extruded samples. (a) General view of the cleaved surface perpendicular to the extrusion axis with triangular etching pits; (bd) annealing at 453 K for 8 h with Ke 1.2 (a), 3 (b), and 5 (c) at 1000 MPa and 300 K.
Figure 6. SEM microphotographs of cleavage features (ac) and fracture surface of a typical sample (d) in Bi0.91Sb0.09 extruded samples. (a) General view of the cleaved surface perpendicular to the extrusion axis with triangular etching pits; (bd) annealing at 453 K for 8 h with Ke 1.2 (a), 3 (b), and 5 (c) at 1000 MPa and 300 K.
Electronicmat 02 00036 g006
Figure 7. Influence of the extrusion ratio on the ultimate bending strength of Bi0.91Sb0.09 crystals. The value of Ke = 1 corresponds to single-crystal samples. Test temperature was 80 K.
Figure 7. Influence of the extrusion ratio on the ultimate bending strength of Bi0.91Sb0.09 crystals. The value of Ke = 1 corresponds to single-crystal samples. Test temperature was 80 K.
Electronicmat 02 00036 g007
Figure 8. Stress–strain curves for Bi0.91Sb0.09 single crystals (a) and extruded polycrystalline material (b) at different temperatures and extrusion ratios. (a) Bending force direction is perpendicular to the trigonal axis OZ and along the binary axis OX and the bisectrix OY axis (Fb || <10 1 ¯ > and Fb || <1 2 ¯ 1>, respectively); 200 K. (b) Ke of 3 and 10 (solid and dashed lines, respectively); 80 K (1, 2) and 300 K (3, 4).
Figure 8. Stress–strain curves for Bi0.91Sb0.09 single crystals (a) and extruded polycrystalline material (b) at different temperatures and extrusion ratios. (a) Bending force direction is perpendicular to the trigonal axis OZ and along the binary axis OX and the bisectrix OY axis (Fb || <10 1 ¯ > and Fb || <1 2 ¯ 1>, respectively); 200 K. (b) Ke of 3 and 10 (solid and dashed lines, respectively); 80 K (1, 2) and 300 K (3, 4).
Electronicmat 02 00036 g008
Figure 9. Elongation-to-fracture of single crystals (Fb || <10 1 ¯ > or <1 2 ¯ 1>) and extruded samples depending on temperature and extrusion ratio Ke.
Figure 9. Elongation-to-fracture of single crystals (Fb || <10 1 ¯ > or <1 2 ¯ 1>) and extruded samples depending on temperature and extrusion ratio Ke.
Electronicmat 02 00036 g009
Figure 10. The effect of applied hydrostatic pressure p on true plastic strain and stress in compression tests of Fe3Si single crystals [20] (a) and AISI 4330 steel [21] (b).
Figure 10. The effect of applied hydrostatic pressure p on true plastic strain and stress in compression tests of Fe3Si single crystals [20] (a) and AISI 4330 steel [21] (b).
Electronicmat 02 00036 g010
Figure 11. Scheme of the formation of a microcrack μ during bending of Bi–Sb single crystals: (a) The formation of a microcrack μ along the binary axis <10 1 ¯ > in the region of intersection of twins S and S′. (b,c) The location of the microcrack μ in the ABCD plane near the line of maximum tensile stresses at Fb || <1 2 ¯ 1> and Fb || <10 1 ¯ >, respectively.
Figure 11. Scheme of the formation of a microcrack μ during bending of Bi–Sb single crystals: (a) The formation of a microcrack μ along the binary axis <10 1 ¯ > in the region of intersection of twins S and S′. (b,c) The location of the microcrack μ in the ABCD plane near the line of maximum tensile stresses at Fb || <1 2 ¯ 1> and Fb || <10 1 ¯ >, respectively.
Electronicmat 02 00036 g011
Table 1. Influence of the extrusion ratio and annealing temperature for 8 h on the content of the crystallographic texture (W<111> × 100%) in the direction of the extrusion axis.
Table 1. Influence of the extrusion ratio and annealing temperature for 8 h on the content of the crystallographic texture (W<111> × 100%) in the direction of the extrusion axis.
KeNo AnnealingAnnealing
150 °C180 °C210 °C
1.288 ± 483 ± 580 ± 572 ± 5
372 ± 562 ± 555 ± 538 ± 4
560 ± 548 ± 441 ± 423 ± 3
1050 ± 433 ± 426 ± 414 ± 2
Table 2. Pole density Phkl of the normals to reflecting planes (hkl) in the direction of the extrusion axis for Bi0.91Sb0.09 crystals, before and after annealing at 180 °C for 8 h.
Table 2. Pole density Phkl of the normals to reflecting planes (hkl) in the direction of the extrusion axis for Bi0.91Sb0.09 crystals, before and after annealing at 180 °C for 8 h.
KeAs ExtrudedAfter Annealing
1.235101.23510
P003292115.210.1251695
P1020.10.30.40.50.20.40.20.4
P1040.10.20.40.60.20.20.40.3
P1050.30.60.70.80.50.40.91.3
P0221.41.61.31.21.71.91.11.6
P1070.30.60.50.70.40.71.01.1
P1160.20.30.30.50.20.30.60.4
P2120.10.50.60.80.30.40.91.4
P1080.81.00.90.81.21.31.21.0
P2140.10.40.40.30.20.60.50.5
Table 3. Bending strength of extruded Bi–Sb polycrystals depending on the test temperature.
Table 3. Bending strength of extruded Bi–Sb polycrystals depending on the test temperature.
Keσb, MPa
300 K200 K150 K80 K
1.227 ± 327 ± 329 ± 331 ± 3
335 ± 336 ± 338 ± 342 ± 3
548 ± 452 ± 455 ± 457 ± 4
1065 ± 468 ± 470 ± 472 ± 4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sidorenko, N.; Unigovski, Y.; Dashevsky, Z.; Shneck, R. A Novel Method to Significantly Improve the Mechanical Properties of n-Type Bi(1−x)Sbx Thermoelectrics Due to Plastic Deformation. Electron. Mater. 2021, 2, 511-526. https://doi.org/10.3390/electronicmat2040036

AMA Style

Sidorenko N, Unigovski Y, Dashevsky Z, Shneck R. A Novel Method to Significantly Improve the Mechanical Properties of n-Type Bi(1−x)Sbx Thermoelectrics Due to Plastic Deformation. Electronic Materials. 2021; 2(4):511-526. https://doi.org/10.3390/electronicmat2040036

Chicago/Turabian Style

Sidorenko, Nikolay, Yaakov Unigovski, Zinovi Dashevsky, and Roni Shneck. 2021. "A Novel Method to Significantly Improve the Mechanical Properties of n-Type Bi(1−x)Sbx Thermoelectrics Due to Plastic Deformation" Electronic Materials 2, no. 4: 511-526. https://doi.org/10.3390/electronicmat2040036

Article Metrics

Back to TopTop