Optical study of charge dynamics in CaCo2As2
Zhang Wei1, 2, Xu Bing2, Yang Run2, Liu Jin-Yun2, Yang Hao3, †, , Qiu Xiang-Gang2, ‡,
College of Physics, Optoelectronics and Energy & Collaborative Innovation Center of Suzhou Nano Science and Technology, Soochow university, Suzhou 215006, China
Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China
College of Science, Nanjing University of Aeronautics and Astronautics, Nanjing 211106, China

 

† Corresponding author. E-mail: yanghao@nuaa.edu.cn

‡ Corresponding author. E-mail: xgqiu@iphy.ac.cn

Project supported by the National Basic Research Program of China (Grant Nos. 2012CB821400, 2012CB921302, and 2015CB921303) and the National Natural Science Foundation of China (Grants Nos. 11274237, 91121004, 51228201, and 11004238). Wei Zhang also thanks the support of the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD).

Abstract
Abstract

We present an infrared spectroscopy study of charge dynamics in CaCo2As2 single crystal. In this material, the optical conductivity can be described by two Drude components with different scattering rates (1/τ): a broad incoherent background and a narrow Drude component. By monitoring the temperature dependence, we find that only the narrow Drude component is temperature-dependent and determines the transport properties. Especially a Fermi liquid behavior of carriers is revealed by the T2 behavior in the dc resistivity ρn and scattering rate 1/τn, indicating a coherent nature of quasiparticles in the narrow Drude subsystem.

1. Introduction

Chemical substitution, or doping, is the most common way to tune the properties of a correlated material. In iron-based superconductors (FeSCs), superconductivity is induced by various doping processes into metallic parent compounds. For example, in BaFe2As2 (122 system), doping with either electrons[1] or holes[2] suppresses the magnetic order, showing superconductivity with a maximum Tc close to the edges of the magnetic dome. The resistivity with different doping levels shows a temperature dependence from T linear to T2, revealing a crossover from non-Fermi liquid to Fermi liquid.[35]

FeSCs are Fe 3d6 multi-orbital systems. All the Fe 3d bands (or orbitals) exist near the Fermi level, resulting in complicated hole- and electron-like Fermi surface sheets,[6,7] and each Fe 3d band turns out to have different fillings and bandwidths.[8] In these multi-band systems the electronic correlations are mainly caused by the Hund rule coupling instead of the strong intraband Coulomb interactions as in cuprates.[9,10] Up to now, FeSCs were thought to be in the intermediate correlation regime. Nevertheless, many features, which are usually observed in some strongly correlated materials, were also found in FeSCs,[11,12] such as the linear temperature-dependent resistivity,[3,4] characteristic of a non-Fermi liquid. The non-Fermi-liquid behavior or linear temperature dependence of the resistivity is often taken as the hallmark of a quantum critical regime.[13,14]

It is well known that the electronic correlations in FeSCs are strongly band-dependent. The coexistence of electrons with different electronic correlations (coherent part and incoherent part) could be reflected in the charge excitation spectra which can be measured by the optical spectroscopy technique. In a multi-band environment, the optical conductivity is a sum of the contribution over the bands, and it happens to be parallel of resistances. Indeed, this is the case found by many previous optical studies on FeSCs,[1519] in which the optical conductivity can be characterized by two Drude terms with different scattering rates; and the total Drude weight is a sum of terms ∼(n/m*)α, where n is a carrier concentration and m* is an effective mass, and α is the band index. Thus, in the situation of heavy differentiation of correlation strength, the transport properties are highly influenced by the bands of the weakly correlated ones (or more coherent ones).

Optical spectroscopy is a bulk-sensitive and energy-resolved probe, and it turns out to be a powerful tool to study charge dynamics of the multi-band system. In this paper, we report a detailed optical study on high quality CaCo2As2 single crystals, a counterpart of the 122 parent compound CaFe2As2. We found that the optical conductivity of CaCo2As2 can be decomposed into a narrow Drude (coherent) component and a broad incoherent background. Upon cooling, only the narrow Drude component is temperature-dependent and determines the transport properties. This subsystem reveals a T2 behavior in the dc resistivity and scattering rate disclosing a Fermi-liquid behavior in this material.

2. Experiments

High-quality CaCo2As2 single crystals were grown with the flux method.[20] The ab-plane reflectivity R(ω) was measured at a near-normal angle of incidence on a Bruker VERTEX 80v FTIR spectrometer. An in situ gold overfilling technique[21] was used to obtain the absolute reflectivity R(ω) of the sample. Data from 50 to 15000 cm−1 were collected at 13 different temperatures from 300 K to 10 K on a freshly cleaved surface on an ARS–Helitran crysostat. Since a Kramers–Kronig analysis requires a broad spectral range, R(ω) was extended to the UV range (up to 50000 cm−1) at room temperature with an AvaSpec-2048×14 fiber optic spectrometer.

3. Results and discussion

Figure 1 shows the reflectivity spectra R(ω) for CaCo2As2 in the far-infrared region at several temperatures. R(ω) shows a typical metallic response, having relatively high value and approaching unity at low frequencies. Two sharp features at ∼ 180 cm−1 and ∼ 260 cm−1 are associated with the symmetry-allowed ab plane infrared-active phonon modes,[2225] which will not be discussed in this paper. The inset displays the room-temperature R(ω) up to 15000 cm−1 for CaCo2As2. R(ω) shows a plasma edge at ∼ 6000 cm−1. The plasma edge is larger than the value ∼ 2000 cm−1 in the typical 122 parent compound BaFe2As2 (red curve). The large value of the plasma edge generally suggests an increase of carrier density, consistent with the expanded electron Fermi pockets with electron doping in 122 system compounds.[7,26]

Fig. 1. Temperature-dependent reflectivity of CaCo2As2 in the far infrared region. Inset: 300-K reflectivity over a wide frequency range for CaCo2As2 and BaFe2As2.

The real part of the optical conductivity σ1(ω), which provides more direct information about the charge dynamics, has been determined via a Kramers–Kronig analysis of the reflectivity spectra R(ω).[27] Below the lowest measured frequency, we used a Hagen–Rubens form for the low-frequency extrapolation. Above the highest measured frequency, we assume a constant reflectivity up to 12.5 eV, followed by a free-electron (ω−4) response.

Figure 2 displays the optical conductivity σ1(ω) at several temperatures between 10 and 300 K in the frequency up to 4000 cm−1. The low-frequency σ1(ω) has a Drude-like peak centered at zero frequency, indicating the metallic nature of the material. This is consistent with the reflectivity analysis. The width of the Drude peak at half maximum represents the quasiparticle scattering rate, while the area under the Drude peak (spectral weight) is proportional to the free carrier density. As the temperature is reduced, this Drude-like peak narrows with a concomitant increase of the low-frequency optical conductivity. The inset of Fig. 2 shows the 300 K conductivity over a broader spectral range for CaCo2As2 and BaFe2As2. In contrast to the small Drude peak in BaFe2As2, the CaCo2As2 has a Drude component with significantly larger spectral weight, indicating that the carrier density increases in CaCo2As2, which is also consistent with the reflectivity analysis. In addition to the Drude-like peak, the optical conductivity of CaCo2As2 shows a rather flat region with a long tail extending to ∼4000 cm−1 and followed by a temperature-independent mid-infrared peak around 8000 cm−1 indicative of strong interband transitions.

Fig. 2. The real part of the optical conductivity at different temperatures for CaCo2As2 up to 4000 cm−1. Inset: 300-K optical conductivity over a wide frequency range for CaCo2As2 and BaFe2As2.

The optical conductivity is conveniently parameterized by a Drude–Lorentz model

where Z0 is the vacuum impedance. The first term in Eq. (1) corresponds to the Drude response characterized by a plasma frequency Ωp,j and a scattering rate 1/τj. The second term is a sum of Lorentz oscillators, each having a resonance frequency Ωk, a line width γk, and a weight Sk. The Drude term accounts for the free carrier response, while the Lorentz contributions represent bound excitations.

In the previous works on pnictides,[1519] the optical conductivity σ1(ω) can be modeled reasonably by the superposition of two Drude components and series of Lorentz terms over a wide frequency range. As shown in Fig. 3(a), we employed a similar fit for CaCo2As2 at 150 K. It is noted that the low frequency σ1(ω) is characterized by a coherent narrow Drude (blue line) with a small scattering rate (1/τn ∼ 314 cm−1) and an incoherent broad Drude (magenta line) with a large scattering rate (1/τb ∼ 3474 cm−1), as well as two Lorentzian oscillators at higher frequencies corresponds to the interband excitation in the mid-infrared region. The fitting parameters at 150 K are summarized in Table 1.

Table 1.

The Drude–Lorentz fitting results of the optical conductivity at 150 K, where Ωp,j and τj are the plasma frequency and scattering rate for the jth Drude component (Dj), and Ωk, Sk, and γ are the oscillator frequency, strength, and line width of the kth Lorentz mode (Lk). All units are in cm−1.

.
Fig. 3. (a) The measured σ1(ω) at 150 K (thick black line) in a wide spectral range for CaCo2As2, fitting with the Drude–Lorentz model (thin red line). The optical conductivity of CaCo2As2 can be separated into temperature-dependent and -independent parts. The temperature-independent part consists from the incoherent broad Drude term σb (magenta line), and two Lorentzian oscillators at higher frequencies (orange and dark yellow lines). The sum of the three contributions results in the light gray shaded area. The unshaded area below the curves is temperature dependent and can be described by a single Drude term σn (blue line). Temperature evolution of σn for CaCo2As2 is shown in panel (b). The thin lines through the data are fits with a single Drude term.

The temperature dependence of the two Drude components provides information about the nature of the two different types of carriers in this material. As shown in Fig. 2, for all data at different temperatures, the long flat tail of σ1(ω) changes very slightly with temperature, meaning that it just contributes a constant background on the optical conductivity. The temperature-independent part of σ1(ω) is indicated by the light-gray shaded area in Fig. 3(a), and it consists of the broad Drude component σb (magenta hatched area) and two oscillators that associate with the high-frequency contributions (orange and dark-yellow hatched area). On the top of the constant background, the remaining temperature-dependent contribution is the narrow Drude component σn(ω)

Upon cooling, the scattering rate 1/τn becomes narrowing, while the plasma frequency Ωp,n is roughly a constant, indicating that there seems to be no considerable transfer of spectral weight between both Drude subsystems. As shown in Table 1, the coherent narrow component σn contains only about 1/4 of the spectral weight compared to the incoherent background σb, but it turns out that it is of superior importance to low-frequency properties and transport properties.

The two Drude terms’ contributions to the total dc resistivity can be considered as a parallel circuit

In order to reveal the temperature dependence of the subsystem described by σn, we simply subtract the temperature-independent part σb by the Drude–Lorentz fits. As shown by the thin lines through the corresponding data in Fig. 3(b), the temperature-dependent part of σ1(ω) can be described well with the single narrow Drude component σn(ω).

Figures 4(a) and 4(b) display the temperature dependence of dc resistivity ρn(T) = 1/σn(ω → 0) and scattering rate 1/τn. Across all the measured temperature range, ρn and 1/τn present a T2-dependent behavior. This T2 temperature dependence is expected for electron–electron scattering that is supposed to be dominant for correlated electron systems at low temperatures and subject to Landau’s theory of a Fermi liquid. The Fermi liquid behavior revealed by the T2 dependence of the scattering rate and dc resistivity in CaCo2As2 indicates a coherent nature of quasiparticles in the narrow Drude subsystem.

Fig. 4. The temperature dependence of the fitting parameters for the Narrow Drude term in CaCo2As2: (a) the reverse dc conductivity ρ = 1/σ1(ω → 0) and (b) the scattering rate 1/τ, respectively. The red lines in panels (a) and (b) are the T2 fits.
4. Conclusion

In summary, we have measured the electrodynamics properties of CaCo2As2 single crystal in a wide frequency range as a function of temperature. The optical conductivity of CaCo2As2 can be decomposed into a broad temperature-independent background and a narrow Drude-type component which solely determines the transport properties. The narrow Drude subsystem reveals a T2 behavior in the dc resistivity and scattering rate up to room temperature, indicating a Fermi-liquid behavior in this material.

Reference
1Ni NTillman M EYan J QKracher AHannahs S TBud’ko S LCanfield P C 2008 Phys. Rev. 78 214515
2Rotter MTegel MJohrendt D 2008 Phys. Rev. Lett. 101 107006
3Doiron-Leyraud NAuban-Senzier PRené de Cotret SBourbonnais CJérome DBechgaard KTaillefer L 2009 Phys. Rev. 80 214531
4Kasahara SShibauchi THashimoto KIkada KTokeya HHirata KTerashima TMatsuda Y 2010 Phys. Rev. 81 184519
5Dai Y MMiao HXing L YWang X CWang P SXiao HQian TRichard PQiu X GYu WJin C QWang ZJohnson P DHomes C CDing H 2015 Phys. Rev. 5 031035
6Ding HRichard PNakayama KSugawara KArakane TSekiba YTakayama ASouma SSato TTakahashi TWang ZDai XFang ZChen G FLuo J LWang N L 2008 Europhys. Lett. 83 47001
7Malaeb WYoshida TFujimori AKubota MOno KKihou KShirage P MKito HIyo AEisaki HNakajima YTamegai TArita R 2009 JPSJ 78 123706
8Lu D HYi MMo S KErickson A SAnalytis JChu J HSingh D JHussain ZGeballe T HFisher I RShen Z X 2008 Nature 455 81
9Yin Z PHaule KKotliar G 2011 Nat. Mater. 10 932
10Misawa TNakamura KImada M 2012 Phys. Rev. Lett. 108 177007
11Mazin I I 2010 Nature 464 183
12Johnston D C 2010 Adv. Phys. 59 803
13Hashimoto KCho KShibauchi TKasahara SMizukami YKatsumata RTsuruhara YTerashima TIkeda HTanatar M AKitano HSalovich NGiannetta R WWalmsley PCarrington AProzorov RMatsuda Y 2012 Science 336 1554
14Walmsley PPutzke CMalone LGuillamón IVignolles DProust CBadoux SColdea A IWatson M DKasahara SMizukami YShibauchi TMatsuda YCarrington A 2013 Phys. Rev. Lett. 110 257002
15Wu DBarisic NKallina PFaridian AGorshunov BDrichko NLi L JLin XCao G HXu Z AWang N LDressel M 2010 Phys. Rev. 81 100512
16Tu J JLi JLiu WPunnoose AGong YRen Y HLi L JCao G HXu Z AHomes C C 2010 Phys. Rev. 82 174509
17Nakajima MIshida SKihou KTomioka YIto TYoshida YLee C HKito HIyo AEisaki HKojima K MUchida S 2010 Phys. Rev. 81 104528
18Dai Y MXu BShen BXiao HWen H HQiu X GHomes C CLobo R P S M 2013 Phys. Rev. Lett. 111 117001
19Yang Y XXiong RFang Z HXu BXiao HQiu X GShi JWang K 2014 Chin. Phys. 23 107401
20Zhang WNadeem KXiao HYang RXu BYang HQiu X G 2015 Phys. Rev. 92 144416
21Homes C CReedyk MCradles D ATimusk T 1993 Appl. Opt. 32 2976
22Akrap ATu J JLi L JCao G HXu Z AHomes C C 2009 Phys. Rev. 80 180502
23Schafgans A APursley B CLaForge A DSefat A SMandrus DBasov D N 2011 Phys. Rev. 84 052501
24Dai Y MXu BShen BXiao HLobo R P S MQiu X G 2012 Chin. Phys. 21 077403
25Xu BDai Y MShen BXiao HYe Z RForget AColson DFeng D LWen H HHomes C CQiu X GLobo R P S M 2015 Phys. Rev. 92 104510
26Sekiba YSato TNakayama KTerashima KRichard PBowen J HDing HXu Y MLi L JCao G HXu Z ATakahashi T 2009 New J. Phys. 11 025020
27Dressel MGrüner G2002Electrodynamics of SolidsCambridgeCambridge University Press