Transport properties of topological nodal-line semimetal candidate CaAs3 under hydrostatic pressure
Li Jing1, 2, Zhao Ling-Xiao1, Wang Yi-Yan1, Wang Xin-Min1, 2, Ma Chao-Yang1, Zhu Wen-Liang1, 2, Gao Mo-Ran1, 2, Zhang Shuai1, Ren Zhi-An1, 2, 3, Chen Gen-Fu1, 2, 3, †
Institute of Physics and Beijing National Laboratory for Condensed Matter Physics, Chinese Academy of Sciences, Beijing 100190, China
School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
SongShan Lake Materials Laboratory, Dongguan 523808, China

 

† Corresponding author. E-mail: gfchen@iphy.ac.cn

Abstract

We report the transport properties of the CaAs3 single crystal, which has been predicted to be a candidate for topological nodal-line semimetals. At ambient pressure, CaAs3 exhibits semiconducting behavior with a small gap, while in some crystals containing tiny defects or impurities, a large “hump” in the resistivity is observed around 230 K. By applying hydrostatic pressure, the samples appear to a tendency towards metallic behavior, but not fully metallized up to 2 GPa. Further high pressure studies are needed to explore the topological characteristics for CaAs3.

1. Introduction

The discovery of topological semimetals (TSMs) has attracted broad attention in condensed matter physics and material science because of the topologically protected band structure. TSMs give rise to some exotic physical properties, including giant magnetoresistance (MR), ultrahigh carrier mobility, nontrivial Berry phase, and negative longitudinal MR.[14] According to the degeneracy of the band crossing points and their distribution in the Brillouin zone (BZ), TSMs can classify into Dirac semimetals (DSMs), Weyl semimetals (WSMs), and nodal-line semimetals (NLSMs).[5] The band crossing points are four-fold degenerate for DSMs, such as Na3Bi and Cd3As2.[68] WSMs exhibit two-fold degenerate band crossing points with definite chirality, and it is located at an even number of discrete points in the BZ, such as TaAs[9,10] and WTe2.[11,12] Different from the discrete points in DSMs and WSMs, the band crossing points form closed loops in the BZ of NLSMs.[13,14]

Recently, it has been reported that black phosphorus (BP) exhibits a topological phase transition from band semiconductor to zero-gap Dirac semimetal state[15,16] under hydrostatic pressure. The semimetal phase in BP is characterized by the large MR, the coexistence of electron and hole pockets, low carrier densities, small cyclotron masses, weak localization-weak antilocalization transition at low magnetic fields and the emergence of a nontrivial Berry phase of π detected by Shubnikov-de Haas (SdH) magneto-oscillations. BP has a puckered honeycomb lattice,[1719] in which the direct band gap is mainly determined by the out-of-plane pz-like orbital of P and can be further reduced by tuning the interlayer coupling.[15,20,21] Interestingly, in CaAs3, the two-dimensional infinite polyanionic nets are derivatives of the BP structure. While a deeper and systematic study on CaAs3 single crystals is still missing except for the semiconducting behavior.[22,23] More recently, CaAs3 was predicted theoretically to be a candidate for topological insulator (TI) or NLSMs when the spin–orbit coupling (SOC) was taken into consideration or not and the topological transition could be engineered via strain, or chemical doping.[5,24] While it should be degenerate to a strong topological insulator (TI) with taking the SOC into consideration.[5,24] Hydrostatic pressure has been recognized as a powerful tool for studying band structures of semiconductors. Therefore, combined with the theoretical prediction, it is necessary to study the pressure effect on the transport properties of CaAs3.

In this work, we have grown the single crystals of CaAs3 and performed the studies of transport properties at atmospheric and hydrostatic pressures. CaAs3 shows a semiconducting-like behavior at atmospheric pressure. A large hump in the resistivity is also observed for some crystals containing tiny defects or impurities. Hydrostatic pressure can effectively reduce its resistivity, but it is not fully metallized up to 2 GPa. However, a small drop of resistivity is observed above 1.5 GPa for some samples, indicating the onset of superconducting transition.

2. Experiment and method

Single crystals of CaAs3 were grown by the modified Bridgman method. Firstly, stoichiometric amounts of Ca and As granule were put into an Al2O3 crucible, and sealed in an evacuated quartz ampule, which was then heated extremely slowly and kept at 900 °C for 10 h before turning the furnace off. Secondly, the precursor was grinded, tableted, put into an Al2O3 crucible, and then sealed again as described in the first step. The quartz ampule was heated to 800 °C, kept at this temperature for 50 h, and then cooled to 400 °C slowly. The phase of obtained single crystals was characterized by x-ray diffraction (XRD) on a PANalytical diffractometer with Cu radiation at room temperature and atmospheric pressure. The elemental compositions were checked by Oxford X-Max energy dispersive x-ray (EDX) spectroscopy analysis in a Hitachi S-4800 scanning electron microscope. Electrical resistivity was measured by the conventional dc four-probe method and the measurements were performed on a Quantum Design physical properties measurement system (PPMS-9). Samples were pressurized in a piston cylinder device made of Be–Cu alloy. The pressure transmitting medium was Daphene Oil 7373. The pressure inside the device was determined by shifting the superconducting transition temperature of Pb strip.

3. Results and discussion

In Fig. 1(a), the powder XRD pattern of CaAs3 can be well indexed to the triclinic structure with the lattice parameters a = 5.88(1) Å, b = 5.85(1) Å, and c = 5.93(1) Å, α = 70.05(1)°, β = 80.16(1)°, and γ = 75.83(1)°, which are in agreement with the previous reported values.[25] figure 1(b) shows the single crystal XRD pattern of the CaAs3. All of peaks can be identified as the (0k0) reflections. For comparison, we also show the crystal structure of BP in Fig. 1(c). BP is a two-dimensional (2D) layered material which is stacked together by weak van der Waals force. Each P atom is covalently bonded with three adjacent P atoms to form a puckered honeycomb structure in a single layer. One can clearly see that the crystal structure of CaAs3 is similar to that of BP. Here CaAs3 can be viewed as a list of 2D infinite puckered polyanionic layers stacking along the b-axis and Ca2+ cations are inserted into the interstitial sites, in which the puckered arsenic layers are directly derived from the orthorhombic structure of BP by removing 1/4 of the P atoms in an ordered way.[23]

Fig. 1. (a) Powder XRD pattern of CaAs3 obtained by grinding the single crystals at room temperature and the refinement results. (b) The single crystal XRD of CaAs3. (c) and (d) The crystal structure of BP and CaAs3, respectively.

All the samples we measured were taken from different parts of a large crystal of CaAs3. We find that the transport properties show a large sample dependence. The crystals peeled from the bottom exhibit semiconducting-like behavior, while some crystals taken from the top show “bad metal” behavior with a big hump at the temperature T1, as shown in Figs. 2(a) and 2(b), respectively. It is noted that the absolute value of resistivity is essentially same between samples #1 and #2 at 300 K. In Fig. 2(a), the ρ(T) in the range from 195 K to 265 K can be fitted by Arrhenius equation, , where kB is the Boltzmann constant. The estimated activation energy Eg is ∼73 meV, which is consistent to the previous report.[22]

Fig. 2. Resistivity ρ plotted as a function of temperature T measured on #1 crystal (a) and #2 (b) at atmospheric pressure, respectively.

To further distinguish between the two type resistivity behaviors in the single crystals (#1 and #2) of CaAs3, we performed the Hall-effect measurements at various temperatures on the same samples, as shown in Figs. 3(a) and 3(d), respectively. The magnetic field is applied perpendicular to the ac-plane. Figure 3(a) presents the magnetic field dependence of the Hall resistivity of the crystal (#1) at several temperatures from 2 K to 300 K. The slope of the curve is all negative, suggesting that the main carriers should be electron-type. One can find that the shows nonlinear behavior at 2 K, indicating the coexistence of two types of carriers at very low temperatures.[8,9] Different from sample #1, the slope of for sample #2 changes from negative to positive with decreasing temperatures, as shown in Fig. 3(d), indicating the transition from electrons dominated transport to holes dominated transport in this sample. All these show that CaAs3 is a multiband system. We try to fit the Hall data obtained at 2 K for sample #1 using the semi-classical two-band model, in which Hall conductivity can be expressed as the following equation:[26] where nh (or ne) is the carrier density of the holes (or the electrons) and μh (or μe) is the mobility of the holes (or the electrons), respectively. For high temperatures, the appears to be approximately linear, and fitting with two-band model could bring uncertainty. Hence, we use the single-band model to estimate the carrier density , and carrier mobility (0 T), where t = e, h indicate electron-type and hole-type carriers, respectively, and the Hall coefficient RH is estimated by the slope of (B) at low fields ( T). By fitting with the single-band and two-band model for different temperatures, the carrier densities and mobilities for sample #1 are summarized in Figs. 3(b) and 3(c), respectively. A large discrepancy is found between the results obtained from the two methods. At 2 K, the concentration and mobility for two kinds of carriers have little difference ( , , , and ). However, the results of fitting with single-band and two-band model are not consistent between 2 K and other temperatures. Above 20 K, μe increases monotonically with increasing temperature, and up to around at 300 K, while ne shows a little temperature dependence from 4.2 ×1017 cm−3 at 20 K to 9.6 ×1017 cm−3 at 300 K. The temperature dependence of the carrier densities and mobilities for sample #2 are shown in Figs. 3(e) and 3(f), respectively. We find that the carrier density is larger than that of sample #1 by 1 order of magnitude. The carriers change from hole-type to electron-type as the temperature goes up to around 230 K, and the transition point is in accord with T1 in the ρ(T) curve. Remarkably, the obvious hump in the ρ(T) curve of the #2 crystal could be attributed to dominated carriers change. We noted that the resistivity curve of HfTe5 also showed a huge hump at about 65 K, at which the dominated carrier type changed, and which was ascribed to the occurrence of the critical topological phase transition.[27] However, in this special case of layered crystal structure, the anomalous transport properties observed in some CaAs3 could be reasonably attributed to the presence of impurities or crystal defects during the crystal growth.

Fig. 3. (a) and (d) The magnetic field dependence of Hall resistivity on samples #1 and #2 at various temperatures, respectively. (b), (c), (e), and (f) The temperature dependence of carrier mobilities and carrier densities of electrons and holes measured on samples #1 and #2, respectively.

CaAs3 was predicted to hold engineered topological phase transitions by applying external tensile or compressive strains.[24] In order to test the pressure effect of CaAs3, we use the piston cylinder device to apply compressive strains to two samples #3 and #4 which are selected with two typical transport behaviors. Figure 4(a) and 4(b) display the temperature dependence of the resistivity ρ(T) under varying hydrostatic pressures for samples #3 and #4, respectively. The values of ρ(T) both for samples #3 and #4 decrease in its entirety under pressure. At low temperatures, the ρ(T) both for two samples changes more significantly, but not metallized up to 2 GPa. In Fig. 4(b), the peak temperature, T1, also shifts to lower temperatures with increasing pressure, while we find that ρ(T) of sample #4 drops with the lowering temperature above 1.5 GPa as shown in the inset. The transition moves to higher temperature with the increasing pressure. It seems natural to ascribe this drop to the onset of superconductivity in such “doped” sample, which provides the possibility for CaAs3 to realize a topological nodal-line semimetal or even topological superconductivity under much higher pressure in the future work.

Fig. 4. (a) and (b) Resistivity ρ of CaAs3 plotted as a function of temperature T on #3 and #4 crystals at various hydrostatic pressures, respectively. Inset of (b): Resistivity ρ above 1.2 GPa in the low temperature region (2 K–6 K).
4. Conclusion

In summary, we have synthesized the single crystals of CaAs3, which has been predicted to be a topological nodal-line semimetal candidate. CaAs3 shows an n-type semiconducting-like behavior at atmospheric pressure. While for some samples with the presence of crystal defect and/or impurities, the resistivity shows an obvious hump around 230 K, at which the dominated carriers change from hole-type to electron-type. The application of hydrostatic pressure results in decrease of the resistivity for both samples, but they are not fully metallized on the whole. A possible superconducting transition is also observed in some “doped” samples above 1.5 GPa. Our work will provide valuable clues for further exploration of the topological phase in CaAs3, and in-depth studies are needed in future.

Reference
[1] He J B Chen D Zhu W L Zhang S Zhao L X Ren Z A Chen G F 2017 Phys. Rev. 95 195165
[2] Weng H M Dai X Fang Z 2016 J. Phys.: Condens. Matter 28 303001
[3] Bansil A Lin H Das T 2016 Rev. Mod. Phys. 88 021004
[4] Chiu C K Teo J C Y Schnyder A P Ryu S 2016 Rev. Mod. Phys. 88 035005
[5] Xu Q N Yu R Fang Z Dai X Weng H M 2017 Phys. Rev. 95 045136
[6] Xiong J Kushwaha S K Liang T Krizan J W Hirschberger M Wang W Cava R J Ong N P 2015 Science 350 413
[7] Liu Z K Zhou B Zhang Y Wang Z J Weng H M Prabhakaran D Mo S K Shen Z X Fang Z Dai X Hussain Z Chen Y L 2014 Science 343 864
[8] Liang T Gibson Q Ali M N Liu M H Cava R J Ong N P 2015 Nat. Mater. 14 280
[9] Huang X C Zhao L X Long Y J Wang P P Chen D Yang Z H Liang H Xue M Q Weng H M Fang Z Dai X Chen G F 2015 Phys. Rev. 5 031023
[10] Zhang C L Xu S Y Belopolski I Yuan Z J Lin Z Q Tong B B Bian G Alidoust N Lee C C Huang S M Chang T R Chang G Q Hsu C H Jeng H T Neupane M Sanchez D S Zheng H Wang J F Lin H Zhang C Lu H Z Shen S Q Neupert T Hasan M Z Jia S 2016 Nat. Commun. 7 10735
[11] Soluyanov A A Gresch D Wang Z J Wu Q S Troyer M Dai X Bernevig B A 2015 Nature 527 495
[12] Cai P L Hu J He L P Pan J Hong X C Zhang Z Zhang J Wei J Mao Z Q Li S Y 2015 Phys. Rev. Lett. 115 057202
[13] Burkov A A Hook M D Balents L 2011 Phys. Rev. 84 235126
[14] Weng H M Liang Y Y Xu Q N Yu R Fang Z Dai X Kawazoe Y 2015 Phys. Rev. 92 045108
[15] Xiang Z J Ye G J Shang C Lei B Wang N Z Yang K S Liu D Y Meng F B Luo X G Zou L J Sun Z Zhang Y Chen X H 2015 Phys. Rev. Lett. 115 186403
[16] Li C H Long Y J Zhao L X Shan L Ren Z A Zhao J Z Weng H M Dai X Fang Z Ren C Chen G F 2017 Phys. Rev. 95 125417
[17] Rodin A S Carvalho A Castro Neto A H 2014 Phys. Rev. Lett. 112 176801
[18] Li D A Jussila H Karvonen L Ye G J Lipsanen H Chen X H Sun Z P 2015 Sci. Rep. 5 15899
[19] Morita A 1986 Appl. Phys. 39 227
[20] Takao Y Asahina H Morita A 1981 J. Phys. Soc. Jpn. 50 3362
[21] Asahina H Shindo K Morita A 1982 J. Phys. Soc. Jpn. 51 1193
[22] Bauhofer W Wittmann M Gmelin E van Schnering H G 1982 Physics of Narrow Gap Semiconductors Berlin Heidelberg Springer 30 10.1007/3-540-11191-3_3
[23] Bauhofer W Wittmann M Vonschnering H G 1981 J. Phys. Chem. Sol. 42 687
[24] Quan Y Yin Z P Pickett W E 2017 Phys. Rev. Lett. 118 176402
[25] Brice J F Courtois A Protas J Aubry J 1976 J. Sol. St. Chem. 17 393
[26] Ashcroft N W Mermin N D 1976 Solid Sate Physics New York Holt, Rinehart and Winston
[27] Zhao L X Huang X C Long Y J Chen D Liang H Yang Z H Xue M Q Ren Z A Weng H M Fang Z Dai X Chen G F 2017 Chin. Phys. Lett. 34 037102