Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

A giant Nernst power factor and figure-of-merit in polycrystalline NbSb2 for Ettingshausen refrigeration

Peng Li ab, Pengfei Qiu *abc, Jie Xiao a, Tingting Deng c, Lidong Chen ab and Xun Shi *ab
aState Key Laboratory of High Performance Ceramics and Superfine Microstructure, Shanghai Institute of Ceramics, Chinese Academy of Science, Shanghai 200050, China. E-mail: qiupf@mail.sic.ac.cn; xshi@mail.sic.ac.cn
bCenter of Materials Science and Optoelectronics Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
cSchool of Chemistry and Materials Science, Hangzhou Institute for Advanced Study, University of Chinese Academy of Sciences, Hangzhou 310024, China

Received 8th May 2023 , Accepted 31st July 2023

First published on 2nd August 2023


Abstract

Ettingshausen refrigeration is a promising solid-state refrigeration technology that can be used in exploring the quantum state of matters and superconducting materials, but its development is greatly limited by the lack of high-performance polycrystalline thermomagnetic materials. In this work, we report that the polycrystalline topological semimetal NbSb2 has a giant Nernst power factor ((PF)N) of 1269 × 10−4 W m−1 K−2 under 9 T at 28 K and a Nernst figure-of-merit (zN) of 28.5 × 10−4 K−1 under 9 T at 18 K, both of which are record-high values among the polycrystalline thermomagnetic materials. The observed high thermomagnetic performance is mainly attributed to the large and unsaturated Nernst thermopower under a high magnetic field. Due to the diminished anisotropy, the polycrystalline NbSb2 has similar high carrier mobility to the single-crystalline NbSb2 in the bc-plane, yielding large Nernst thermopower in low temperature ranges. Combining the excellent thermomagnetic performance and low-cost and time-saving fabrication process, polycrystalline NbSb2 is a very competitive candidate material for Ettingshausen refrigeration.



Broader context

The demand for thermoelectric applications at low temperatures is heightened now by the exploration of quantum state of matters, quantum information science, and space science. However, today good thermoelectric and thermomagnetic materials in the low temperature range, particularly at temperatures below the liquid nitrogen boiling point, are still very rare. Discovering high performance thermoelectric and thermomagnetic materials below the liquid nitrogen temperature has become a very active field worldwide. In this work, we demonstrate that polycrystalline NbSb2 has a giant Nernst power factor of 1269 × 10−4 W m−1 K−2 under 9 T at 28 K and a Nernst figure-of-merit of 28.5 × 10−4 K−1 under 9 T at 18 K. These values are not only the highest among the polycrystalline thermomagnetic materials, but are also superior to many single-crystalline thermomagnetic materials reported before. More importantly, polycrystalline NbSb2 can be easily synthesized by using a low-cost and time-saving fabrication process, which is more competitive than the single-crystalline thermomagnetic materials for the fabrication of an Ettingshausen refrigerator. This work provides a new optional material for solid-state heat pumping below liquid nitrogen temperature based on Ettingshausen refrigeration.

Introduction

A low-temperature environment is essential for the investigation of quantum states of matter,1 superconducting materials,2 and space science.3 Beyond the traditional compressor refrigerator based on gaseous working substances, thermoelectric (TE) refrigeration provides an alternative way to obtain a low-temperature environment.4–7 When a current flows through the thermoelectric (TE) device, the temperature at one end of the device can be reduced. According to the working principle, TE refrigeration can be divided into two categories, termed as Peltier refrigeration and Ettingshausen refrigeration.8 Besides the electric current, the latter one also needs the assistance of a magnetic field.

Compared with Peltier refrigeration, which has already realized commercialization several decades before,4,5 the development of Ettingshausen refrigeration is relatively slow. The main reason is the lack of high performance thermomagnetic materials with a high Nernst power factor ((PF)N = Syx2σyy, determining the transverse pumping power, where Syx is the Nernst thermopower and σyy is the longitudinal electrical conductivity) and large Nernst figure-of-merit (zN = Syx2σyy/κxx, determining the maximum temperature gradient that can be built across the device, where κxx is the transverse thermal conductivity).8 To maximize (PF)N and zN, the material for Ettingshausen refrigeration requires electrons and holes possessing nearly identical concentrations and high mobilities, which greatly limit the scope of optional materials.9 Thus, for a long time, the investigation of Ettingshausen refrigeration was limited in a few thermomagnetic materials (e.g. Bi–Sb alloys10,11 and In–Sb alloys12). The peak (PF)N and zN are about 324 × 10−4 W m−1 K−2 and 55 × 10−4 K−1 for single-crystalline Bi97Sb3 under 1 T at 115 K, respectively.10

Recently, the investigation of Ettingshausen refrigeration was rejuvenated due to the great progress achieved in topological semimetals.13–15 A series of topological semimetals have been reported with high thermomagnetic performance, such as Cd3As2,16 ZrTe5,17 PtSn4,18 Mg2Pb,19 NbSb2,20 and WTe2.21 Some of them exhibit superior (PF)N to the Bi–Sb alloys, such as 425 × 10−4 W m−1 K−2 for single-crystalline Mg2Pb under 10 T at 30 K,19 3800 × 10−4 W m−1 K−2 for single-crystalline NbSb2 under 5 T at 25 K,20 and 3 W m−1 K−2 for single-crystalline WTe2 under 9 T at 11.3 K.21 The maximum zN of thermomagnetic materials was boosted up to 265 × 10−4 K−1 for single-crystalline WTe2 under 9 T at 11.3 K.21 However, all these high (PF)N and zN are achieved in single crystals, in which the preparation methods are usually time-consuming and expensive. Furthermore, the crystal sizes are usually small, which greatly limit the real application of Ettingshausen refrigeration. In contrast to single crystals, the polycrystals can be easily prepared into large bulks by using simple and low-cost preparation methods, which are more suitable for real application. However, currently the investigation on polycrystalline thermomagnetic materials is very rare.22–25 Particularly, as shown in Fig. 1, the (PF)N and zN of the reported polycrystalline thermomagnetic materials are much lower than those of single-crystalline thermomagnetic materials. Discovering the polycrystalline thermomagnetic materials with high (PF)N and zN has already become an important task for the investigation of Ettingshausen refrigeration.


image file: d3ee01450a-f1.tif
Fig. 1 (a) Comparisons of the (a) Nernst power factor (PF)N and (b) Nernst figure-of-merit zN of polycrystalline NbSb2 and some typical thermomagnetic materials reported before.10,16–21,23–28 The maximum (PF)N and zN of each thermomagnetic material are used in the figures.

In this work, we report the discovery of a colossal (PF)N of 1269 × 10−4 W m−1 K−2 under 9 T at 28 K and a high zN of 28.5 × 10−4 K−1 under 9 T at 18 K in polycrystalline NbSb2, both of which are record-high values in polycrystalline thermomagnetic materials (Fig. 1). The excellent performance of polycrystalline NbSb2 is caused by the nearly identical electron and hole concentrations near the Fermi level, ultrahigh mobilities close to 1 m2 V−1 s−1, and a strong phonon-drag effect in low temperature ranges. This work provides a novel potential candidate material for Ettingshausen refrigeration.

Results and discussion

Polycrystalline NbSb2 was fabricated by using the solid-state reaction and spark plasma sintering technique. Fig. 2a shows the powder X-ray diffraction pattern of polycrystalline NbSb2. All diffraction peaks can be indexed belonging to the centrosymmetric monoclinic structure (space group C2/m) of NbSb2 (the inset in Fig. 2a). Backscattered electron (BSE) image (Fig. 2b) and energy dispersive X-ray spectroscopy (EDS) mapping (Fig. 2c and d) indicate that Nb and Sb are homogeneously distributed inside the matrix, confirming that the prepared polycrystalline NbSb2 is phase pure. Fig. 2e shows that the grain sizes are about several micrometers. No obvious orientation is observed. Such high homogeneity and isotropy provide the basis for the accurate measurement of thermomagnetic performance in this work.
image file: d3ee01450a-f2.tif
Fig. 2 (a) Powder X-ray diffraction pattern for polycrystalline NbSb2. The inset shows the crystal structure of NbSb2. (b) Backscattered electron (BSE) image and (c) and (d) energy dispersive X-ray spectroscopy (EDS) mapping performed on polycrystalline NbSb2. (e) Cross-section image of fractured polycrystalline NbSb2.

Fig. 3a shows the temperature dependence of the Seebeck thermopower Sxx of polycrystalline NbSb2 under different magnetic fields B. In the absence of a magnetic field, the Sxx of polycrystalline NbSb2 below 100 K is almost zero. This is consistent with the band structure of NbSb2, with the schematics shown in the inset of Fig. 3a. The Fermi level crosses both the valence and conduction bands, demonstrating a typical feature of semimetals. Above 100 K, the Sxx decreases linearly with increasing temperature, reaching −20 μV K−1 at 300 K. When the magnetic field is applied, the Sxx below 100 K is significantly enhanced, while that above 100 K is only slightly increased. Similar temperature-dependence and magnetic field-dependence have been also observed in single-crystalline NbSb2,20 indicating that the grain boundary has little influence on the Sxx.


image file: d3ee01450a-f3.tif
Fig. 3 Temperature dependences of (a) Seebeck thermopower Sxx, (b) Nernst thermopower Syx, (c) electrical conductivity σyy, and (d) Nernst power factor (PF)N for polycrystalline NbSb2 under different magnetic fields. The inset in (a) represents the schematics of the band structure of NbSb2 near the Fermi level.

Fig. 3b shows the temperature dependence of the Nernst thermopower Syx of polycrystalline NbSb2 under different magnetic fields B. Under each magnetic field, the absolute value of Syx firstly increases with the increase of temperature, reaching a peak around 21 K, and then decreases for higher temperature. Likewise, with the increase of magnetic field, the Syx is monotonously increased, which can be more intuitively reflected in Fig. S2 (ESI). Under the same magnetic field and temperature, the Syx of polycrystalline NbSb2 is smaller than that of single-crystalline NbSb2. For example, the Syx of polycrystalline NbSb2 is 396 μV K−1 under 9 T at 21 K, which is 36% lower than that of single-crystalline NbSb2 (616 μV K−1) under the same conditions.20 Thus, the grain boundaries in polycrystalline NbSb2 lead to decreased Syx.

To obtain the Nernst power factor (PF)N, a longitudinal electrical conductivity σyy is required. Under a magnetic field, the motion of the carriers is deflected by the Lorentz force. The motion of the carriers is subject to the influence from other directions. In this case, the σyy of polycrystalline NbSb2 can be described as29

 
image file: d3ee01450a-t1.tif(1)
where ρxx is the transverse electrical resistivity and ρyx is the Hall resistivity. In the absence of magnetic field, the ρxx of polycrystalline NbSb2 increases with increasing temperature, which is consistent with the semimetal feature of NbSb2 (Fig. S3a, ESI). When the magnetic field is applied, the ρxx is significantly enhanced, particularly below 100 K, indicating the existence of a strong magnetoresistance effect. The ρyx shown in Fig. S3b (ESI) is negative over the entire temperature range, indicating that the dominant carriers are electrons in polycrystalline NbSb2. Under the same magnetic field and temperature, |ρyx| is much smaller than ρxx, indicating that the contributions of electrons and holes in the electrical transports are close. Fig. 3c shows the calculated σyy for polycrystalline NbSb2. The temperature-dependence and magnetic field-dependence are similar to those of single-crystalline NbSb2, but the value is slightly lower than that of single-crystalline NbSb2 under the same magnetic field and temperature. For example, the σyy of polycrystalline NbSb2 is 0.65 × 106 S m−1 under 9 T at 21 K, which is 20% lower than that of single-crystalline NbSb2 (0.81 × 106 S m−1) under the same conditions.20

Based on the Syx and σyy, the (PF)N of polycrystalline NbSb2 is calculated and plotted in Fig. 3d. Due to the significantly enhanced Syx, the (PF)N is also greatly increased upon applying magnetic field. When B = 9 T, the peak (PF)N of polycrystalline NbSb2 is 1269 × 10−4 W m−1 K−2 at 28 K. Likewise, Fig. S4 (ESI) shows that the (PF)N tends to become saturated above 9 T, and thus further increasing the magnetic field will not significantly increase the (PF)N. The peak (PF)N of polycrystalline NbSb2 is a very exciting result since it is a record-high value among the polycrystalline thermomagnetic materials (Fig. 1a and Fig. S5, ESI). Actually, this value is even higher than most single-crystalline thermomagnetic materials, such as single-crystalline Bi97Sb3,10 Cd3As2,16 ZrTe5,17 PtSn4,18 Mg2Pb,19 and TbPtBi28 (Fig. 1a and Fig. S5, ESI), and thermoelectric materials, such as Bi2Te3-based alloys,30,31 AgSbTe2,32 TaSiTe4,33 and Ag2S-based materials.34–36

The high (PF)N of polycrystalline NbSb2 is mainly contributed by the large Syx. Although the peak Syx of polycrystalline NbSb2 is lower than the single-crystalline NbSb2, it is still much higher than most thermomagnetic materials reported before, such as 124 μV K−1 for polycrystalline NbP under 9 T at 140 K,23 127 μV K−1 for polycrystalline Mg3Bi2 under 13 T at 13.5 K,24 125 μV K−1 for single-crystalline Cd3As2 under 3 T at 350 K,16 45 μV K−1 for single-crystalline PtSn4 under 9 T at 10.3 K,18 200 μV K−1 for single-crystalline Mg2Pb under 10 T at 30 K,19 and 225 μV K−1 for single-crystalline TbPtBi under 14 T at 240 K.28 One main reason for the large Syx is originated from its unsaturated behavior under a high magnetic field. As shown in Fig. S2 (ESI), the Syx linearly increases with increasing magnetic field, without showing obvious deflection even under 9 T.

When both the electrons and holes take part in electrical transports, the Syx can be expressed as37

 
image file: d3ee01450a-t2.tif(2)
where σexx and σhxx are the electrical conductivity contributed by electrons and holes, Sexx and Shxx are the Seebeck thermopower of electrons and holes under the magnetic field, μe and μh are electron mobility and hole mobility, respectively. ne (or nh) and μe (or μh) can be estimated by using the two-carrier model to fit the measured ρxx(B) and ρyx(B)38
 
image file: d3ee01450a-t3.tif(3)
 
image file: d3ee01450a-t4.tif(4)
Fig. S6 (ESI) plots the ρxx(B) and ρyx(B) for polycrystalline NbSb2 under different temperatures and magnetic field. At low temperatures (below 100 K), ρxx(B) varies with the magnetic field B by a factor of 2 and ρyx(B) also significantly deviates off linearity as B increases, which is consistent with the scenario described using eqn (3) and (4) when μB ≫ 1. At high temperatures (above 100 K), the ρxx(B) and ρyx(B) exhibit weak magnetic field dependence since the decreased carrier mobility at high temperatures reduces the contribution of the nonlinear terms in eqn (3) and (4). The fitted ne and nh of polycrystalline NbSb2 are very close in the whole temperature range (Fig. 4a). A similar phenomenon was also observed in single-crystalline NbSb2.20 This is consistent with the band structure of NbSb2, which shows that the electron pocket and hole pocket near the Fermi level have nearly the same volume. Likewise, the fitted μe and μh of polycrystalline NbSb2 are also very close in the whole temperature range (Fig. 4b). In this case, eqn (2) can be simplified into
 
image file: d3ee01450a-t5.tif(5)
where Syx is in proportional to B. This can well explain the observed unsaturated behavior of Syx under a high magnetic field for polycrystalline NbSb2.


image file: d3ee01450a-f4.tif
Fig. 4 (a) Carrier concentrations (ne and nh) and (b) carrier mobilities (μe and μh) of polycrystalline NbSb2. Angular dependence of electrical resistivity ρ measured (c) along the b-axis of single-crystalline NbSb2 and (d) the direction perpendicular the sintering pressure of polycrystalline NbSb2 under 5 T and 9 T at 25 K. (e) Temperature dependence of the difference of Seebeck thermopower of electrons and holes (ShxxSexx) for polycrystalline NbSb2 under different magnetic fields. (f) The difference of Seebeck thermopower of electrons and holes related to the charge carrier diffusion processes (Sed − Shd) and phonons (SepShp) of polycrystalline and single-crystalline NbSb2 at 5 T, respectively. The data of single-crystalline NbSb2 are included in (a), (b), (e), and (f) for comparison.20

Beyond the high magnetic field, eqn (5) indicates that the high carrier mobility is also necessary for the large Syx. The μe and μh of polycrystalline NbSb2 are about 1.7 m2 V−1 s−1 and 1.2 m2 V−1 s−1 at 5 K, which are quite high values among those for the reported thermomagnetic materials, such as Cd3As2 (μe = 6.5 m2 V−1 s−1, μh = 0.5 m2 V−1 s−1 at 10 K)16, PtSn4 (μe = 7.6 m2 V−1 s−1, μh = 7.6 m2 V−1 s−1 at 10 K),18 Mg3Bi2 (μe = 0.48 m2 V−1 s−1, μh = 0.14 m2 V−1 s−1 at 15 K),24 and TbPtBi (μe = 0.3 m2 V−1 s−1, μh = 0.3 m2 V−1 s−1 at 20 K).28 Generally, it is considered that the polycrystals have lower carrier mobility than the single crystals due to the additional grain boundary scattering to carriers. Interestingly, herein the polycrystalline NbSb2 have similar μe and μh with the single-crystalline NbSb2 in the bc-plane. This abnormal phenomenon is related to the anisotropic transport properties of single crystalline NbSb2 at low temperatures. Herein, we measured the angular dependence of the resistivity ρ of single-crystalline NbSb2 under a magnetic field to investigate the anisotropy. As shown in Fig. S7a and b (ESI), for the single-crystalline NbSb2, the angle between the magnetic field and the normal to the bc-plane is noted as θ. When θ = 0°, the position of the magnetic field is perpendicular to the bc-plane. This is the condition used to measure the carrier mobility of single-crystalline NbSb2 in a previous work.20 For the polycrystalline NbSb2, the angle between the magnetic field and the direction of the sintering pressure is noted as θ. When θ = 0°, the position of the magnetic field is parallel to the direction of the sintering pressure. In the case that the carrier mobility is isotropic, the measured angle dependence of ρ should possess a circle shape.

As shown in Fig. 4c, the measured angle dependence of ρ for single-crystalline NbSb2 under 5 T or 9 T at 25 K has a typical two-fold rotational symmetry, which is similar to the crystal structure projection of NbSb2 in the ac-plane (the inset in Fig. 2a). Under 9 T, the ρ at θ = 0° is just 94 × 10−8 Ω m, which is about 47% that at θ = 75° or 255° (198 × 10−8 Ω m). Since the carrier concentration is isotropic, the anisotropic ρ indicates that the carrier mobility of single-crystalline NbSb2 varies considerably in the [010] crystal zone and that in the bc-plane is much lower than those in other planes. In contrast, the measured angle dependence of ρ in polycrystalline NbSb2 at 25 K is nearly a circle (Fig. 4d). At 9 T, the maximum ρ (143 × 10−8 Ω m at around θ = 94° or 274°) is just about 1.09 times the minimum ρ (131 × 10−8 Ω m at around θ = 5° or 185°). This indicates the carrier mobility in the polycrystalline NbSb2 has weak anisotropy, which is consistent with the randomly distributed grains shown in Fig. 2e. Thus, the measured carrier mobility of polycrystalline NbSb2 should simultaneously include the contributions from all planes and grain boundaries. This can explain why the polycrystalline NbSb2 has a high carrier mobility comparable with the single-crystalline NbSb2 in the bc-plane.

Although the polycrystalline NbSb2 has similar carrier mobilities to the single-crystalline NbSb2, it has lower (ShxxSexx). As shown in Fig. 4e, under 9 T, the peak (ShxxSexx) of polycrystalline NbSb2 is about 137 μV K−1, which is about 57% that of single-crystalline NbSb2.20 The (ShxxSexx) includes two parts, termed as the (ShdSed) related to the diffusion process of charge carriers and the (ShpSep) related to the phonon-drag effect.39Fig. 4f shows the (ShdSed) and (ShpSep) of the polycrystalline NbSb2, with the calculation details shown in the Supplementary Information. The former is comparable with that of single-crystalline NbSb2, which is consistent with their similar carrier concentrations shown in Fig. 4a. However, the latter is much lower than that of single-crystalline NbSb2, indicating the weaker phonon-drag effect in polycrystalline NbSb2. The phonon-drag effect is caused by the interaction between the long-wavelength acoustic phonons and the carriers on the Fermi surface.8 In polycrystals,40,41 due to the grain boundary scattering, the relaxation time of the participating long-wavelength phonons is reduced, leading to the suppressed phonon-drag effect.

Fig. 5a and b show the temperature dependence and magnetic field dependence of thermal conductivity κxx for polycrystalline NbSb2. As the temperature increases, the κxx first increases, reaching a peak around 35 K, and then decreases at higher temperature. Upon applying a magnetic field, κxx shows obvious reduction, especially below 30 K (Fig. 5a). This is caused by the suppressed carrier thermal conductivity κc under magnetic field. The κc and lattice thermal conductivity κl can be obtained by fitting the measured κxx according to an empirical formula18,42,43

 
image file: d3ee01450a-t6.tif(6)
where η and s are the two factors related to the thermal mobility and scattering mechanism, respectively. As shown in Fig. 5b, the fitting results agree well with the experimental data. The fitting parameters are shown in Table S1 (ESI). The fitted κl values of polycrystalline NbSb2 are plotted in the inset of Fig. 5a. The peak κl, appearing around 50 K, is 57.1 W m−1 K−1, which is slightly lower than that of the single-crystalline NbSb2 along the b-axis (64.3 W m−1 K−1 at 25 K).20 At 300 K, the κxx is 23.5 W−1 m−1 K−1, which is still much higher than the state-of-the-art materials for Peltier refrigeration around room temperature, such as 0.7 W m−1 K−1 for Bi2Te3,31 0.7 W m−1 K−1 for Mg3Bi2,5 and 0.34 W m−1 K−1 for Ag2Se.44 Under a high magnetic field, such as 9 T, κc is already very low (Fig. S8, ESI), and κl dominates the measured κxx. Thus, reducing the overhigh κl is an important task for the investigation of NbSb2 in the future.


image file: d3ee01450a-f5.tif
Fig. 5 (a) Temperature dependence of thermal conductivity κxx and lattice thermal conductivity κl under different magnetic fields for polycrystalline NbSb2. (b) Magnetic field dependence of thermal conductivity κxx at different temperatures. The symbols are experimental data and the lines are the fitting curves. (c) The ratio of Lorentz number and the Sommerfeld value (L/L0) for polycrystalline NbSb2. (d) Temperature dependence of the Nernst figure-of-merit zN for polycrystalline NbSb2 under different magnetic fields. The data of single-crystalline NbSb2 are included in (c) and (d) for comparison.20

Similar to most topological semimetal materials, the polycrystalline NbSb2 also exhibits the abnormal Lorenz number L deviating off the classic Sommerfeld value L0 below room temperature. Based on the fitted κc, the L can be calculated by using the Wiedemann–Franz law L = κc/(σxxT). Fig. 5c plots the temperature dependence of L/L0 for polycrystalline NbSb2. With increasing temperature, the L/L0 firstly decreases, reaching a minimum around 50 K, and then increases at higher temperature. The violation of the Wiedemann–Franz law at intermediate temperature range might be caused by the inelastic scattering,45–47 while the upturn of L/L0 below and above 50 K might be caused by the changed carrier scattering mechanism to the elastic scattering from the impurities and elastic electron–phonon scattering, respectively. Likewise, the L/L0 of polycrystalline NbSb2 is larger than that of single-crystalline NbSb2, which might be caused by the dilution of inelastic scattering by grain boundary scattering.

Based on the Syx, σyy, and κxx, the Nernst figure-of-merit zN (= Syx2σyy/κxx) is calculated for polycrystalline NbSb2. The detailed derivation process of zN can be found in the Supplementary Information. Similar to the figure-of-merit z for thermoelectric materials, the zN is also a parameter that is independent of the material's geometric factors. As shown in Fig. 5d, the zN increases with increasing temperature, reaches a maximum around 18 K, and then decreases at higher temperature. Likewise, Fig. S9 (ESI) shows that the zN tends to saturate above 9 T, thus further increasing the magnetic field will not significantly increase zN. The peak zN of polycrystalline NbSb2 is already much higher than all the polycrystalline thermomagnetic materials reported before (Fig. 1b and Fig. S10, ESI), such as 0.93 × 10−4 K−1 for polycrystalline NbP under 9 T at 220 K23 and 14.6 × 10−4 K−1 for polycrystalline Mn-doped Mg3Bi2 under 14 T at 14 K.25 Actually, this value is also higher than those of many single-crystalline thermomagnetic materials reported before, such as 8 × 10−4 K−1 for PtSn4 under 9 T at 10.3 K.18 If the overhigh κl can be reduced without deteriorating the electrical transport properties, a higher zN can be expected.

Based on the physical properties of polycrystalline NbSb2, the maximum temperature difference (ΔTmax) and maximum specific heat pumping power (Pmax) of a rectangular Ettingshausen refrigerator made of polycrystalline NbSb2 can be estimated.8,19 Under B = 9 T and Tc = 30 K, the ΔTmax of the NbSb2-based Ettingshausen refrigerator with the thickness of 1 mm along the heat flow direction is about 0.9 K and the theoretical Pmax is about 6.5 W g−1. The latter value is much higher than the compression refrigerator with gas refrigerants (e.g. Pmax = 0.05 W g−1 for He at 5 K, 0.12 W g−1 for H2 at 26 K, and 1.01 W g−1 for N2 at 93 K).19 Thus, the polycrystalline NbSb2 has a great potential to be used for Ettingshausen refrigeration.

Conclusions

In summary, we found that the polycrystalline NbSb2 has an unsaturated Nernst thermopower Syx of 396 μV K−1 under 9 T at 21 K and a giant Nernst power factor (PF)N of ∼1269 × 10−4 W m−1 K−2 under 9 T at 28 K. The latter is almost an order of magnitude larger than the reported (PF)N of other polycrystalline thermomagnetic materials. The zN of polycrystalline NbSb2 is 28.5 × 10−4 K−1 under 9 T at 18 K, which is also a record-high value among the reported polycrystalline thermomagnetic materials. The superior (PF)N and zN are attributed to the nearly identical electron and hole concentrations near the Fermi level, ultrahigh mobilities close to 1 m2 V−1 s−1, and the strong phonon-drag effect in low temperature ranges. Combining the low-cost and time-saving fabrication process, polycrystalline NbSb2 is a very competitive candidate material for Ettingshausen refrigeration.

Experimental section

Sample synthesis

Polycrystalline NbSb2 was synthesized via the solid-stated reaction and spark plasma sintering technique. The niobium powder (alfa, 99.99%) and antimony shot (alfa, 99.9999%) were encapsulated in a vacuum quartz tube and reacted at 1023 K for 48 h. The obtained product was ground into fine powders, loaded into a graphite mold and compacted by spark plasma sintering (SPS, SPS-725, Sojitz, Japan) in a vacuum. The sintering temperature was 1123 K and the sintering pressure was 65 MPa. The holding time at 1123 K is 10 min. The relative density of the sintered product was 98%.

Sample characterization

The phase composition was characterized by X-ray diffraction (XRD, D/max-2550 V, Rigaku, Japan). The elemental distribution was characterized by scanning electron microscopy (SEM, ZEISS supra-55, Germany) with energy dispersive X-ray spectroscopy (EDS, Oxford, UK). The bulk sample was cut into the strips of about 8 × 2 × 2 mm3 and 8 × 2 × 1 mm3 for electrical and thermal transport measurements under the magnetic field, respectively. All measurements were carried out on a physical property measurement system (PPMS, Quantum design, USA). The electrical resistivity and Hall resistivity were measured using a four-probe method and five-probe method, respectively. The Seebeck thermopower and Nernst thermopower were measured using the four-probe method on a standard thermal transport option (TTO) platform and a modified one, respectively. The details can be found in ref. 20. The Syx data used to calculate Nernst power factor and Nernst figure-of-merit are obtained as Syx = [Syx (+B) − Syx (−B))]/2 to eliminate the effects of contact misalignment.

Author contributions

P. L., P. Q., and X. S. designed the work. P. L. synthesized the samples and performed the transport property measurements with the help of J. X., and P. L. and T. D. analyzed the transport properties. L. C. provided helpful discussion. P. L., P. Q., and X. S. analyzed the data and wrote the manuscript.

Conflicts of interest

The authors declare no competing financial interests.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (grants 52122213 and 91963208), the Shanghai Pilot Program for Basic Research-Chinese Academy of Science, Shanghai Branch (JCYJ-SHFY-2022-002), and the Shanghai Government (20JC1415100).

Notes and references

  1. J. Chan, T. P. M. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Gröblacher, M. Aspelmeyer and O. Painter, Nature, 2011, 478, 89–92 CrossRef CAS PubMed.
  2. C. Coats, J. S. Rogers and P. McIntyre, IEEE Trans. Appl. Supercond., 2023, 33, 1–5 Search PubMed.
  3. Y. Han and A. Zhang, Sci. Rep., 2022, 12, 2349 CrossRef CAS PubMed.
  4. J. Mao, G. Chen and Z. Ren, Nat. Mater., 2021, 20, 454–461 CrossRef CAS PubMed.
  5. J. Mao, H. Zhu, Z. Ding, Z. Liu, A. Gamage Geethal, G. Chen and Z. Ren, Science, 2019, 365, 495–498 CrossRef CAS PubMed.
  6. Y. Qin, B. Qin, D. Wang, C. Chang and L. Zhao, Energy Environ. Sci., 2022, 15, 4527–4541 RSC.
  7. Y. Zheng, X. Han, J. Yang, Y. Jing, X. Chen, Q. Li, T. Zhang, G. Li, H. Zhu, H. Zhao, G. J. Snyder and K. Zhang, Energy Environ. Sci., 2022, 15, 2374–2385 RSC.
  8. H. J. Goldsmid, Thermoelectric Refrigeration, Plenum Press, Wembley, 1964 Search PubMed.
  9. H. J. Goldsmid, in CRC Handbook of Thermoelectrics, ed. D. M. Rowe, CRC Press LLC, Boca Raton, 1995, ch. 8, pp. 73–74 Search PubMed.
  10. K. F. Cuff, R. B. Horst, J. L. Weaver, S. R. Hawkins, C. F. Kooi and G. M. Enslow, Appl. Phys. Lett., 1963, 2, 145–146 CrossRef.
  11. W. M. Yim and A. Amith, Solid-State Electron., 1972, 15, 1141–1165 CrossRef CAS.
  12. B. Madon, J. E. Wegrowe, H. J. Drouhin, X. Liu, J. Furdyna and G. A. Khodaparast, J. Appl. Phys., 2016, 119, 025701 CrossRef.
  13. C. Fu, Y. Sun and C. Felser, APL Mater., 2020, 8, 040913 CrossRef CAS.
  14. X. Li, Z. Zhu and K. Behnia, Adv. Mater., 2021, 33, 2100751 CrossRef CAS.
  15. K. Behnia and H. Aubin, Rep. Prog. Phys., 2016, 79, 046502 CrossRef PubMed.
  16. J. Xiang, S. Hu, M. Lyu, W. Zhu, C. Ma, Z. Chen, F. Steglich, G. Chen and P. Sun, Sci. China: Phys., Mech. Astron., 2019, 63, 237011 Search PubMed.
  17. P. Wang, C. Cho, F. Tang, P. Wang, W. Zhang, M. He, G. Gu, X. Wu, Y. Shao and L. Zhang, Phys. Rev. B, 2021, 103, 045203 CrossRef CAS.
  18. C. Fu, S. N. Guin, T. Scaffidi, Y. Sun, R. Saha, S. J. Watzman, A. K. Srivastava, G. Li, W. Schnelle, S. S. P. Parkin, C. Felser and J. Gooth, Research, 2020, 2020, 4643507 CrossRef CAS PubMed.
  19. Z. Chen, X. Zhang, J. Ren, Z. Zeng, Y. Chen, J. He, L. Chen and Y. Pei, Nat. Commun., 2021, 12, 3837 CrossRef CAS PubMed.
  20. P. Li, P. Qiu, Q. Xu, J. Luo, Y. Xiong, J. Xiao, N. Aryal, Q. Li, L. Chen and X. Shi, Nat. Commun., 2022, 13, 7612 CrossRef CAS PubMed.
  21. Y. Pan, B. He, T. Helm, D. Chen, W. Schnelle and C. Felser, Nat. Commun., 2022, 13, 3909 CrossRef CAS PubMed.
  22. C. Fu, S. N. Guin, S. J. Watzman, G. Li, E. Liu, N. Kumar, V. Süβ, W. Schnelle, G. Auffermann, C. Shekhar, Y. Sun, J. Gooth and C. Felser, Energy Environ. Sci., 2018, 11, 2813–2820 RSC.
  23. W. Liu, Z. Wang, J. Wang, H. Bai, Z. Li, J. Sun, X. Zhou, J. Luo, W. Wang, C. Zhang, J. Wu, Y. Sun, Z. Zhu, Q. Zhang and X. Tang, Adv. Funct. Mater., 2022, 32, 2202143 CrossRef CAS.
  24. T. Feng, P. Wang, Z. Han, L. Zhou, W. Zhang, Q. Liu and W. Liu, Adv. Mater., 2022, 34, 2200931 CrossRef CAS PubMed.
  25. T. Feng, P. Wang, Z. Han, L. Zhou, Z. Wang, W. Zhang, Q. Liu and W. Liu, Energy Environ. Sci., 2023, 16, 1560–1568 RSC.
  26. S. J. Watzman, T. M. McCormick, C. Shekhar, S.-C. Wu, Y. Sun, A. Prakash, C. Felser, N. Trivedi and J. P. Heremans, Phys. Rev. B, 2018, 97, 161404 CrossRef CAS (R).
  27. C. Shekhar, A. K. Nayak, Y. Sun, M. Schmidt, M. Nicklas, I. Leermakers, U. Zeitler, Y. Skourski, J. Wosnitza, Z. Liu, Y. Chen, W. Schnelle, H. Borrmann, Y. Grin, C. Felser and B. Yan, Nat. Phys., 2015, 11, 645–649 Search PubMed.
  28. H. Wang, Z. Zhou, J. Ying, Z. Xiang, R. Wang, A. Wang, Y. Chai, M. He, X. Lu, G. Han, Y. Pan, G. Wang, X. Zhou and X. Chen, Adv. Mater., 2023, 35, 2206941 CrossRef CAS PubMed.
  29. R. T. Delves, Rep. Prog. Phys., 1965, 28, 249–289 CrossRef.
  30. X. Liu, T. Xing, P. Qiu, T. Deng, P. Li, X. Li, X. Li and X. Shi, J. Materiomics, 2023, 9, 345–352 CrossRef.
  31. R. Deng, X. Su, Z. Zheng, W. Liu, Y. Yan, Q. Zhang, V. P. Dravid, C. Uher, M. G. Kanatzidis and X. Tang, Sci. Adv., 2018, 4, eaar5606 CrossRef PubMed.
  32. Y. Wu, P. Qiu, Y. Yu, Y. Xiong, T. Deng, O. Cojocaru-Mirédin, M. Wuttig, X. Shi and L. Chen, J. Materiomics, 2022, 8, 1095–1103 CrossRef.
  33. Q. Xu, C. Ming, T. Xing, P. Qiu, J. Xiao, X. Shi and L. Chen, Mater. Today Phys., 2021, 19, 100417 CrossRef CAS.
  34. Z. Gao, Q. Yang, P. Qiu, T.-R. Wei, S. Yang, J. Xiao, L. Chen and X. Shi, Adv. Energy Mater., 2021, 11, 2100883 CrossRef CAS.
  35. Q. Yang, S. Yang, P. Qiu, L. Peng, T.-R. Wei, Z. Zhang, X. Shi and L. Chen, Science, 2022, 377, 854–858 CrossRef CAS PubMed.
  36. M. Jin, X. Bai, R. Zhang, L. Zhou and R. Li, J. Inorg. Mater., 2021, 37, 101–106 CrossRef.
  37. R. T. Delves, Br. J. Appl. Phys., 1964, 15, 105–106 CrossRef.
  38. L. Guo, Y. K. Liu, G. Y. Gao, Y. Y. Huang, H. Gao, L. Chen, W. Zhao, W. Ren, S. Y. Li, X. G. Li, S. Dong and R. K. Zheng, J. Appl. Phys., 2018, 123, 155103 CrossRef.
  39. F. J. Blatt, P. A. Schroeder, C. L. Foiles and D. Greig, Thermoelectric power of metals, Plenum Press, New York and London, 1976 Search PubMed.
  40. J. Zhou, B. Liao, B. Qiu, S. Huberman, K. Esfarjani, M. S. Dresselhaus and G. Chen, Proc. Natl. Acad. Sci. U. S. A., 2015, 112, 14777–14782 CrossRef CAS PubMed.
  41. M. Pokharel, H. Zhao, K. Lukas, Z. Ren, C. Opeil and B. Mihaila, MRS Commun., 2013, 3, 31–36 CrossRef CAS.
  42. R. Ocaña and P. Esquinazi, Phys. Rev. B, 2002, 66, 064525 CrossRef.
  43. F. Han, N. Andrejevic, T. Nguyen, V. Kozii, Q. T. Nguyen, T. Hogan, Z. Ding, R. Pablo-Pedro, S. Parjan, B. Skinner, A. Alatas, E. Alp, S. Chi, J. Fernandez-Baca, S. Huang, L. Fu and M. Li, Nat. Commun., 2020, 11, 6167 CrossRef CAS PubMed.
  44. J. Liang, P. Qiu, Y. Zhu, H. Huang, Z. Gao, Z. Zhang, X. Shi and L. Chen, Research, 2020, 2020, 6591981 CrossRef CAS PubMed.
  45. J. Gooth, F. Menges, N. Kumar, V. Subeta, C. Shekhar, Y. Sun, U. Drechsler, R. Zierold, C. Felser and B. Gotsmann, Nat. Commun., 2018, 9, 4093 CrossRef CAS PubMed.
  46. A. Jaoui, B. Fauqué, C. W. Rischau, A. Subedi, C. Fu, J. Gooth, N. Kumar, V. Süβ, D. L. Maslov, C. Felser and K. Behnia, npj Quantum Mater., 2018, 3, 64 CrossRef CAS.
  47. N. Kumar, Y. Sun, M. Nicklas, S. J. Watzman, O. Young, I. Leermakers, J. Hornung, J. Klotz, J. Gooth, K. Manna, V. Süβ, S. N. Guin, T. Förster, M. Schmidt, L. Muechler, B. Yan, P. Werner, W. Schnelle, U. Zeitler, J. Wosnitza, S. S. P. Parkin, C. Felser and C. Shekhar, Nat. Commun., 2019, 10, 2475 CrossRef PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ee01450a

This journal is © The Royal Society of Chemistry 2023