LETTERS PUBLISHED ONLINE: 21 APRIL 2013 | DOI: 10.1038/NMAT3621 Extremely high electron mobility in a phonon-glass semimetal S. Ishiwata1 *, Y. Shiomi1 , J. S. Lee1 , M. S. Bahramy2 , T. Suzuki2 , M. Uchida1 , R. Arita1 , Y. Taguchi2 and Y. Tokura1,2 The electron mobility is one of the key parameters that characterize the charge-carrier transport properties of materials, as exemplified by the quantum Hall effect1 as well as high-efficiency thermoelectric and solar energy conversions2,3 . For thermoelectric applications, introduction of chemical disorder is an important strategy for reducing the phonon-mediated thermal conduction, but is usually accompanied by mobility degradation. Here, we show a multilayered semimetal β-CuAgSe overcoming such a trade-off between disorder and mobility. The polycrystalline ingot shows a giant positive magnetoresistance and Shubnikov de Haas oscillations, indicative of a high-mobility small electron pocket derived from the Ag s-electron band. Ni doping, which introduces chemical and lattice disorder, further enhances the electron mobility up to 90,000 cm2 V−1 s−1 at 10 K, leading not only to a larger magnetoresistance but also a better thermoelectric figure of merit. This Ag-based layered semimetal with a glassy lattice is a new type of promising thermoelectric material suitable for chemical engineering. High-carrier-mobility semiconductors and semimetals often exhibit the giant magnetoresistance (GMR) effect, a pronounced increase in resistivity ρxx by the application of a magnetic field B. The classical magnetoresistance associated with the orbital motion of carriers can be described in such a way that the magnetoresistance ratio defined as 1ρxx (B)/ρxx (0) ≡ [ρxx (B) − ρxx (0)]/ρxx (0) is proportional to (µB)2 (µ is electron mobility) at small B, followed by saturation at large B. Although GMR showing a quadratic B dependence has been known since the early twentieth century for semimetallic Bi (ref. 4), particular interest has recently been directed to positive GMR with a linear B dependence in a narrow-gap semiconductor β-Ag2+δ X (X = Se and Te; refs 5,6). The origin of the linear GMR without saturation in a wide temperature range was discussed within the framework of the classical magnetoresistance based on the electron trajectory due to chemical inhomogeneity in addition to the high mobility5,7,8 . As an alternative mechanism, quantum magnetoresistance has been proposed by assuming a zerogap semiconductor with a Dirac-cone-like linearly dispersing band structure, which allows all conduction carriers to condensate at the lowest Landau level9,10 . Although the band-structure calculations suggest that β-Ag2 Se is a narrow-gap semiconductor with parabolic electron and hole bands located at the 0 point11 , the linear dispersion is expected to be realized through the inhomogeneity caused by the slight off-stoichiometry of Ag (refs 9,12). In either the quantum or classical case, chemical inhomogeneity plays an essential role in the GMR in β-Ag2+δ Se. The quantum and classical GMR have recently been discussed also for Dirac fermion systems such as graphene13 , the topological insulator Bi2 Te3 (ref. 14) and narrow-gap semiconductors such as InSb and HgTe (refs 15–17). The multilayered semimetal β-Cu1−x Nix AgSe reported here possesses both high-mobility electrons and lattice disorder as evidenced by the GMR deviating from the quadratic B dependence and the high thermoelectric performance. Whereas classical semiconductors such as doped PbTe have been extensively studied as high-efficiency thermoelectrics18 , semimetals have received less attention, because the thermal broadening of the Fermi distribution function almost equally activates two kinds of carrier with opposite signs. However, the semimetal β-CuAgSe is found to possess a significant electron–hole asymmetry in the band structure near the Fermi level, as revealed by first-principles calculations. β-CuAgSe lies in the middle of the solid solution of β-Ag2 Se and β-Cu2 Se but has a unique layered structure consisting of alternate stacking of the layers of Ag and CuSe (Fig. 1a)19 . Contrary to the naive expectation, the Fermi surface has been found to be nearly isotropic, which allows simple analyses of experimental data for the polycrystalline samples. It should be noted here that β-Ag2 Se is a promising thermoelectric material at room temperature20,21 . The other end compound Cu2 Se also shows a high thermoelectric figure of merit ZT ∼ 1.5 at 1,000 K, where the system forms a cubic lattice with statistically distributed copper ions, and becomes superionic22 . Likewise, CuAgSe forms a cubic lattice with superionic conductivity in the α phase above 470 K (ref. 19). As shown in Fig. 1b, β-Cu1−x Nix AgSe shows metallic behaviour down to the lowest temperature. The residual resistivity ratio ρxx (300 K)/ρxx (2 K) is 25 for the undoped sample (x = 0) and 17 for the Ni-doped sample (x = 0.1). The negative sign of the Seebeck coefficient S indicates the dominance of electron-type conduction (Fig. 1c). However, the sign of S for the undoped sample changes from negative to positive at 400 K (see the inset), indicative of a semimetallic band structure, as will be discussed later. Note that the jump in ρxx as well as S at 470 K corresponds to the anomaly of the phase transition to the superionic phase (α-CuAgSe). Lattice thermal conductivity (κlattice = κ − κelectron ) as estimated from the Wiedemann–Franz law is as low as 10 mW cm−1 K−1 for both compounds (Fig. 1d). Such a small κlattice implies the presence of a large lattice anharmonicity due to disorder, which is reminiscent of the amorphous solids23 . Interestingly, the 10% Ni doping decreases ρxx by a factor of 3–4, while reducing κlattice and maintaining S at an almost comparable, or even larger value at low temperatures. As a result, the Ni doping increases ZT from 0.15 to 0.25 at room temperature, and from 0.02 to 0.1 at 100 K (Fig. 1e). Figure 2a,b shows a positive GMR in β-Cu1−x Nix AgSe with x = 0 and 0.1. With decreasing temperature, 1ρxx (B)/ρxx (0) shows a 1 Department of Applied Physics and Quantum-Phase Electronics Center (QPEC), University of Tokyo, Hongo, Tokyo 113-8656, Japan, 2 RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan. *e-mail: ishiwata@ap.t.u-tokyo.ac.jp. NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials © 2013 Macmillan Publishers Limited. All rights reserved. 1 NATURE MATERIALS DOI: 10.1038/NMAT3621 LETTERS a Se Ag Cu 10¬3 ρxx (Ω cm) 4 Cu1¬xNixAgSe x=0 10¬5 0 100 S (μV K¬1) ¬50 6 4 0 2 ¬100 8 6 10¬3 300 x = 0.1 x=0 400 500 Temperature (K) ¬100 κ (mW cm¬1 K¬1) d 50 ρxx (Ω cm) 4 2 c x = 0.1 2 10¬4 S (μV K¬1) b x=0 x = 0.1 40 30 20 x=0 10 0 e 0.2 ZT x = 0.1 0.1 0.0 0 x=0 100 200 Temperature (K) 300 Figure 1 | Structure and transport properties. a, Crystal structure of β-CuAgSe with a pseudo-tetragonal unit cell (shown by dotted lines). The occupancy of the Cu site is 0.5. b–e, Resistivity ρxx (b), Seebeck coefficient S (c), thermal conductivity κ (d) and dimensionless figure of merit ZT (e) as a function of temperature for β-Cu1−x Nix AgSe. The blue shaded areas in d correspond to the estimated lattice thermal conductivity on the assumption of the Wiedemann–Franz law. The inset shows the Seebeck coefficient and resistivity above room temperature for β-CuAgSe below 470 K and α-CuAgSe above 470 K. significant increase especially below 100 K and reaches 24 at 2 K under a transverse B of 14 T (B⊥ current) and that for the Ni-doped sample exceeds 140 under the same conditions. The magnitudes of magnetoresistance at low temperatures are 2–14 times larger than that of β-Ag2+δ Se (ref. 5). As longitudinal magnetoresistances (shown by dashed lines) are rather small, the spin-polarization effect can be ruled out as a major origin of the transverse GMR. In Fig. 2c,d, the negative slope of the Hall resistivity ρyx at low fields reconfirms the predominance of the electron conduction below 300 K. However, nonlinear field dependence of ρyx can be found for β-Cu1−x Nix AgSe with both x = 0 and 0.1 at low temperatures, signalling two-carrier (electron and hole) conduction. For x = 0, the sign of the slope ρyx /B changes from negative to positive 2 with increasing B at 50 and 2 K. In the high-B limit, ρyx /B is approximately equal to (e(nh − ne ))−1 , where e and nh (ne ) denote the elementary charge and the carrier concentration of holes (electrons), respectively. Thus, the concentration of hole carrier is expected to be larger than that of the electron for x = 0 and vice versa for x = 0.1. The presence of the two kinds of carrier with different τ can be confirmed for x = 0 by the deviation of 1ρxx (B)/ρxx (0) from a scaling function f (B/ρxx (0)) (Fig. 2e). This scaling relation, called Kohler’s rule, holds for the system with a constant scattering time τ over the Fermi surfaces. Therefore, considering the negative slope of ρyx at low B and its highly nonlinear B dependence, it is reasonable to assume that β-CuAgSe has two kinds of carrier with different τ , one is the high-mobility electron carrier and the other is the hole carrier. As for the Ni-doped sample (Fig. 2f), the magnetoresistance data at low temperatures tend to fall on a single scaling curve with a power factor of 1.3, almost obeying Kohler’s rule. This result implies the further predominance of the electron conduction in the Ni-doped sample. The conductivity tensors are analysed by the two-carrier model as shown in Fig. 3a,b. All of the longitudinal and Hall conductivities are fitted with the following formulae24 . neµ + Cxx (1) σxx = 1 + (µB)2 1 σxy = nH eµ2H B + Cxy (2) 1 + (µH B)2 Here, n(nH ) and µ(µH ) denote the carrier concentration and mobility for the high-mobility carriers estimated from the longitudinal conductivity, σxx (Hall conductivity σxy ), respectively. Cxx (Cxy ) denotes the low-mobility component estimated from σxx (σxy ). The fitting parameters are summarized in Supplementary Tables SI and SII. We assume that the high-mobility electron carriers dominate the positive magnetoresistance effect, and that the fitting parameters, which are dependent on temperature, are independent of B. Using the fitting results both for σxx (data not shown) and σxy , we independently extract the longitudinal and Hall mobilities, µ and µH (Fig. 3c), and carrier concentrations, n and nH (see Supplementary Information), respectively. µ and µH coincide perfectly with each other in a wide temperature range, indicating the validity of our fitting procedure based on the two-carrier model. The carrier concentrations at 10 K, n = 7.4 × 1018 cm−3 and nH = 6.5 × 1018 cm−3 , are also comparable to each other, signalling the semimetallic nature of the compound. On decreasing the temperature to 20 K, the mobilities for both compounds increase in proportion to T −1 (dashed lines in Fig. 3c). Thus, the carrier scattering is dominated not by the defect scattering but by the electron–phonon scattering in a wide temperature range down to low temperatures. Notably, the mobilities at 10 K are about 20,000 cm2 V−1 s−1 and 90,000 cm2 V−1 s−1 for β-Cu1−x Nix AgSe with x = 0 and 0.1, respectively, the latter of which is comparable to the high-quality thin film of doped HgTe (ref. 15) suitable for studying the quantum Hall effect. Although the deviation of the magnetoresistance curves from the quadratic B dependence indicates the effect of inhomogeneity, the main origin of the positive GMR in this system can be ascribed to the extremely high carrier mobility. As seen in Fig. 3d,e, the ratios of low-mobility hole conductivity to high-mobility electron conductivity, Cxx /σxx and Cxy , are fairly small and decrease either by doping Ni or decreasing temperature, in accordance with the enhancement of magnetoresistance. Despite the random orientation of domains in the polycrystalline β-CuAgSe, Shubnikov de Haas (SdH) oscillations are observed for ρyx at low temperatures (Fig. 4a), reflecting the high carrier mobility. The oscillatory behaviour cannot be discerned for ρxx because of the GMR masking the SdH signals. The periodicity of the NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials © 2013 Macmillan Publishers Limited. All rights reserved. NATURE MATERIALS DOI: 10.1038/NMAT3621 a LETTERS b CuAgSe Cu0.9Ni0.1AgSe B⊥I 8 100 K 2 50 K B || I 2K 4 2 ρxx (Ω cm) ρxx (Ω cm) 4 8 6 200 K 6 10¬4 10¬3 300 K 10¬3 B⊥I 300 K 200 K 100 K 10¬4 6 4 2K 2 6 4 50 K 2K B || I 2K 10¬5 6 ¬10 c ¬5 0 5 Magnetic field (T) ¬10 10 d 2 ¬5 0 5 Magnetic field (T) 10 1.5 1.0 1 0 ρyx (mΩ cm) ρyx (mΩ cm) 0.5 2K 50 K 0.0 2K 50 K ¬0.5 ¬1 ¬2 100 K 200 K 300 K ¬10 ¬5 0 5 Magnetic field (T) ¬1.5 10 e f 0.1 300 K 0.01 1014 1016 ¬5 0 5 Magnetic field (T) 2K 20 K 50 K 100 K 200 K 300 K 1 0.1 0.001 1018 B2/ρxx(0)2 (Oe2 Ω¬2 cm¬2) 10 100 0.01 H2 H 1.3 0.001 300 K ¬10 10 2K 20 K 50 K 100 K 200 K Δ ρ (B)/ρ xx(0) Δ ρ (B)/ρxx(0) 10 1 100 K 200 K ¬1.0 H2 H 1.3 1014 1016 1018 B2/ρ xx(0)2 (Oe2 Ω¬2 cm¬2) 1020 Figure 2 | GMR and Hall resistance. a–f, Magnetic field dependence of resistivity ρxx (a,b) and Hall resistivity ρyx (c,d), and magnetoresistance 1ρ/ρ0 plotted as a function of B2 /ρ02 (e,f) for β-CuAgSe and β-Cu0.9 Ni0.1 AgSe, respectively (ρ0 ≡ ρxx (B = 0)). The resistivity under a transverse magnetic field (B perpendicular to the current I) is shown by solid lines and the resistivity under a longitudinal magnetic field (BkI) is shown by the dashed lines in a,b. SdH oscillation minima 1(1/B), 0.0076 T−1 , gives a Fermi surface cross-section area of 0.013 Å−2 (Fig. 4a, inset). Assuming a spherical Fermi surface, the carrier concentration at 2 K was estimated to be 8.6 × 1018 cm−3 . As this value is close to that estimated by the two-carrier analyses, one can ascribe the origin of the SdH oscillation not to an impurity phase such as a silver metal but to the intrinsic bulk nature of β-CuAgSe. The temperature dependence of the SdH oscillation amplitude at 11 T is fitted to the standard Lifshitz–Kosevich theory adopted for the Hall resistivity25 , 1ρyx 2π2 kB T /h̄ωc = exp(−2π2 kB TD /h̄ωc ) 4ρyx sinh(2π2 kB T /h̄ωc ) (3) where TD and ωc denote the Dingle temperature and the cyclotron frequency, respectively. The fitting gives the rather small cyclotron mass mc (= eB/ωc ) of 0.13 m0 (m0 being the mass of free electron). To further characterize the conduction carriers in β-CuAgSe, we measured the reflectivity spectra at room temperature. Figure 4b shows the plasmon loss function given by −Im[1/] = 2 /(12 + 22 ). From the peak position at 0.06√eV, corresponding to the reduced plasma frequency ωp∗ (= ωp / ∞ ), the effective mass m∗ is estimated, using the values n = 7.0 × 1018 cm−3 and ∞ = 21, to be 0.12 m0 , which is comparable to the cyclotron mass. Reflecting the semimetallic nature with low carrier density, the optical conductivity spectrum (Fig. 4b, inset) shows a small NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials © 2013 Macmillan Publishers Limited. All rights reserved. 3 NATURE MATERIALS DOI: 10.1038/NMAT3621 LETTERS a c 6 20 K μ (cm2 V¬1 s¬1) 10 30 K 5 50 K 0 0.0 0.5 40 2.0 Cu0.9Ni0.1AgSe 104 6 CuAgSe 100 200 Temperature (K) 2 CuAgSe 10¬3 6 Cu0.9Ni0.1AgSe 4 0 20 K e 100 200 Temperature (K) 10¬4 Cxy 300 K 10¬5 Cu0.9Ni0.1AgSe 100 K 0.5 1.0 1.5 Magnetic field (T) 10¬6 2.0 300 CuAgSe 20 0 0.0 300 d Cu0.9Ni0.1AgSe 50 K μH 2 0 b 10 K μ μH 2 300 K 1.0 1.5 Magnetic field (T) μ 4 4 100 K σ xy (× 105 Ω¬1 cm¬1) 105 CuAgSe Cxx /σ xx σ xy (× 105 Ω¬1 cm¬1) 10 K 0 100 200 Temperature (K) 300 Figure 3 | Two-carrier model analyses for the conductivity tensors. a,b, Magnetic field dependence of Hall conductivity σxy for β-CuAgSe (a) and β-Cu0.9 Ni0.1 AgSe (b). c–e, Temperature dependence of drift mobility µ and Hall mobility µH of high-mobility electron carriers (c), Cxx /σxx (d) and Cxy (e) (see text), obtained from the fits for conductivity tensors. Δ ρyx (μΩ cm) a 0.5 Δ ρyx /ρyx (%) 0.4 2 4 c 2K 18 K 4 3 0 ¬2 0.3 2 0.08 0.09 0.10 0.11 1/B (1/T) 0.2 1 0.1 ¬Im [ε (ω )] (× 10¬3 m F¬1) 60 50 40 30 10 15 Temperature (K) σ (ω) (Ω¬1 cm¬1) b 5 20 25 1,200 ¬1 ¬2 800 400 0 0.0 E ¬ E F (eV) 0 0.0 0 ¬3 0.5 1.0 Photon energy (eV) ¬4 20 ¬5 10 0 0.0 0.1 0.2 Photon energy (eV) 0.3 ¬6 Γ X M Γ Z Figure 4 | Experimental and theoretical characterization of conduction electrons. a,b, Temperature dependence of Shubnikov de Haas (SdH) oscillation amplitude 1ρyx /ρyx (%) at 11 T (a) and plasmon loss function as a function of the photon energy (b) for β-CuAgSe. The inset in a shows the oscillatory part of ρyx , and the inset of b shows optical conductivity at room temperature. c, Calculated electronic band structure of β-CuAgSe (see text). Drude weight and a pseudogap structure at around 0.1 eV, above which interband transitions, for instance from Te 4p to Ag 5s bands, are dominant. 4 Having the experimental evidence for the semimetallic band structure with the light-mass electron carriers, we have performed band-structure calculations for β-CuAgSe as shown in Fig. 4c. NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials © 2013 Macmillan Publishers Limited. All rights reserved. NATURE MATERIALS DOI: 10.1038/NMAT3621 Along the 0–M and 0–X lines, a highly dispersive Ag 5s band is crossing the Fermi level, whereas a less dispersive band mainly consisting of a Te 4p band and a Cu 3d band is located just below the Fermi level. This feature is consistent with the two-carrier analyses on the conductivity tensors and the sign change in S from negative to positive on increasing temperature. Despite the layered structure, the highly dispersive band is also discerned along the 0−Z line. The isotropic Fermi surface topology, that is, electron pocket, is compatible with the observation of the SdH oscillation in the polycrystalline sample. The averaged mass of the Ag s band is about 0.4m0 , which is small but larger than the experimental values. This discrepancy may arise from the inaccuracy of the structure model used for the band calculation; for example, the fivefold superlattice modulation in the c plane was confirmed by transmission electron microscope observations26 . As the band-edge structure is sensitive to external perturbations, the superlattice modulation, the chemical disorder in β-CuAgSe and the chemical pressure by Ni doping would cause non-trivial effects on the carrier conduction in this semimetallic system. Let us summarize the important features of β-CuAgSe. First, it shows a positive GMR, neither linear nor quadratic to B, and the magnitude of the magnetoresistance increases markedly as the mobility increases at low temperatures. This is in contrast to β-Ag2 Se showing positive and linear GMR without such a remarkable temperature dependence5 . Second, it has a semimetallic band structure with a small pocket of extremely high-mobility Ag s electrons. The high mobility and relatively low carrier concentration are the key factors for the large thermoelectric power factor. Third, the high-mobility electron conduction is rather robust against the chemical disorder and the randomly oriented domains, as evidenced by the enhanced electron mobility in the Ni-doped sample and by the observation of SdH oscillation in the polycrystalline β-CuAgSe. Moreover, β-CuAgSe may have substantial atomic-lattice disorder due to the proximity to the superionic phase27,28 . For the origin of the compatibility of the high electron mobility and the significant chemical disorder, we propose two possibilities. One is the fairly large screening length due to the large polarizability of Cu and Ag ions, which is advantageous also for the superionic conduction. The other is the effective band-edge engineering by the chemical inhomogeneity9,12,17 . Although our band calculation suggests that the highly dispersive conduction band and the less dispersive valence band are not very close to each other, non-parabolic and gapless (Dirac-cone-like) energy dispersion with a strong electron–hole asymmetry may be realized by the inhomogeneity. In addition, the multilayered structure is suitable for band tuning by chemical substitution. Thus, the multilayered semimetal β-CuAgSe provides a fruitful playground not only for new magnetoresistance phenomena but for a promising thermoelectric performance near or below room temperature, where the strong electron–hole imbalance near the Fermi level remains intact. We regard this material as a candidate for a new type of phonon-glass electron crystal29,30 , characterized by high-mobility Ag s electrons in the presence of atomic and macroscopic inhomogeneity, by which lattice thermal conductivity is reduced effectively. Methods The polycrystalline ingots were prepared by a solid-state reaction in evacuated quartz tubes (∼5 × 10−3 Pa). A mixture of Cu, Ni, Ag and Se powders at a molar ratio 1−x: x: 1 : 1.01 (x = 0 and 0.1) was slowly heated to 673 K at a rate of 20 K h−1 and then heated to 1,073 K at a rate of 50 K h−1 and maintained at this temperature for 24 h, followed by cooling to room temperature at a rate of 100 K h−1 . The obtained sample was ground, pressed into pellets and subjected to the same heat treatment process again. The transport measurements were performed on bar-shaped rectangular samples with dimensions of about 2×0.5×0.2 mm3 . Air-dried gold paste was used as the electrodes. The Seebeck coefficient and thermal conductivity were measured by a conventional steady-state method. The reflectivity spectrum was obtained in LETTERS the energy range between 0.01 and 6 eV with coated aluminium and silver mirrors as references. After extrapolation of the reflectivity in unmeasured spectral regions by adopting the Hagen–Rubens relation for the low-energy region, the optical conductivity was derived through the Kramers–Kronig transformation. The electronic structure calculations were carried out using the full-potential augmented plane-wave plus local orbital method, as implemented in the WIEN2K code31 . The exchange-correlation part of the potential was treated using the modified Becke–Johnson functional32 . The muffin-tin radii RMT of Ag, Cu and Se were set to 2.5, 2.04 and 1.81 Bohr, respectively, and the maximum modulus of reciprocal vectors Kmax was chosen such that RMT Kmax = 7. The lattice parameters and atomic positions were taken from experiment26 and the corresponding Brillouin zone was sampled by a 11 × 11 × 7 k-mesh. In our calculations, the unit cell contains two formula units of CuAgSe with two Cu atoms occupying (3/4, 1/4, 0.895) and (1/4, 3/4, 0.105) sites, two Se atoms occupying (3/4, 1/4, 0.127) and (1/4, 3/4, 0.873) sites, and finally two Ag atoms located at (3/4, 3/4, 0.551) and (1/4, 1/4, 0.449). Received 20 October 2012; accepted 5 March 2013; published online 21 April 2013 References 1. Novoselov, K. S. et al. Two-dimensional gas of massless Dirac fermions in graphene. Nature 438, 197–200 (2005). 2. Mahan, G., Sales, B. & Sharp, J. Thermoelectric materials: New approaches to an old problem. Phys. Today 50, 42–47 (March, 1997). 3. Shah, A. V. et al. Thin-film silicon solar cell technology. Prog. Photovolt. 12, 113–142 (2004). 4. Kapitza, P. L. Bismuth crystals and its change in strong the study of the specific resistance of magnetic fields and some allied problems. Proc. R. Soc. Lond. A 119, 358–443 (1928). 5. Xu, R. et al. Large magnetoresistance in non-magnetic silver chalcogenides. Nature 390, 57–60 (1997). 6. Husmann, A. et al. Megagauss sensors. Nature 417, 421–424 (2002). 7. Parish, M. M. & Littlewood, P. B. Classical magnetotransport of inhomogeneous conductors. Phys. Rev. B 72, 094417 (2005). 8. Herring, C. Effect of random inhomogeneities on electrical and galvanomagnetic measurements. J. Appl. Phys. 31, 1939–1953 (1960). 9. Abrikosov, A. A. Quantum linear magnetoresistance. Phys. Rev. B 58, 2788–2794 (1998). 10. Lee, M., Rosenbaum, T. F., Saboungi, M-L. & Schnyders, H. S. Band-gap tuning and linear magnetoresistance in the silver chalcogenides. Phys. Rev. Lett. 88, 066602 (2002). 11. Fang, C. M., de Groot, R. A. & Wiegers, G. A. Ab initio band structure calculations of the low-temperature phases of Ag2 Se. J. Phys. Chem. Solids 63, 457–464 (2002). 12. Wang, X-L, Dou, S. X. & Zhang, C. Zero-gap materials for future spintronics, electronics and optics. NPG Asia Mater. 2, 31–38 (2010). 13. Friedman, A. L. et al. Quantum linear magnetoresistance in multilayer epitaxial graphene. Nano Lett. 10, 3962–3965 (2010). 14. Wang, X., Du, Y., Dou, S. & Zhang, C. Room temperature giant and linear magnetoresistance in topological insulator Bi2 Te3 nanosheets. Phys. Rev. Lett. 108, 266806 (2012). 15. Thio, T. et al. Giant magnetoresistance in zero-band-gap Hg1−x Cdx Te. Phys. Rev. B 57, 12239–12244 (1998). 16. Solin, S. A., Thio, T., Hines, D. R. & Heremans, J. J. Enhanced room-temperature geometric magnetoresistance in inhomogeneous narrow-gap semiconductors. Science 289, 1530–1532 (2000). 17. Hu, J. & Rosenbaum, T. F. Classical and quantum routes to linear magnetoresistance. Nature Mater. 7, 697–700 (2008). 18. Heremans, J. P. et al. Enhancement of thermoelectric efficiency in PbTe by distortion of the electronic density of states. Science 321, 554–557 (2008). 19. Miyatani, S. Electronic and ionic conduction in (Ag1−x Cux )2 Se. J. Phys. Soc. Jpn 34, 423–432 (1973). 20. Conn, J. B. & Taylor, R. C. Thermoelectric and crystallographic properties of Ag2 Se. J. Electrochem. Soc. 107, 977–982 (1960). 21. Ferhat, M. & Nagao, J. Thermoelectric and transport properties of Ag2 Se compounds. J. Appl. Phys. 88, 813–816 (2000). 22. Liu, H. et al. Copper ion liquid-like thermoelectrics. Nature Mater. 11, 422–425 (2012). 23. De Yoreo, J. J., Knaak, W., Meissner, M. & Pohl, R. O. Low-temperature properties of crystalline (KBr)1−x (KCN)x : A model glass. Phys. Rev. B 34, 8828–8842 (1986). 24. Takahashi, H., Okazaki, R., Yasui, Y. & Terasaki, I. Low-temperature magnetotransport of the narrow-gap semiconductor FeSb2 . Phys. Rev. B 84, 205215 (2011). 25. Shoenberg, D. Magnetic Oscillations in Metals (Cambridge Univ. Press, 1984). 26. Frueh, A. J., Czamanke, G. K. & Knight, C. H. The crystallography of eucairite, CuAgSe. Z. Kristallogr. 108, 389–396 (1957). NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials © 2013 Macmillan Publishers Limited. All rights reserved. 5 NATURE MATERIALS DOI: 10.1038/NMAT3621 LETTERS 27. Ramasesha, S. A phenomenological model for semiconductor–metal transition in mixed conductors. J. Solid State Chem. 41, 333–337 (1982). 28. Fortner, J., Saboungi, M-L. & Enderby, J. E. Carrier density enhancement in semiconducting NaSn and CsPb. Phys. Rev. Lett. 74, 1415–1418 (1995). 29. Slack, G. A. in CRC Handbook of Thermoelectrics (ed. Rowe, M.) 407–440 (CRC, 1995). 30. Snyder, G. J. & Toberer, E. S. Complex thermoelectric materials. Nature Mater. 7, 105–114 (2008). 31. Blaha, P., Schwarz, K., Madsen, G., Kvasnicka, D. & Luitz, J. WIEN2K package, Version 10.1, http://www.wien2k.at. 32. Tran, F. & Blaha, P. Accurate band gaps of semiconductors and insulators with a semilocal exchange-correlation potential. Phys. Rev. Lett. 102, 226401 (2009). Acknowledgements The authors thank D. Okuyama and T. Arima for experimental support and thank J. G. Checkelsky, A. Tsukazaki, F. Kagawa and N. Kanazawa for useful comments. This study was in part supported by a Grant-in-Aid for Scientific Research (Grant No. 6 23685014) from the MEXT, and by the Funding Program for World-Leading Innovative R&D on Science and Technology (FIRST Program), Japan. Author contributions S.I. and Y. Tokura conceived the study and wrote the paper. S.I. prepared the samples and performed the transport measurements. Y.S. and M.U. designed the thermoelectric measurement systems. T.S. and Y. Taguchi performed thermoelectric measurements at high temperatures. J.S.L. worked on the optical study. M.S.B. and R.A. performed band calculations. Additional information Supplementary information is available in the online version of the paper. Reprints and permissions information is available online at www.nature.com/reprints. Correspondence and requests for materials should be addressed to I.S. Competing financial interests The authors declare no competing financial interests. NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials © 2013 Macmillan Publishers Limited. All rights reserved.