This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

Large magnetoresistance and quantum oscillations in Sn0.05Pb0.95Te

, , , and

Published 28 June 2021 © 2021 IOP Publishing Ltd
, , Citation K Shrestha et al 2021 J. Phys.: Condens. Matter 33 335501 DOI 10.1088/1361-648X/ac06ed

0953-8984/33/33/335501

Abstract

We have synthesized high-quality single crystals of SnxPb1−xTe and carried out detailed studies of the magnetotransport properties of one of the samples, Sn0.05Pb0.95Te. Longitudinal magnetoresistance increases almost linearly with increasing applied field (H) and reaches ∼310% at H = 13 T. At higher fields, both longitudinal and Hall resistance show clear Shubnikov de Haas oscillations. The oscillations are smooth and periodic, and there exists only one frequency, fα ∼ 57 T. However, an additional frequency, fβ∼ 69 T, appears as the angle between the field direction and the normal to the sample surface (θ) is increased. Both fα and fβ exhibit θ-dependence; fα decreases whereas fβ increases gradually with increasing θ. The presence of two frequencies in Sn0.05Pb0.95Te indicates that there exist two Fermi surface pockets (α and β pockets). We have constructed the Landau-level fan plot and determined the Berry phase (δ) for the α pocket to be δ ∼ 0.1. This δ value is very close to the expected value of 0 for a topologically trivial system.

Export citation and abstract BibTeX RIS

1. Introduction

Topological phases of materials have attracted enormous attention in recent years because of the novel properties of their surface states [14]. A topological insulator (TI) has an insulating bulk and highly conducting surface states arising from the non-trivial topology of the bulk states. TIs that have been predicted are experimentally verified using various techniques such as angle-resolved photoemission spectroscopy (ARPES) [58], quantum oscillations [911], weak antilocalization (WAL) [1216], etc. ARPES is a commonly used technique for studying topological materials as it can directly

visualize the Dirac dispersion of surface states. Magnetoresistance (MR)/magnetization under high magnetic fields show quantum oscillations known as Shubnikov de Haas (SdH)/de Haas–van Alphen oscillations [1719]. Surface electrons are Dirac fermions and have a linear dispersion in the band structure. The two-dimensional Fermi surface of surface states can be identified by carrying out measurements of the angle dependence of quantum oscillations and Berry phase calculations [3, 16, 20]. The origin of surface states in TIs is both fascinating and pivotal for understanding fundamental physics, and surface states have potential applications in modern electronic and computer devices. However, only a few TIs are known to exist in the real world, including bismuth- and thallium-based TIs (also known as Z2-TIs) [20, 21]. This is due to the requirement of strong spin–orbit interaction and time-reversal symmetry for Z2-TIs. Also, most of the Z2-TIs studied thus far do not have an insulating bulk due to crystal defects (antisite or point). The bulk conduction channel interferes with the surface conduction channel, which makes the transport characterization of surface states very challenging, although not impossible [22, 23].

In the search for other topological compounds, a new system, known as a topological crystalline insulator (TCI), has been discovered [24]. A TCI has a simple rock salt crystal structure and includes elements from groups IV–VI, namely Snx Pb1−x Se/Te [2528]. Like the Z2-TIs, these materials possess semiconducting bulk states and metallic gapless surface states. However, there are some fundamental differences between the surface states in the two types of TIs. First, surface states in TCIs are protected by the mirror symmetry or reflection symmetry present in the crystal, whereas those in Z2-TIs are protected by the time-reversal symmetry [29, 30]. Second, there is an even number of Dirac cones on the surface terminations in TCIs, and they are oriented perpendicular to the mirror symmetry planes, whereas Z2-TIs possess an odd number of Dirac cones [21, 31].

Recent studies have shown that the topological properties of Snx Pb1−x Se/Te can be tuned either via doping [32, 33] or external pressure [34], making it an ideal system for investigating the topological phase transition. With increasing Sn content, Snx Pb1−x Se/Te shows a band inversion at critical doping of xc ∼ 0.35 for Snx Pb1−x Te [27] and ∼0.17 for Snx Pb1−x Se [35]. This means that the system transforms from a topologically trivial to non-trivial phase at xc. The existence of surface states and topological phase transition in TCIs has been evidenced using methods such as ARPES [25, 27, 36], magneto-optical absorption [37], WAL [3840], etc. However, there have been limited magnetotransport studies [4145] of TCIs, especially of their quantum oscillations, although this is one of the most powerful techniques for studying topological surface states. One of the reasons could be the higher magnetic field range needed to observe the Landau-level quantization, typically above 9 T, which is not easily attainable in most labs. By analyzing the quantum oscillations, we can determine the Fermi surface topology in addition to its shape, size, and dimensionality. Therefore, it would be interesting to explore how the Fermi surface topology changes with x in Snx Pb1−x Se/Te.

Here, we have synthesized high-quality single crystals of Snx Pb1−x Te, and we present the results of detailed magnetotransport studies of one of the samples, Sn0.05Pb0.95Te. The sample shows a large positive MR without showing any sign of saturation. At higher fields, it exhibits clear SdH oscillations. We have analyzed these oscillations and we discuss the Fermi surface topology.

2. Experimental details

Single crystals of Snx Pb1−x Te were grown by the chemical vapor transport method. All of the elements used in the crystal growth, lead, tin, and tellurium, were of high purity (99.9999%). First, high-purity powder samples of PbTe (62 at.% Pb and 39 at.% Te) and SnTe (48 at.% Sn and 52 at.% Te) were prepared by solid-state reaction. Powder mixtures with nominal concentrations estimated by the Pb/Sn mixture weight ratio were each heated in a clean evacuated quartz tube maintained at 900°C over 48 h prior to crystal growth. Afterward, each powder mixture was sealed in an evacuated quartz tube of dimensions 2.2 × 2 × 40 cm3 along with a transporting agent of ∼150 mg iodine (I2). The quartz tube was placed in a horizontal 2 zone furnace maintained for 10 d at 900°C–800°C, before being further cooled down to room temperature.

Longitudinal and Hall resistance under applied magnetic fields up to 13 T were measured in an Oxford cryostat. A special homemade probe was built for rotating the sample in the magnetic field. A linear resistance bridge (LR-700) and a Lake Shore (LS336) temperature controller were used for measuring the resistance and temperature, respectively, of the sample.

3. Results and discussion

Figures 1(a)–(c) shows the x-ray diffraction (XRD) analysis of Sn-doped PbTe single-crystal powder along with the calculated pattern at selected Sn doping (x). It is well known that PbTe crystallizes in the cubic structure; when Sn is doped at the Pb site of PbTe, the lattice parameter decreases systematically. The figure 1(d) inset shows the lattice parameter (a) determined from the Rietveld refinement at different x values. A systematic decrease of the lattice parameter with x indicates that a periodic doping was successfully implemented. The temperature dependence of the longitudinal resistance (Rxx ) for Snx Pb1−x Te (x = 0.05) is displayed in figure 1(d). The resistance decreases with decreasing temperature, showing typical metallic behavior.

Figure 1.

Figure 1.  XRD and electrical resistance. ((a)–(c)) Room-temperature powder XRD pattern (black squares), together with the Rietveld refinement (red line) and Bragg positions (vertical lines) for Snx Pb1−x Te single crystals at selected doping levels (x). XRD patterns are consistent with cubic crystal structure. (d) Temperature dependence of electrical resistance for Sn0.05Pb0.95Te. Resistance decreases with decreasing temperature, indicating metallic behavior. Inset: decreasing lattice parameter as a function of x. The dashed line is a guide for the eye.

Standard image High-resolution image

To investigate the magnetotransport properties of this system, we have measured MR in one of the samples, Sn0.05Pb0.95Te. The tilt angle (θ) is defined as the angle between H and the normal to the [001] surface. Figure 2 shows the longitudinal (Rxx ) and Hall resistance (Rxy ) at θ = 0°. MR is expressed as a percentage, i.e. MR (%) = $\frac{{R}_{xx}\left(H\right)-{R}_{xx}\left(0\right)}{{R}_{xx}\left(0\right)}\enspace {\times}$ 100%, where Rxx (0) and Rxx (H) are the resistance values at 0 and H applied fields, respectively. As shown in figure 2(a), MR is positive and increases almost linearly with H. It reaches 310% at 13 T without showing any sign of saturation. Such unsaturated large positive MR in Snx Pb1−x Te has also been reported in previous magnetotransport studies [41, 44, 46]. Roychowdhury et al [46] observed a systematic decrease in the MR value with increasing x. Here, we have estimated MR ∼ 210% at H = 9 T for x = 0.05, which is slightly higher than the previously reported MR ∼ 190% for x = 0.4. Therefore, our MR data lies within the same doping dependence trend reported by Roychowdhury et al [46]. The Hall resistance, Rxy , also increases linearly with H, as shown in figure 2(b). The positive slope of the Rxy vs H plot confirms the presence of the hole-like bulk charge carriers in Sn0.05Pb0.95Te. It should be noted that both longitudinal and Hall resistance show a signature of SdH oscillations at higher fields, and these oscillations can be clearly resolved in the polynomial background-subtracted data, as shown in the insets to figures 2(a) and (b), respectively. The oscillations are periodic and well-defined, and appear to have a single frequency. The frequency of the oscillations can be determined by obtaining the fast Fourier transform (FFT).

Figure 2.

Figure 2.  MR and quantum oscillations. (a) MR expressed as a percentage and (b) Hall resistance (Rxy ) of single-crystal Sn0.05Pb0.95Te as functions of applied field (H). MR is positive and reaches ∼310% under an applied field of 13 T at 4.2 K. The positive slope of the Rxy vs H plot indicates the presence of hole-like bulk charge carriers. Both MR and Rxy show a signature of quantum oscillations at higher fields that is clearly observed in the polynomial background-subtracted data, as shown in the insets to (a) and (b), respectively. Landau levels (N) in the quantum oscillations are labeled as marked by the arrows. The oscillations are periodic and appear to have a single frequency.

Standard image High-resolution image

Figure 3 shows the frequencies of the SdH oscillations at θ = 0°. Both the longitudinal and transverse oscillations have a single frequency located at fα = 57 T. As the frequency of SdH oscillations is directly proportional to the cross-section of the Fermi surface, the single frequency indicates the presence of one Fermi surface in Sn0.05Pb0.95Te. Using the Onsager's relation [1719, 47], we can determine the Fermi-wave vector as

Equation (1)

where = $\frac{h}{2\pi }$, h is Planck's constant, and KF is the Fermi wave vector. Using equation (1), we have estimated that KF = 4.12 × 106 cm−1.

Figure 3.

Figure 3.  Frequency analyses of SdH oscillations. FFT of SdH oscillations at θ = 0°. Both the longitudinal and Hall components have a single frequency located at fα = 57 T. Inset: illustration of the tilt angle, θ, between the applied field and the normal to the [001] surface.

Standard image High-resolution image

In order to map the Fermi surface, we have measured the angle dependence of the SdH oscillations. Figures 4(a) and (b) show the SdH oscillations after polynomial background subtraction for longitudinal and Hall resistance, respectively, at different θ values. The oscillations at θ = 0° appear smooth and periodic, and they exhibit a single frequency of 57 T (see figure 3). At higher θ values, the oscillations in the longitudinal resistance (figure 4(a)) are similar to those at θ = 0°, but they are aperiodic, possibly due to the superposition of more than one frequency. Similar aperiodic oscillations are also seen in the Hall component (figure 4(b)). The superposition of multiple frequencies in the oscillations can be separated by obtaining the Fourier transform.

Figure 4.

Figure 4.  Quantum oscillations at higher angles. (a) MR and (b) Hall resistance at different θ values obtained after polynomial background subtraction. The oscillations at θ = 0° are periodic and have a single frequency. At higher θ, the SdH oscillations seem to have a mixture of more than one frequency. The data were obtained at T = 4.2 K and the plots are shifted vertically for clarity.

Standard image High-resolution image

Figure 5 shows the frequency plots at different θ values for both longitudinal and Hall oscillations. The frequency values for the longitudinal and Hall components are comparable to one another. At θ = 0°, there exists a single frequency, fα = 57 T, and it decreases to 45 T as θ is increased to 15° (figure 5(b)). It should be noted that an additional frequency, fβ = 69 T, appears at θ = 15°. A third frequency of 118 T is seen at θ = 45° (figure 4(d)) and is nearly the sum of fα and fβ , i.e. (fα + fβ ). Although there exists a single frequency at θ = 0°, we have observed two distinct frequencies (fα and fβ ) at higher θ values. The observation of two frequencies suggests that there exist two distinct Fermi wave vectors representing two different sections (α and β pockets) of the Fermi surface.

Figure 5.

Figure 5.  Frequency plot at higher angles. Fourier transforms of quantum oscillations for both longitudinal and Hall resistance at θ values of (a) 0°, (b) 15°, (c) 30°, and (d) 45°. The frequency values obtained from oscillations in longitudinal and transverse resistance are comparable to one another. At θ = 0°, there exists only one frequency. The number of frequencies increases at higher angles, suggesting the existence of more than one Fermi wave vector.

Standard image High-resolution image

Previous magnetotransport studies [41, 44] have also reported two distinct frequencies in PbTe, although the frequency values that were found are slightly different than those observed here. Ma et al [41] reported two frequencies, one ≈ 39 T and another that appears as a broad shoulder ≈70 T. Furthermore, Akiba et al [44] found two frequencies at 22.6 T and 45.2 T, although they claimed that the second frequency is the second harmonic of the first. Since the frequency of the SdH oscillations in PbTe depends on the bulk carrier concentration [48], the slight difference in the frequency values in our Sn0.05Pb0.95Te sample could be due to a different bulk carrier concentration arising from the small amount of Sn doping. However, detailed studies at higher fields are required to determine unambiguously both the number of frequencies and their values.

The angle dependence of fα and fβ is shown in the inset to figure 6(a). With increasing θ, fα decreases gradually, whereas fβ first increases and then decreases slightly. In order to obtain information on the Fermi surface topology, we have calculated the Berry phase (δ). We chose the oscillation data at θ = 0° for this calculation to avoid interference from the second frequency. Figure 6(a) shows the SdH oscillations in conductance, Δσxx , which were obtained after subtracting the polynomial background. Here, σxx = $\frac{{\rho }_{xx}}{\left({\rho }_{xx}^{2}+{\rho }_{xy}^{2}\right)}$, where ρxx and ρxy represent the longitudinal and Hall resistivity, respectively. Minima and maxima positions of the oscillations were assigned as integer (N) and half-integer (N + $\frac{1}{2}$) values, respectively, and the constructed Landau-level fan diagram is shown in figure 6(b). From the linear extrapolation of the data at the limit of $\frac{1}{H}\enspace \to $ 0, we have obtained δ = 0.11 ± 0.07 and fα = (56.8 ± 0.6) T. This value of δ is very close to 0, which is expected for a topologically trivial system [911, 49]. The frequency obtained from the linear extrapolation, 56.8 T, is also consistent with the fα = 57 T obtained by the FFT of the SdH oscillations (figure 3). It should be noted that our Sn0.05Pb0.95Te sample has Sn content, x = 0.05, that is below xc = 0.35. Therefore, the trivial topology of the α pocket determined from the analyses of SdH oscillations is consistent with the results from other measurement techniques [27, 28, 33, 38]. The topology of the α pocket is well understood based on the current data. However, it will be necessary to acquire more SdH oscillations data above the current maximum field of 13 T to calculate δ for, and determine the topology of, the β pocket.

Figure 6.

Figure 6.  Berry phase analyses. (a) Quantum oscillations after background subtraction at θ = 0°. Inset: angle dependence of frequencies fα and fβ from figure 5. The error bar for each data point is defined as the half-width at half-height of the respective peak in the frequency plot (figure (5)). (b) Landau-level fan diagram constructed using maxima and minima of the oscillations in (a). From the linear extrapolation, the Berry phase and frequency values are found to be δ = 0.1 and fα = 56.8 T, respectively.

Standard image High-resolution image

4. Summary

We have synthesized high-quality single crystals of Snx Pb1−x Te, and studied the Fermi surface topology of one of the samples, Sn0.05Pb0.95Te. The sample exhibits a large positive MR that increases almost linearly with increasing applied field and does not show any sign of saturation. At higher fields, SdH oscillations appear for both longitudinal and Hall resistance. At θ = 0°, the SdH oscillations are smooth and periodic, and they consist of a single frequency, fα = 57 T. At higher θ values, the SdH oscillations are complex and consist of two frequencies, fα and fβ . The presence of two frequencies suggests that two Fermi surface pockets (α and β pockets) are present in the Sn0.05Pb0.95Te sample. For the α pocket, we have estimated the Berry phase (δ) to be ∼0.1, which is very close to the expected value of 0 for a topologically trivial system. In order to construct the Landau-level plot for the β pocket and hence determine its topology, studies will need to be conducted above the current maximum field of 13 T. Although our motivation is to explore the topological phase transition in Snx Pb1−x Te, this work only presents the Fermi surface topology of the sample below the critical doping level (xc ∼ 0.35). Detailed magnetotransport studies at higher fields and of samples with higher Sn content (particularly above xc) are underway, and the obtained results will be reported elsewhere.

Acknowledgments

We are grateful to Dr. F C Chou for providing high-quality single crystals of Snx Pb1−x Te. This work is supported in part by the US Air Force Office of Scientific Research Grant FA9550-15-1-0236, the TLL Temple Foundation, the John J and Rebecca Moores Endowment, and the State of Texas through the Texas Center for Superconductivity at the University of Houston. The research at West Texas A&M University is supported by a start-up Grant from the Paul Engler College of Agriculture and Natural Sciences and by the Welch Foundation (Grant No. AE-025).

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

Please wait… references are loading.
10.1088/1361-648X/ac06ed